10.07.2015 Views

Quantitative structural analyses and numerical modelling of ...

Quantitative structural analyses and numerical modelling of ...

Quantitative structural analyses and numerical modelling of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

“It strikes me that all our knowledge about the structure <strong>of</strong> our Earth is very muchlike what an old hen would know <strong>of</strong> the hundred-acre field in a corner <strong>of</strong> which she isscratching.”Charles Darwin


AcknowledgementsThis habilitation thesis is the result <strong>of</strong> several years <strong>of</strong> research, <strong>and</strong> many people have, inone way or another, influenced this work. First <strong>and</strong> foremost, I would like to express mydeepest thanks to Pr<strong>of</strong>. Karel Schulmann at whose department I have been studying<strong>and</strong> working since 1995, <strong>and</strong> whose support <strong>and</strong> encouragement in all phases <strong>of</strong> myacademic career have been invaluable to me. I have always pr<strong>of</strong>ited greatly from thestimulating discussions with Pr<strong>of</strong>. Schulmann, <strong>and</strong> his ideas, advice, <strong>and</strong> critical readinghave boosted the present work enormously.Secondly, sincere thanks go to Drs. P. Štípská, J. Konopásek, S. Ulrich, P. Jeřábek,J. Ježek, F. Hrouda <strong>and</strong> V. Janoušek, with whom I had, or have, the great pleasureto collaborate on the subjects presented in this thesis. I also thank my former Ph.D.students <strong>and</strong> collegues P. Hasalová, M. Racek, P. Závada, J. Franěk, L. Baratoux forperforming joint research with me on their interesting research topics.I feel very much enriched by all my visits in Strasbourg. I would like to thank J.-B.Edel, G. Manatschall, J. Lehmann, E. Skrzypek, F. Chopin, <strong>and</strong> all the other people atthe CGS/EOST University <strong>of</strong> Strasbourg, for the countless <strong>and</strong> motivative discussions.Important part <strong>of</strong> this thesis was written while I was a distinguished postdoctoral fellowat this institute.It was furthermore a great pleasure to organize an international conference Granulites &granulites 2009 on the field <strong>of</strong> research partly covered by this thesis, which was pursuedtogether with Pr<strong>of</strong>s. K. Schulmann, R. White, M. Brown, <strong>and</strong> P. O’Brien. Thanks goto all <strong>of</strong> them for their great enthusiasm, help <strong>and</strong> advice.I wish to express my gratitude to the directors, to the staff, <strong>and</strong> to all the other membersat the Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology in Prague for providing excellentworking conditions <strong>and</strong> a fantastic working atmosphere as well as to Mrs. M. Wontrobováfor her kind <strong>and</strong> efficient administrative support.So here I am at the end. On my personal treasures. I would never have achieved thishabilitation without the support <strong>of</strong> my wife Markéta <strong>and</strong> her infinite underst<strong>and</strong>ing formy staying late at work, disappearing for fieldwork, traveling on a conferences, beingabroad. . . And I would never be where I am without my family, <strong>and</strong> without my friends. . .v


Contentsviii4.7 Schulmann, Konopásek, Janoušek, Lexa, Lardeaux, Edel, Štípská, <strong>and</strong>Ulrich 2009b . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1594.8 Lexa, Schulmann, Janoušek, Štípská, Guy, <strong>and</strong> Racek 2011 . . . . . . . . 1814.9 Franěk, Schulmann, Lexa, Tomek, <strong>and</strong> Edel 2011a . . . . . . . . . . . . . 2054.10 Lexa, Štípská, Schulmann, Baratoux, <strong>and</strong> Kröner 2005 . . . . . . . . . . . 2314.11 Baratoux, Schulmann, Ulrich, <strong>and</strong> Lexa 2005b . . . . . . . . . . . . . . . . 2494.12 Závada, Schulmann, Konopásek, Ulrich, <strong>and</strong> Lexa 2007 . . . . . . . . . . . 2794.13 Schulmann, Martelat, Ulrich, Lexa, Štípská, <strong>and</strong> Becker 2008b . . . . . . 2954.14 Hasalová, Schulmann, Lexa, Štípská, Hrouda, Ulrich, Haloda, <strong>and</strong> Týcová2008a . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3154.15 Franěk, Schulmann, Lexa, Ulrich, Štípská, Haloda, <strong>and</strong> Týcová 2011b . . 341


List <strong>of</strong> Figures1.1 Strain evolution <strong>modelling</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . 61.2 Mechanical <strong>and</strong> rheological evolution <strong>of</strong> orthogneisses <strong>and</strong> migmatites . . 71.3 Parameter space for homogeneous transpression . . . . . . . . . . . . . . . 81.4 Schematic 3d structure <strong>of</strong> Gemer unit . . . . . . . . . . . . . . . . . . . . 101.5 Numerical model for continental indentation . . . . . . . . . . . . . . . . . 101.6 Oblique fold sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111.7 Schematic 3d structure <strong>of</strong> Jeseníky mountains . . . . . . . . . . . . . . . . 121.8 Fold analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121.9 AMS data <strong>of</strong> Třebíč massif . . . . . . . . . . . . . . . . . . . . . . . . . . 141.10 Numerical simulation <strong>of</strong> AMS fabric development . . . . . . . . . . . . . . 142.1 Schematic pr<strong>of</strong>ile across Central European Variscides . . . . . . . . . . . . 202.2 Kinematic model <strong>of</strong> synchronous folding <strong>and</strong> erosion . . . . . . . . . . . . 212.3 N-S pr<strong>of</strong>ile <strong>of</strong> eastern margin . . . . . . . . . . . . . . . . . . . . . . . . . 212.4 Numerical <strong>modelling</strong> <strong>of</strong> Rayleigh-Taylor instability within orogenic root . 222.5 Example <strong>of</strong> gravity inversion <strong>modelling</strong> . . . . . . . . . . . . . . . . . . . . 232.6 Model <strong>of</strong> thickening from above process . . . . . . . . . . . . . . . . . . . 242.7 Geodynamic model <strong>of</strong> crustal indentation . . . . . . . . . . . . . . . . . . 253.1 Crystal size distribution development . . . . . . . . . . . . . . . . . . . . . 313.2 Contrasting microstructures <strong>of</strong> amphibolites . . . . . . . . . . . . . . . . . 333.3 Melt pockets preferred orientation . . . . . . . . . . . . . . . . . . . . . . 343.4 Grain boundary statistics evolution . . . . . . . . . . . . . . . . . . . . . . 343.5 CSD in channel flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353.6 Example <strong>of</strong> digitized microstructures . . . . . . . . . . . . . . . . . . . . . 363.7 Coupled analysis <strong>of</strong> textures <strong>and</strong> microstructures . . . . . . . . . . . . . . 37ix


Dedicated to my wife Markéta, daughter Eliška <strong>and</strong> sonJáchym.... . .xi


ForewordThis habilitation thesis consists <strong>of</strong> three main parts, which aim to underline the principaldirections <strong>and</strong> features <strong>of</strong> my scientific career <strong>and</strong> gives the overview <strong>of</strong> new scientificdirections, which I would like to follow as an associate pr<strong>of</strong>essor <strong>of</strong> the Charles University.My research career is based on collaborative work <strong>of</strong> a larger research team, whichworked at the Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology from 1995 to 2011 <strong>and</strong> myresearch interest <strong>and</strong> focus was necessarily influenced by the overall aim <strong>of</strong> that group.This was particularly expressed in mathematical <strong>modelling</strong> <strong>of</strong> continental deformationin the micro-, meso - <strong>and</strong> macro-scale <strong>and</strong> the development <strong>of</strong> the necessary s<strong>of</strong>twaretools for specific geologically oriented quantitative studies, which were published underthe Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology mostly in collaboration with K. Schulmann,J. Konopásek, P. Jeřábek, P. Štípská, S. Ulrich <strong>and</strong> others, who initiate manysubjects further discussed in this text. My skills in computer programs development,computer-based processing <strong>and</strong> statistical evaluation <strong>of</strong> <strong>structural</strong> data, implementation<strong>of</strong> database <strong>and</strong> geographic information systems <strong>and</strong> ability to model processes <strong>numerical</strong>lyallowed me substantially contribute to several manuscripts <strong>of</strong> different authors thatare cited in this habilitation thesis. As these techniques were progressively more <strong>and</strong>more used by IPSG research team, it became the driving force for three major researchdirections <strong>of</strong> my career presented here. These are following themes: 1) quantitativeanalysis <strong>of</strong> deformation structures <strong>and</strong> <strong>numerical</strong> <strong>modelling</strong> <strong>of</strong> deformation,2) mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications<strong>and</strong> 3) quantitative analysis <strong>of</strong> metamorphic microstructures, their visualization<strong>and</strong> statistical processing. At this point I would like to stress that I highlyappreciated multidisciplinary approach <strong>of</strong> collective scientific collaboration, which ruledin the last ten years <strong>of</strong> my pr<strong>of</strong>essional career in the Institute <strong>of</strong> Petrology <strong>and</strong> StructuralGeology <strong>and</strong> that helped me to identify the essentials <strong>of</strong> my personal approach, as wellas the pr<strong>of</strong>essional qualities <strong>of</strong> my colleagues. It is my aim further transfer this spirit <strong>of</strong>active cooperation <strong>and</strong> stimulating discussions to students <strong>and</strong> cultivate this approachat Charles University in the future.1


Foreword 2First topic “<strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> <strong>numerical</strong> <strong>modelling</strong><strong>of</strong> deformation” includes mainly <strong>numerical</strong> <strong>modelling</strong> <strong>of</strong> progressive development <strong>of</strong> finitestrain <strong>and</strong> superposition <strong>of</strong> deformation events in collision zones. This part <strong>of</strong> myresearch was originally related to the simple kinematic <strong>modelling</strong> <strong>of</strong> deformation pathsin the Flinn space (Konopásek et al., 2001, 2003) <strong>and</strong> <strong>modelling</strong> <strong>of</strong> finite strain withintranspressional zones (Schulmann et al., 2003). Later, I focused on a “thin-sheet” <strong>numerical</strong><strong>modelling</strong> <strong>of</strong> deformation in collision zones characterized by complex geometries<strong>and</strong> variable boundary conditions (Lexa et al., 2003). Part <strong>of</strong> my research activities wasdevoted to quantitative geometrical evaluation <strong>of</strong> so-called “extensional shear b<strong>and</strong>s”,which are commonly used as tectonic transport indicators (Lexa et al., 2004) <strong>and</strong> possibleuse <strong>of</strong> fold shape <strong>analyses</strong> as a proxy <strong>of</strong> degree <strong>of</strong> mechanical anisotropy <strong>and</strong> bulkdeformation intensity (Baratoux et al., 2005a). Recently, my research activities areoriented towards underst<strong>and</strong>ing <strong>of</strong> deformation overprints in magmatic <strong>and</strong> solid staterocks, which are based on my early research (Konopásek et al., 2001) <strong>and</strong> whose resultsappear in some recent works <strong>of</strong> my co-workers (e.g. Kratinová et al., 2010, Lehmannet al., 2011, Schulmann et al., 2009b). This topic is crucial in underst<strong>and</strong>ing the symmetry<strong>and</strong> intensity <strong>of</strong> deformation in orogenic zones <strong>and</strong> it will be discussed in thechapter “Perspectives <strong>of</strong> <strong>numerical</strong> <strong>modelling</strong> <strong>of</strong> deformation in geology”. In all studies,where I am not listed as first author I have decisively contributed to the results by means<strong>of</strong> creating s<strong>of</strong>tware, own contribution in formulating the manuscript as well as field researchincluding the West Carpathians, Bohemian Massif, Central Asian Orogenic Belt<strong>and</strong> the Vosges Mountains.The second topic “Mechanism <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanicalimplications” is based on field study <strong>of</strong> two areas: the eastern margin <strong>of</strong> the BohemianMassif <strong>and</strong> central Western Carpathians. Research related to channel flow model alongthe eastern margin <strong>of</strong> the Bohemian Massif is the subject <strong>of</strong> numerous publicationswhich I co-authored with my colleagues <strong>and</strong> students, for example, Schulmann et al.(2005), Racek et al. (2006) <strong>and</strong> Franěk et al. (2006). Our research gained new stimulusby recently published description <strong>of</strong> kinematics <strong>of</strong> the channel flow model (Schulmannet al., 2008a), where the channel geometry, thermal structure <strong>and</strong> polyphase nature <strong>of</strong><strong>structural</strong> evolution were firstly identified within Variscan orogenic system. New developmentsbased exclusively on my own research are further discussed in “Perspectivesfor the thermomechanical <strong>modelling</strong> <strong>of</strong> channel flow in the orogeny”, where I illustratethe plausibility <strong>of</strong> our model in terms <strong>of</strong> <strong>numerical</strong> <strong>modelling</strong> for different boundaryconditions. However, there is a second type <strong>of</strong> channel flow, which is linked to influx <strong>of</strong>crustal material into the orogenic root, which is subsequently extruded during exhumationstage. This model was firstly proposed on the example <strong>of</strong> the Western Carpathians,where some aspects <strong>of</strong> the lower crustal flow were <strong>numerical</strong>ly simulated (Jeřábek et al.,


Foreword 32008, 2007) <strong>and</strong> subsequently applied to describe evolution <strong>of</strong> orogenic root in BohemianMassif. Here we used available geochemical <strong>and</strong> geochronological data to argue that thefelsic lower crust is an allochthonous body emplaced underneath pre-existing mafic lowercrust during continental subduction (Schulmann et al., 2009b). Similar scenario was recentlyproposed for evolution Orlica-Śnieżnik dome in Western Sudetes (Chopin et al.,2011a). It appears that influx <strong>of</strong> crustal material to the orogenic root zone have a majorimpact on his future thermal <strong>and</strong> rheological evolution. Detailed study <strong>of</strong> geodynamicconsequences caused by subduction <strong>of</strong> felsic <strong>and</strong> radioactively productive material underthe orogenic system is described in Lexa et al. (2011) <strong>and</strong> this aspect will be furtherelaborated in the chapter “Perspectives for the thermomechanical <strong>modelling</strong> <strong>of</strong> channelflow in the orogeny”.The last topic “<strong>Quantitative</strong> analysis <strong>of</strong> metamorphic microstructures, its visualization<strong>and</strong> statistical description” introduce a comprehensive set <strong>of</strong> quantitative methodsfor analysis <strong>of</strong> deformation microstructures <strong>and</strong> it is based on PolyLX s<strong>of</strong>tware developedby myself. This chapter will deal with techniques <strong>of</strong> microstructure digitalization <strong>and</strong>subsequent statistical processing, including evaluation <strong>of</strong> the grain size, grain shapes <strong>and</strong>statistics <strong>of</strong> their mutual contacts. The PolyLX covers several methods <strong>of</strong> quantification<strong>of</strong> preferred orientation <strong>of</strong> grains <strong>and</strong> boundaries as well as their statistical processing.This method allows us not only visualize microstructure characteristics <strong>and</strong> their statisticaltreatment, but also address some kinetic problems <strong>of</strong> metamorphic microstructures.This method was originally described in my Ph.D. thesis <strong>and</strong> firstly applied in the work<strong>of</strong> Lexa et al. (2005). Subsequently, the PolyLX s<strong>of</strong>tware was successfully used in numerouspublications <strong>of</strong> our research group (Baratoux et al., 2005a,b, Hasalová et al., 2008a,Jeřábek et al., 2007, Kratinová et al., 2010, Lexa et al., 2005, Machek et al., 2007, Schulmannet al., 2008b, Závada et al., 2007, 2009). In addition, there is a number <strong>of</strong> laboratoriesin France, where I worked during my post-doctoral fellowship <strong>and</strong> I led specializedcourses in this subject, which are routinely using PolyLX s<strong>of</strong>tware in complex micro<strong>structural</strong>problems (Barraud, 2006, Chopin et al., 2011b, Martelat et al., 2011, Oliot et al.,2011). It appears, that PolyLX becomes valuable tool for projects, where quantification<strong>of</strong> the spatial relationships <strong>of</strong> minerals is needed. Recently I leed three specialized workshops(Granulites & Granulites 2009, Prague; 2010 Tromsø; Geomaterials 2011, Prague)focused on quantitative microstructure <strong>analyses</strong>. Recently, PolyLX has interfaces toexchange data with s<strong>of</strong>twares like ELLE (http://www.materialsknowledge.org/elle) orMTEX (http://code.google.com/p/mtex) that allows simulating the evolution <strong>of</strong> deformationmicrostructures or quantitative textural <strong>analyses</strong> <strong>of</strong> natural samples. Futherideas as some <strong>of</strong> the new kinetic <strong>and</strong> petrogenetic outputs that could be depicted bythis method (Franěk et al., 2011b) are given in the chapter “Perspectives for <strong>numerical</strong><strong>modelling</strong> <strong>of</strong> deformation microstructures <strong>and</strong> their quantification”. As PolyLX is open


Foreword 4source <strong>and</strong> freely available on the internet (http://petrol.natur.cuni.cz/˜ondro) we hopethat it will become a widely used method in petrology <strong>and</strong> <strong>structural</strong> geology.Ondrej Lexa: lexa@natur.cuni.cz


Chapter 1<strong>Quantitative</strong> analysis <strong>of</strong>deformation structures <strong>and</strong> their<strong>numerical</strong> <strong>modelling</strong>Finite deformation <strong>modelling</strong> is an approach that is used in <strong>structural</strong> geology for decadesas a proxy to study distribution <strong>and</strong> nature <strong>of</strong> deformation within crustal collision zones,zones <strong>of</strong> transpression or transtension <strong>and</strong> lithospheric extension (S<strong>and</strong>erson <strong>and</strong> Marchini,1984). The main question arises in situations, where rock masses do not behaveaccording to “simple rules” <strong>of</strong> homogeneous deformation, i.e. where considerablevariation in lithology <strong>and</strong> rheology introduce heterogeneities, which are likely to controltheir complex deformation behaviour accompanied with strain superposition (Burg,1999). Konopásek et al. (2001) described in western part <strong>of</strong> Erzgebirge steeply dippinglineations contained by steeply dipping eclogite foliations <strong>and</strong> horizontal lineations insurrounding <strong>structural</strong>ly conformable orthogneisses. This conflict has led us to modelevolution <strong>of</strong> deformation pattern within “s<strong>of</strong>t” incompetent orthogneiss during verticalflattening. In our model, these orthogneisses initially showed a similar orientation <strong>of</strong>finite strain axes as eclogites but after subsequent vertical shortening suffered viscousdeformations, which does not affected competent eclogites. Our <strong>numerical</strong> simulationshave shown conclusively that the superposition <strong>of</strong> deformation stages is a powerful mechanismthat allows us to underst<strong>and</strong> some strain paradoxes as one described above. Theresult <strong>of</strong> our <strong>modelling</strong> (Fig. 1.1) shows that the contrast in deformation record <strong>of</strong> eclogites<strong>and</strong> orthogneisses results from common deformation history, which is selectively <strong>and</strong>to different extent recorded in individual lithologies.Our results were challenged by German colleagues Krohe <strong>and</strong> Willner (2003) <strong>and</strong> wewere obliged to defend our hypothesis in the manuscript Konopásek et al. (2003) using5


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 6Figure 1.1: Calculated evolution <strong>of</strong> the finite strain as a result <strong>of</strong> homogeneous verticalshortening <strong>of</strong> previous fabric ellipsoid marked by vertical X-axis. Black <strong>and</strong> grey linesshow the progressive strain being labelled by elongation <strong>of</strong> former X-axis. Results <strong>of</strong>calculations is shown for deformation fabrics marked by contrasting strain symmetries<strong>and</strong> intensities as observed in the field. a) low strain domain with plane strain symmetryb) high strain domain with prolate shape.our <strong>numerical</strong> <strong>modelling</strong> as a backbone <strong>of</strong> the discussion. More recently, the problem<strong>of</strong> strain superimposition was elaborated again in the study <strong>of</strong> orthogneiss migmatitesdeformation in the French Vosges (Schulmann et al., 2009a). Here, the originally verticalstructures <strong>of</strong> orthogneisses embedded by metasedimentary migmatites are overprintedby vertical shortening due to larger scale crustal extension, which resulted in development<strong>of</strong> prolate strain fabrics in the so-called “internal margin” <strong>and</strong> oblate strainfabrics in the “external margin” <strong>of</strong> the crustal scale orthogneiss inclusion. Our detailed<strong>structural</strong> <strong>and</strong> AMS study revealed that these changes correlate well with melt fraction<strong>and</strong> orientation <strong>of</strong> mechanical anisotropy relative to stress axes applied to orthogneissbody by flow <strong>of</strong> surrounding migmatites (Fig. 1.2). It turns out that in order to explainprolate finite strain ellipsoids within orogenic zones, one must account to the model <strong>of</strong>


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 7Figure 1.2: Three stages <strong>of</strong> the mechanical <strong>and</strong> rheological evolution <strong>of</strong> orthogneissbodies <strong>and</strong> host metasedimentary migmatites: (a) Early stage <strong>of</strong> layer parallel shortening.(b) Later stage in the deformation when the margin <strong>of</strong> the orthogneiss body hasbeen transformed to an orthogneiss/granite multilayer. (c) In the final stage furtherhomogeneous deformation <strong>of</strong> host diatexites is accompanied by flattening <strong>and</strong> intensification<strong>of</strong> the orthogneiss fabric.deformation overprints, otherwise we are exposed to the problem <strong>of</strong> strain compatibility(formation <strong>of</strong> open spaces within rock masses is not possible for obvious reasons). In thiswork, <strong>numerical</strong> models were set-up to reproduce the variation in the finite strain <strong>and</strong>orientation <strong>of</strong> the original mechanical anisotropy with respect to external stress field.Although the model successfully explained the <strong>structural</strong> zonality <strong>of</strong> rigid domains inthe migmatites, it was not robust enough to underst<strong>and</strong> general dependencies <strong>of</strong> thefinite strain intensity <strong>and</strong> symmetry to the degree <strong>and</strong> symmetry <strong>of</strong> initial mechanicalanisotropy. In the following I will show some other perspectives <strong>of</strong> the problem.Relative motion <strong>of</strong> lithospheric plates on a spherical surface is such that the plateconvergence vectors are <strong>of</strong>ten not orthogonal to plate boundaries (Dewey, 1975, McKenzie<strong>and</strong> Parker, 1967). These plate boundaries experience combined transcurrent <strong>and</strong>convergent displacements associated with development <strong>of</strong> deformation zones <strong>of</strong> differentsize. Within continental blocks, the deformation is not only restricted to active plate


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 8Figure 1.3: Superposition <strong>of</strong> strain intensity (D), strain symmetry (K), <strong>and</strong> sampleelevation for various values <strong>of</strong> rigid floor depth <strong>and</strong> initial sample depth z 0 , plotted interms <strong>of</strong> angle <strong>of</strong> convergence (α) <strong>and</strong> time parameter (k t ). The values <strong>of</strong> D show astrong dependence on ratio between convergence velocity <strong>and</strong> width <strong>of</strong> deformed zone(R vd ) <strong>and</strong> time for α = 20 ◦ − 90 ◦ ; for α < 20 ◦ the dependence <strong>of</strong> D on R vd increases.K is strongly dependent on a but not much on R vd <strong>and</strong> time.boundaries but occurs within zones <strong>of</strong> weakness inside rigid continental domains <strong>and</strong>can be approximately described as a deformation <strong>of</strong> a weak tabular zone bounded byrigid blocks with steep parallel walls. All the mentioned types <strong>of</strong> deformation zonescan be more or less described by a model called “transpression”, introduced first byHarl<strong>and</strong> (1971), developed by S<strong>and</strong>erson <strong>and</strong> Marchini (1984), <strong>and</strong> elaborated by manyothers. In the paper Schulmann et al. (2003), we summarize our experience with thisissue <strong>and</strong> we have modelled strain evolution in such zones including various deformationpartitioning schemes, namely 1) discrete partitioning (lateral component <strong>of</strong> motion isaccommodated by discrete faults), 2) ductile partitioning (lateral component <strong>of</strong> movementis accommodated by narrow zones <strong>of</strong> ductile flow) <strong>and</strong> 3) viscous partitioning(deformation zone is divided to domains where deformation is accumulated at differentstrain rate). In this model, we also included the concept <strong>of</strong> a rigid floor, which controls


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 9elevation <strong>and</strong> exhumation <strong>of</strong> rocks. The results <strong>of</strong> our model (Fig. 1.3) shows systematicvariability in strain pattern, depending on the convergence angle <strong>and</strong> strain-rate withinthe transpressive zones. This fundamental behaviour could be used to inverse field datainto controlling parameters in both active <strong>and</strong> fossil deformation zone.One <strong>of</strong> the most important published contributions to this topic (Lexa et al., 2003)is based on our detailed <strong>structural</strong> analysis <strong>of</strong> polyphase deformation within the Gemerunit in the Western Carpathians. Here, the Paleozoic rocks deformed during LowerCretaceous convergence form steep cleavage fan between the more rigid blocks <strong>of</strong> eastern<strong>and</strong> western part <strong>of</strong> the Vepor crystalline basement. Moreover, the so-called Trans-Gemeric shear zone is developed along eastern margin <strong>of</strong> the Vepor promontory <strong>and</strong>further east separate Gemer unit into two domains. By contrast, the eastern part <strong>of</strong>the Vepor basement suffered more frontal convergence with Gemer unit documented bydevelopment <strong>of</strong> Eastern Gemer thrust (Fig. 1.4). We tried to justify this complicatedtectonic model using viscous sheet formulation modelled by finite element method, wherewe simulate deformation in front <strong>of</strong> the moving indenter, which progressively deformedviscous domain in between the stationary blocks <strong>of</strong> eastern <strong>and</strong> western Vepor basement<strong>and</strong> which was free to thicken <strong>and</strong> outflow among these blocks.The result <strong>of</strong> <strong>numerical</strong> <strong>modelling</strong> (Fig. 1.5) was more than surprising, since both thesymmetry <strong>and</strong> distribution <strong>of</strong> viscous deformation <strong>of</strong> thin viscous sheet corresponds withsufficient precision to our observations. We consider this model as particularly valuablecontribution to the simulation <strong>of</strong> continental deformation. This example suggests thatthe concept developed since the eighties <strong>of</strong> last century (Engl<strong>and</strong> <strong>and</strong> Houseman, 1986,Houseman <strong>and</strong> Engl<strong>and</strong>, 1986) for deformation lithospheric scale deformations (deformation<strong>of</strong> Tibet in front <strong>of</strong> India indenter) could be successfully applied to significantlysmaller scale <strong>of</strong> geological units.Our simulations showed that our model for transpression was satisfying startingestimate for the deformation evolution in areas with complex kinematic framework. Webelieve that the deformation <strong>of</strong> continental crust is dominantly controlled by the presence<strong>of</strong> lateral inhomogeneities <strong>of</strong> different shape <strong>and</strong> that the deformation is subsequentlypartitioned <strong>and</strong> channelized into narrow zones. Underst<strong>and</strong>ing polyphase deformationas consequence <strong>of</strong> principles <strong>of</strong> continuum mechanics is a challenge that will be testedin the near future theoretically as well as on field examples.Our field studies <strong>and</strong> observations rise the question <strong>of</strong> underst<strong>and</strong>ing the flow kinematicsin anisotropic rocks. The most common kinematic indicators used by field geologistsare “extensional shear b<strong>and</strong>s”, which form basis <strong>of</strong> argumentation in number<strong>of</strong> kinematic orogenic models (e.g. Behrmann, 1988, Neubauer et al., 1999, Platt <strong>and</strong>


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 10Figure 1.4: Geometry <strong>of</strong> main tectonic units <strong>of</strong> Central Western Carpathians<strong>and</strong> schematic evolution <strong>of</strong> major structures during progressive deformation used toparametrize subsequent <strong>numerical</strong> model.Figure 1.5: Finite strain pattern developed in Gemer unit after 5 Myr <strong>of</strong> shortening.(a) Distribution <strong>of</strong> strain intensity (D) <strong>and</strong> orientation <strong>of</strong> finite strain ellipsoid. (b)Distribution <strong>of</strong> probability <strong>of</strong> possible simple shear reactivation <strong>of</strong> evolved anisotropy.(c) Distribution <strong>of</strong> finite strain symmetry (K). (d) Distribution <strong>of</strong> finite topography (inmeters) in front <strong>of</strong> the indenter.


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 11Figure 1.6: Stereographic projection showing influence <strong>of</strong> inter-limb angle <strong>and</strong> anglebetween fold axial plane <strong>and</strong> main anisotropy. The range <strong>of</strong> section planes showingshear-b<strong>and</strong> geometry increases with inter-limb angle <strong>of</strong> folds affecting main anisotropy.Vissers, 1989). Some field observations, however, were cautionary: Are geologists alwayssure about mutual orientation <strong>of</strong> observed plane <strong>and</strong> kinematic frame? Are all the“extensional shear b<strong>and</strong>s” really extensional structures? I tried to answer these questionsin a simple fold section model (Lexa et al., 2004) which showed that from all thepossible sections <strong>of</strong> 3D fold structure, only a small percentage allows us to observe theactual fold shape. In fact, in most <strong>of</strong> the sections, asymmetric fold structure appears as“extensional shear b<strong>and</strong>s”. Our own observation from the central Western Carpathians<strong>and</strong> from eastern margin <strong>of</strong> Bohemian Massif revealed that most <strong>of</strong> “extensional shearb<strong>and</strong>s” represent only the oblique sections across compressive folds (Fig. 1.6).This important observation seriously modifies the today view on origin or core complexes,which can be only the oblique sections across the crustal scale folds. For instance,the discussion <strong>of</strong> origin <strong>of</strong> Montagne Noire gneiss dome as a core complex or steep crustalmegafold is debated for more than twenty years by French geologists (e.g. Aerden, 1998,Matte et al., 1998, Van den Driessche <strong>and</strong> Brun, 1992). We believe that the analysis<strong>of</strong> kinematics <strong>of</strong> such orogenic systems requires a thorough revision in the light <strong>of</strong> theseobservations. Similar conclusion could be made about the origin <strong>of</strong> Tauren Window as itis shown by contrasting interpretations (e.g. Ratschbacher et al., 1989, Rosenberg et al.,November, 2004).


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 12Figure 1.7: 3d blockdiagram <strong>of</strong> general structure <strong>of</strong> northern termination <strong>of</strong> Silesi<strong>and</strong>omain. Figure shows progressive increase <strong>of</strong> late deformation <strong>and</strong> folding superimposedon previous flat fabric in conjunction with increase in metamorphic grade toward west.Figure 1.8: Selected fold assemblages from the garnet, staurolite <strong>and</strong> sillimanite zones,respectively. (a) Dip isogon patterns. (b) b1 (fold shape) vs. F (degree <strong>of</strong> fold flattening)plots constructed for the presented folds showing the relative importance <strong>of</strong> foldflattening <strong>and</strong> active fold amplification.


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 13Some models <strong>of</strong> viscous deformation <strong>of</strong> thin plates show the characteristic deformationgradient, which is generally difficult to identify. Are we able to recognise sucha gradient within anisotropic rock masses? And what is the effect <strong>of</strong> lateral temperaturegradient on deformation <strong>of</strong> the crust? We have tried to answer those interestingquestions in the paper Baratoux et al. (2005b), where we observed <strong>structural</strong> variationsin shape <strong>and</strong> flattening <strong>of</strong> mesoscopic folds across 40 km long section (Fig. 1.7). Thispr<strong>of</strong>ile was characterized by buckle <strong>and</strong> chevron folds in the east, flattened buckle foldsin the centre <strong>and</strong> flow folds in the west. <strong>Quantitative</strong> analysis <strong>of</strong> fold shape in terms<strong>of</strong> Fourier harmonic analysis method <strong>and</strong> the Ramsay’s method <strong>of</strong> dip isogons pointedto systematic decay in degree <strong>of</strong> mechanical anisotropy from east to west. This trendwas correlated with PolyLX <strong>analyses</strong> <strong>of</strong> microstructures <strong>of</strong> amphibolites, which revealexistence <strong>of</strong> gradient in shape anisotropy on grain scale. This detailed quantitative workconcludes that the character <strong>of</strong> folding is directly controlled by degree <strong>of</strong> anisotropy <strong>of</strong>microstructure, so that the grain shape anisotropy governed mechanical properties <strong>of</strong>folded macroscopic layers. In another words, lateral variations in fold style correlatewith lateral variations in the mechanical anisotropy which reflect metamorphic zonalityi.e. thermal history <strong>of</strong> previous tectonometamorphic event.1.1 Perspectives <strong>of</strong> <strong>numerical</strong> <strong>modelling</strong> <strong>of</strong> deformation ingeologyUse <strong>of</strong> <strong>numerical</strong> <strong>modelling</strong> in <strong>structural</strong> geology introduced new possibilities especiallyin the area <strong>of</strong> deformation overprints. The range <strong>of</strong> deformation structures, which aretraditionally explained as the result <strong>of</strong> “magic” deformation partitioning into zones <strong>of</strong>pure shear (prolate <strong>and</strong> oblate) <strong>and</strong> simple shear dominated deformation can not explainfabric variations encountered in orogens. This concerns the problem <strong>of</strong> strain compatibility<strong>and</strong> second the problem <strong>of</strong> significance <strong>of</strong> deformation intensities. Therefore wedecided to elaborate methods to quantify <strong>and</strong> <strong>numerical</strong>ly simulate these deformationprocesses (e.g. Kratinová et al., 2010). For example, long-st<strong>and</strong>ing study <strong>of</strong> Třebíč syenitemassif led us to conclude is that the magma body suffered multiple deformationsduring magmatic stage, which led to the distinctive variance <strong>of</strong> the magmatic fabricsthat are not explainable in terms <strong>of</strong> st<strong>and</strong>ard approaches (Lexa et al., in prep.). OurAMS study shows that northern part <strong>of</strong> the body exhibit NW-SE trending lineation accompaniedwith oblate shape <strong>of</strong> AMS ellipsoid, while the southern part is characterizedby NE-SW trending lineation <strong>and</strong> prolate shape <strong>of</strong> AMS ellipsoid (Fig. 1.9).Our <strong>numerical</strong> <strong>modelling</strong> revealed that the variance in the symmetry <strong>and</strong> direction <strong>of</strong>lineations could be successfully explained by the superposition <strong>of</strong> horizontal flow (within


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 14Figure 1.9: Map <strong>of</strong> P <strong>and</strong> T parameters <strong>of</strong> AMS fabrics with selected homogeneousAMS data showing two domains within Třebíč massif.Figure 1.10: Numerical simulation <strong>of</strong> progressive deformation <strong>of</strong> uniformly dispersedbiotite grains for horizontal shortening (green frame) superimposed by vertical shortening<strong>and</strong> perpendicular elongation (orange frame). Strain paths in PT space show highvariability <strong>of</strong> AMS fabric shape (T) for similar values <strong>of</strong> degree <strong>of</strong> AMS (P).


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 15crustal channel) on older vertical magmatic fabrics that are folded <strong>and</strong> therefore rotatedduring early stages <strong>of</strong> channel flow. This <strong>numerical</strong> model shown that routine utilisation<strong>of</strong> methods <strong>of</strong> AMS fabric interpretation could easily fail to correctly interpret plutonsemplacement histories. It appears that the superposition <strong>of</strong> two distinct plane straindeformations with various mutual geometrical relations could produce broad spectrum<strong>of</strong> finite deformation symmetries as well as finite deformation intensities which are commonlyobserved in real rocks (Fig. 1.10). Recently we also investigated intrusions <strong>of</strong>magmatic sheets related to exhumation <strong>of</strong> crustal scale gneiss dome Lehmann et al.(2011). The magmatic sheets are emplaced along boundary between core <strong>of</strong> gneiss dome<strong>and</strong> its mantle in contrasting tectonic regimes. Our <strong>modelling</strong> showing that systematicfabric transitions are related to superpositions <strong>of</strong> strain in magmatic state. It should bepointed out that without these experiences <strong>and</strong> deep underst<strong>and</strong>ing <strong>of</strong> strain accumulationin magmatic rocks, geologists can easily commit wrong geological models.


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 161.2 Accompanying publicationsKonopásek, J., Schulmann, K., Lexa, O., 2001. Structural evolution <strong>of</strong> the centralpart <strong>of</strong> the Krušné Hory (Erzgebirge) Mountains in the Czech Republic; evidencefor changing stress regime during Variscan compression. Journal <strong>of</strong> Structural Geology23 (9), 1373–1392. 49Schulmann, K., Thompson, A. B., Lexa, O., Ježek, J., 2003. Strain distribution<strong>and</strong> fabric development modeled in active <strong>and</strong> ancient transpressive zones. Journal<strong>of</strong> Geophysical Research, B, Solid Earth <strong>and</strong> Planets 108(B1), 2023, doi:10.1029/2001JB000632. 69Lexa, O., Schulmann, K., Ježek, J., 2003. Cretaceous collision <strong>and</strong> indentationin the Western Carpathians: View based on <strong>structural</strong> analysis <strong>and</strong> <strong>numerical</strong>modeling. Tectonics 22 (6), 1066, doi:10.1029/2002TC001472. 85Lexa, O., Cosgrove, J., Schulmann, K., 2004. Apparent shear-b<strong>and</strong> geometry resultingfrom oblique fold sections. Journal <strong>of</strong> Structural Geology 26 (1), 155–161.101Baratoux, L., Lexa, O., Cosgrove, J. W., Schulmann, K., 2005a. The quantitativelink between fold geometry, mineral fabric <strong>and</strong> mechanical anisotropy; as exemplifiedby the deformation <strong>of</strong> amphibolites across a regional metamorphic gradient.Journal <strong>of</strong> Structural Geology 27 (4), 707–730. 1091.3 Related publicationsKonopásek, J., Schulmann, K., Lexa, O., 2003. Reply to comments by A. Krohe <strong>and</strong>A.P. Willner on ”Structural evolution <strong>of</strong> the central part <strong>of</strong> the Kru né Hory (Erzgebirge)Mountains in the Czech Republic—evidence for changing stress regime duringVariscan compression”. Journal <strong>of</strong> Structural Geology 25 (6), 1005–1007. Schulmann, K., Edel, J.-B., Hasalová, P., Cosgrove, J., Ježek, J., Lexa, O., 2009a.Influence <strong>of</strong> melt induced mechanical anisotropy on the magnetic fabrics <strong>and</strong> rheology<strong>of</strong> deforming migmatites, Central Vosges, France. Journal <strong>of</strong> Structural Geology31 (10), 1223–1237.Lehmann, J., Schulmann, K., Edel, J.-B., Hrouda, F., Lexa, O., Štípská, P., Haloda,J., 2011. AMS <strong>and</strong> <strong>structural</strong> record <strong>of</strong> granitoid sills emplacement during growth<strong>of</strong> continental gneiss dome: implications for exhumation <strong>of</strong> deep crust. Journal <strong>of</strong>Structural Geology, submitted.


1. <strong>Quantitative</strong> analysis <strong>of</strong> deformation structures <strong>and</strong> their <strong>numerical</strong> <strong>modelling</strong> 17 Kratinová, Z., Ježek, J., Schulmann, K., Hrouda, F., Shail, R. K., Lexa, O., 2010.Noncoaxial K-feldspar <strong>and</strong> AMS subfabrics in the L<strong>and</strong>’s End granite, Cornwall:Evidence <strong>of</strong> magmatic fabric decoupling during late deformation <strong>and</strong> matrix crystallization.Journal <strong>of</strong> Geophysical Research, B, Solid Earth <strong>and</strong> Planets 115,B09104, doi:10.1029/2009JB006714.


Chapter 2Mechanisms <strong>of</strong> lower crustal flow<strong>and</strong> its thermal <strong>and</strong> mechanicalimplicationsMechanisms <strong>of</strong> lower crustal flow are reflected in two different crustal levels linked bothto the processes <strong>of</strong> orogenic root building <strong>and</strong> exhumation <strong>of</strong> the orogenic lower crust.These processes have been evaluated on the basis <strong>of</strong> our own <strong>structural</strong>, petrological<strong>and</strong> geochronological studies along the eastern margin <strong>and</strong> from the central part <strong>of</strong> theBohemian Massif (Franěk et al., 2006, Racek et al., 2006, Schulmann et al., 2005). Theprinciple <strong>of</strong> this model is the large-scale folding <strong>of</strong> the bottom crust as a response tolateral forcing. Amplification <strong>of</strong> folds gradually passed to vertical extrusion, which wassubsequently transferred to horizontal flow just below the brittle-ductile transition. Thehorizontal flow is driven by two processes: 1) collapse <strong>of</strong> vertical extrusion related fabricsdue to excess <strong>of</strong> gravity potential energy <strong>of</strong> elevated orogenic lid <strong>and</strong> 2) horizontalsubsurface flow driven by indentation <strong>of</strong> Brunia basement generating hot <strong>and</strong> heterogeneousfold-nappe (length <strong>of</strong> 200 km). The initial folding <strong>and</strong> extrusion <strong>of</strong> lower crustis probably controlled by the ongoing subduction <strong>of</strong> Saxothuringian continental lithosphere(Schulmann et al., 2009b) while the horizontal channel flow is possibly controlledby change in lithospheric plates configuration (Fig. 2.1). The whole concept is clearlyexplained in the work <strong>of</strong> Schulmann et al. (2008a), where we report a full geochronological,<strong>structural</strong> <strong>and</strong> petrological inventory supporting the model <strong>of</strong> vertical extrusion<strong>and</strong> subsequent subsurface horizontal flow processes.The subsurface horizontal flow or horizontal spreading <strong>of</strong> exhumed orogenic lowercrust are generally interpreted as a result <strong>of</strong> ductile rebound <strong>of</strong> buoyant crustal root19


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 20Figure 2.1: Model <strong>of</strong> laterally forced overturns related to Saxothuringian convergencein Bohemian Massif.(Koyi et al., 1999). In detail the process was described by Rey et al. (2001) or V<strong>and</strong>erhaeghe<strong>and</strong> Teyssier (2001) who discuss possible <strong>structural</strong> levels <strong>of</strong> horizontal flowresulting from rheological stratification <strong>and</strong> boundary conditions <strong>of</strong> their models. Thistype <strong>of</strong> orogenic collapse was recognized in two specific areas <strong>of</strong> Bohemian Massif: thesouth Bohemian granulites (Franěk et al., 2006, 2011a) <strong>and</strong> the Lugian domain <strong>of</strong> WesternSudetes (Dumicz, 1979, Štípská et al., 2004). In our new studies we have describedmetamorphic evolution related to burial <strong>of</strong> orogenic middle crust contemporaneous withextrusion <strong>of</strong> lower crustal omphacite bearing granulites (Skrzypek et al., 2011b). Subsequentsubhorizontal flow led to exhumation <strong>of</strong> the crust from 30 km depth to about 10km during which horizontal foliation associated with sequential growth <strong>of</strong> cordierite <strong>and</strong><strong>and</strong>alusite originated. This particular <strong>structural</strong> evolution resembles ductile thinningmechanism described in subduction wedges (Ring et al 2009, Br<strong>and</strong>on <strong>and</strong> Fe,...XXX).In order to image this process I have modelled the vertical exchange <strong>of</strong> crustal materialas laterally forced gravity overturns followed by ductile thinning <strong>of</strong> crust bellow thebrittle-ductile transition (Fig. 2.2). The thermal field <strong>and</strong> P-T-t paths were simulatedfor both exhumation mechanisms in order to correlate petrological <strong>and</strong> geochronologicaldata with <strong>modelling</strong> results. We shown that the horizontal spreading is a localised processoperating in several kilometres wide layer bellow orogenic lid which is progressivelydestroyed during ductile thinning process (Skrzypek et al., 2011b, Štípská et al., 2011).However, different type <strong>of</strong> large-scale horizontal flow <strong>of</strong> hot lower crustal materialoperates at the eastern margin <strong>of</strong> Bohemian Massif (Fig. 2.3). Here, the hot orogeniclower crust is thrust over the continental basement which is documented by gravityinversion <strong>modelling</strong> (Guy et al., 2011). In this work we have shown that the crustalbasement unit is covered by thin layer <strong>of</strong> low density migmatitic material (Fig. 2.5). Inaddition, new petrological studies <strong>of</strong> Hasalová et al. (2008a,b) <strong>and</strong> Štípská et al. (2008)show that the flow <strong>of</strong> migmatites occurred above the basement at the depth <strong>of</strong> 25 kmcorresponding to 7 - 8 kbar pressures. We argue, that the migmatitic rocks are flowingabove the basement in form <strong>of</strong> ductile hot nappe which is progressively fed by lower


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 21Figure 2.2: Schematic evolution <strong>of</strong> vertical displacement <strong>of</strong> folded single-layer duringbulk shortening for (a) “no erosion”, (b) “increasing erosion” <strong>and</strong> (c) “full erosion”models.Figure 2.3: Schematic geological cross-section showing main features <strong>of</strong> continentalchannel flow along eastern margin <strong>of</strong> Bohemian Massif.crustal material as the basement promontory is progressively wedging the thickenedorogenic root. We discuss the characteristics <strong>of</strong> the deformation fabrics in the hot nappe<strong>and</strong> defined the metamorphic <strong>and</strong> <strong>structural</strong> pattern as a result <strong>of</strong> the crustal scalechannel flow (Schulmann et al., 2008a).I have performed series <strong>of</strong> <strong>numerical</strong> models in order to answer the questions <strong>of</strong> theheat source origin, evolution <strong>of</strong> transitional rheology <strong>of</strong> the lithosphere <strong>and</strong> existence


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 22Figure 2.4: Results <strong>of</strong> <strong>numerical</strong> simulations showing distribution <strong>of</strong> lithologies <strong>and</strong>temperature field developed after 20 Ma. First correspond to low radioactive heatproduction, while second, third <strong>and</strong> fourth columns shows results <strong>of</strong> simulations withhigh heat production <strong>and</strong> different solidus for lower crustal layer.<strong>of</strong> gravitational instabilities controlling vertical extrusion or channel flow mechanismLexa et al. (2011). In this work we show that the thermal structure <strong>of</strong> the orogenis controlled by a position <strong>of</strong> felsic lower crust between the MOHO <strong>and</strong> the formermafic lower crust resulting from lower Paleozoic underplating. This layer, likely richin radioactive elements, is responsible for elevation <strong>of</strong> geotherm up to solidus <strong>of</strong> felsiccrust leading to its dramatic s<strong>of</strong>tening. Finally, the Rayleigh-Taylor instability developsleading to exhumation <strong>of</strong> felsic partially molten material into mid-crustal levels (Fig. 2.4).In this study we discuss as whether: 1) the felsic crust had a proportion <strong>of</strong> radioactiveelements high enough, or 2) as whether the heat source originate from below either from


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 23Figure 2.5: Results <strong>of</strong> gravity inversion <strong>modelling</strong> showing the typical distribution <strong>of</strong>crustal layers in the Bohemian Massif.metasomatised upper mantle, or 3) due to up-welling <strong>of</strong> hot asthenospheric material. It isimportant to note that realistic data <strong>of</strong> thickness <strong>of</strong> felsic lower crustal layer, hanging wallmafic lower crust <strong>and</strong> middle crust derived from geophysical data (Fig. 2.5), includingrealistic values <strong>of</strong> concentration <strong>of</strong> the radioactive elements in the felsic lower crust allowsus to “rapidly” generate large temperature anomaly, which is able to trigger the verticaldoming process by gravitational instability. This exhumation mechanism is driven bothby gravity <strong>and</strong> by the horizontal shortening <strong>and</strong> we call it “Laterally Forced Overturns”process. Relics <strong>of</strong> this felsic lower crust are still preserved in the Moldanubian domainin the depth <strong>of</strong> 40 to 30 km (Guy et al., 2011), while the dense mafic crust occur inmiddle crustal depths (Fig. 2.5).Finally, the remaining problem is the underst<strong>and</strong>ing <strong>of</strong> influx <strong>of</strong> the felsic crust intothe root. We studied this problem in the Western Carpathians on the example <strong>of</strong> Vepordome (Jeřábek et al., 2008, 2007) <strong>and</strong> Orlica-Śnieżnik dome in West Sudetes (Chopinet al., 2011a). The goal <strong>of</strong> these studies is underst<strong>and</strong>ing <strong>of</strong> formation <strong>of</strong> burial orogenicfabrics during continental collision processes. Petrological studies <strong>of</strong> deep gneissesfrom the Vepor basement suggest that the origin <strong>of</strong> first subhorizontal metamorphicfabric is clearly related to the prograde P-T path (prograde zonation <strong>of</strong> garnets in Vepororthogneiss <strong>and</strong> pelites), <strong>and</strong> therefore to progressive burial. This mechanism is inobvious contrast to the generally accepted model <strong>of</strong> flat fabric development during theexhumation, namely the origin <strong>of</strong> “core complexes” due to orogen parallel extension. I


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 24Figure 2.6: Petrological data showing distribution <strong>of</strong> different prograde P-T pathswithin crustal column. Schematic drawing <strong>of</strong> thickening from above model showingprogressive evolution <strong>of</strong> thermal structure in Vepor unit.tried to resolve the paradox, i.e. the development <strong>of</strong> flat fabric during crustal thickeningby <strong>numerical</strong> calculations <strong>and</strong> I <strong>of</strong>fered a model <strong>of</strong> thermal evolution within thickenedcrust controlled Gemer crystalline overthrust (Jerabek,2009). This process <strong>of</strong> “thickeningfrom above” plays an essential role for the thermal evolution <strong>of</strong> a root zone, whichis supported by petrological data. In conclusion, model <strong>of</strong> thickening <strong>of</strong> Vepor crustindicates that successful conjunction <strong>of</strong> petrological, <strong>structural</strong> <strong>and</strong> <strong>numerical</strong> studies isonly possible if they are underpinned by thorough field knowledge.2.1 Perspectives for the <strong>numerical</strong> <strong>modelling</strong> <strong>of</strong> channelflow in the orogens <strong>and</strong> development <strong>of</strong> infra- <strong>and</strong> superstructuretransition zoneExtrusion <strong>of</strong> the lower crust <strong>and</strong> its lateral spreading below brittle-ductile transitionrequires a thorough underst<strong>and</strong>ing <strong>of</strong> the rheological <strong>and</strong> thermal states prevailing duringthe process. First, it is necessary to h<strong>and</strong>le the problem <strong>of</strong> heat source which may be<strong>of</strong> various origins. It is likely, that key role is played by process <strong>of</strong> “relamination” <strong>of</strong>felsic lower crust above the MOHO resulting from massive material influx as shown in


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 25Figure 2.7: Example <strong>of</strong> finite element <strong>modelling</strong> <strong>of</strong> crustal indentation <strong>of</strong> orogenicroot. This model was calculated using modified Elmer s<strong>of</strong>tware (Maierová, 2011) <strong>and</strong>includes temperature <strong>and</strong> strain rate dependent rheology, erosion, isostasy <strong>and</strong> gravitationalforces to simulate more realistic geodynamic evolution.our study from West Carpathians. To underst<strong>and</strong> this process we need to examine avariety <strong>of</strong> heat sources such as the mantle delamination, radiocative mantle or radioactivelower crust. We will orient our research in future to underst<strong>and</strong> relative role <strong>of</strong> variableheat sources, the rheology <strong>of</strong> crust, kinematics <strong>of</strong> gravity overturns <strong>and</strong> more realisticgeometries <strong>of</strong> rock bodies.Horizontal flow <strong>of</strong> orogenic lower crust over the Brunia indenter along eastern marginis a problem that also requires the rheological <strong>and</strong> thermal scaling. For this purpose weset up finite element models (Fig. 2.7), which could solve the problem <strong>of</strong> material transferefficiency <strong>of</strong> channel flow, velocity <strong>and</strong> strain-rate field patterns as well as temperature


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 26relationship between the channel, overburden rigid lid <strong>and</strong> underlying basement. We testvarious combinations <strong>of</strong> rheological, thermal properties <strong>and</strong> boundary conditions whichallow us to underst<strong>and</strong> principal behaviour within parameter space <strong>and</strong> to identify firstorder factors, which may cause the development <strong>of</strong> the channel structure, such as thoseobserved along the eastern margin <strong>of</strong> the Bohemian Massif (FIg ISTRA). It turns outthat the temperature contrast between the continent <strong>and</strong> the root is not the majorparameter to initiate indentation process, but the rheology contrast <strong>of</strong> rocks, whichare partially molten within root zone <strong>and</strong> strong in the indenter represent the mainparameter controlling channel flow extrusion.The last topic that I consider to be very promising is relationship between infra<strong>and</strong>superstructure <strong>of</strong> orogens in terms <strong>of</strong> their contrasting <strong>structural</strong>, rheological <strong>and</strong>thermal evolution. It is the model <strong>of</strong> Culshaw et al. (2006) which first introduced theconcept <strong>of</strong> mechanical coupling during early stages <strong>and</strong> decoupling during late stages<strong>of</strong> orogen evolution. It is during the transition from coupled to decoupled behaviour <strong>of</strong>the crust when the infra <strong>and</strong> superstructure are formed <strong>and</strong> when the infrastructure <strong>and</strong>superstructure transition zone (ISTZ) is established. The coupling/decoupling <strong>and</strong> ISTZformation are clearly associated with extreme localization <strong>of</strong> deformation accompaniedby major change in mechanical properties <strong>of</strong> crustal rocks. Therefore in the light <strong>of</strong>mechanical reasons for the ISTZ formation, several important questions related to themodel <strong>of</strong> Culshaw et al. (2006) emerge. (1) Is the origin <strong>of</strong> ISTZ exclusively controlledby the thermal structure <strong>of</strong> an orogen or can it also be influenced by other rheologicalparameters such as composition, fluids, differential stress, lithostatic pressure <strong>and</strong> strainrate? And if yes, to what extent? (2) What is the exact depth <strong>of</strong> ISTZ break throughin orogens with typically complex internal orogenic architecture? And can we predictits depth in active orogens?In this area, I would like to put efforts <strong>of</strong> my future master’s students <strong>and</strong> doctoralc<strong>and</strong>idates. I believe that areas like contact <strong>of</strong> Orlica-Śnieżnik dome with Zábřeh crystalline,contact <strong>of</strong> Teplá crystalline <strong>and</strong> Barr<strong>and</strong>ian Proterozoic in Bohemian Massif <strong>and</strong>the relationship <strong>of</strong> Vepor dome with Gemer Paleozoic in the Western Carpathians arebest c<strong>and</strong>idates to solve these problems. Relationship between long-lasting <strong>structural</strong>record <strong>of</strong> superstructure <strong>and</strong> juvenile infrastructure (possibly influenced by channel flowprocess) is first-order problem <strong>of</strong> continental <strong>structural</strong> geology <strong>and</strong> in combination withappropriate geochronological <strong>and</strong> petrological methods promises to bring new resultsnot only for Bohemian Massif <strong>and</strong> Carpathians but also for other orogenic systems likeCanadian Cordillera, Grenville tract <strong>and</strong> others. This research should be linked with adetailed micro<strong>structural</strong> studies combined with thermomechanical <strong>modelling</strong> allowing acombination <strong>of</strong> field <strong>and</strong> laboratory research which may be attractive for students withrather limited mathematical background.


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 272.2 Accompanying publicationsSchulmann, K., Lexa, O., Štípská, P., Racek, M., Tajčmanová, L., Konopásek, J.,Edel, J.-B., Peschler, A., Lehmann, J., 2008a. Vertical extrusion <strong>and</strong> horizontalchannel flow <strong>of</strong> orogenic lower crust: key exhumation mechanisms in large hotorogens? Journal <strong>of</strong> Metamorphic Geology 26 (2), 273–297. 133Schulmann, K., Konopásek, J., Janoušek, V., Lexa, O., Lardeaux, J.-M., Edel, J.-B., Štípská, P., Ulrich, S., 2009b. An Andean type Palaeozoic convergence in theBohemian Massif. Comptes Rendus Geosciences 341 (2-3), 266–286. 159 Lexa, O., Schulmann, K., Janoušek, V., Štípská, P., Guy, A., Racek, M., Jan. 2011.Heat sources <strong>and</strong> trigger mechanisms <strong>of</strong> exhumation <strong>of</strong> HP granulites in Variscanorogenic root. Journal <strong>of</strong> Metamorphic Geology 29 (1), 79–102. 181Franěk, J., Schulmann, K., Lexa, O., Tomek, v., Edel, J.-B., 2011a. Model <strong>of</strong>syn-convergent extrusion <strong>of</strong> orogenic lower crust in the core <strong>of</strong> the Variscan belt:implications for exhumation <strong>of</strong> high-pressure rocks in large hot orogens. Journal<strong>of</strong> Metamorphic Geology 29 (1), 53–78. 2052.3 Related publicationsSchulmann, K., Kröner, A., Hegner, E., Wendt, I., Konopásek, J., Lexa, O., Štípská,P., 2005. Chronological constraints on the pre-orogenic history, burial <strong>and</strong> exhumation<strong>of</strong> deep-seated rocks along the eastern margin <strong>of</strong> the Variscan orogen,Bohemian Massif, Czech Republic. American Journal <strong>of</strong> Science 305 (5), 407–448.Franěk, J., Schulmann, K., Lexa, O., 2006. Kinematic <strong>and</strong> rheological model <strong>of</strong>exhumation <strong>of</strong> high pressure granulites in the Variscan orogenic root: Example<strong>of</strong> the Blanský les granulite, Bohemian Massif, Czech Republic. Mineralogy <strong>and</strong>Petrology 86 (3-4), 253–276.Racek, M., Štípská, P., Pitra, P., Schulmann, K., Lexa, O., 2006. Metamorphicrecord <strong>of</strong> burial <strong>and</strong> exhumation <strong>of</strong> orogenic lower <strong>and</strong> middle crust: a new tectonothermalmodel for the Drosendorf window (Bohemian Massif, Austria). Mineralogy <strong>and</strong>Petrology 86 (3-4), 221–251.Jeřábek, P., Faryad, W. S., Schulmann, K., Lexa, O., Tajčmanová, L., 2008. Alpineburial <strong>and</strong> heterogeneous exhumation <strong>of</strong> Variscan crust in the West Carpathians:insight from thermodynamic <strong>and</strong> argon diffusion <strong>modelling</strong>. Journal <strong>of</strong> the GeologicalSociety 165, 479–498.


2. Mechanisms <strong>of</strong> lower crustal flow <strong>and</strong> its thermal <strong>and</strong> mechanical implications 28Skrzypek, E., Štípská, P., Schulmann, K., Lexa, O., Lexová, M., 2011b. Prograde<strong>and</strong> retrograde metamorphic fabrics - a key for underst<strong>and</strong>ing burial <strong>and</strong> exhumationin orogens (Bohemian Massif). Journal <strong>of</strong> Metamorphic Geology 29 (4), 451–472.Skrzypek, E., Schulmann, K., Štípská, P., Chopin, F., Lehmann, J., Lexa, O.,Haloda, J., 2011a. Tectono-metamorphic history recorded in garnet porphyroblasts:insights from thermodynamic <strong>modelling</strong> <strong>and</strong> electron backscatter diffractionanalysis <strong>of</strong> inclusion trails. Journal <strong>of</strong> Metamorphic Geology 29 (4), 473–496.Chopin, F., Schulmann, K., Štípská, P., Martelat, J., Pitra, P., Lexa, O., Petri,B., 2011b. Micro<strong>structural</strong> <strong>and</strong> petrological evolution <strong>of</strong> a high pressure graniticorthogneiss during continental subduction (Orlica-Śnieżnik dome, NE BohemianMassif). Journal <strong>of</strong> Metamorphic Geology, submitted.Štípská, P., Chopin, F., Skrzypek, E., Schulmann, K., Lexa, O., Pitra, P., Martelat,J., Bollinger, C., 2011. Juxtaposition <strong>of</strong> eclogite-facies <strong>and</strong> mid-crustal rocksduring exhumation: relative role <strong>of</strong> erosion <strong>and</strong> crustal scale folding in a continentalwedge (Orlica-Śnieżnik dome, Bohemian Massif). Journal <strong>of</strong> MetamorphicGeology, submitted.


Chapter 3<strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong>metamorphic microstructures,their visualization <strong>and</strong> statisticsWithin the framework <strong>of</strong> micro<strong>structural</strong> <strong>analyses</strong> carried out at the Institute <strong>of</strong> Petrology<strong>and</strong> Structural Geology I developed a code to characterise <strong>and</strong> quantify deformation<strong>and</strong> metamorphic microstructures. This early version allowed quantification <strong>of</strong> strainintensity using digitized grain boundaries similar to PAROR <strong>and</strong> SURFOR methods <strong>of</strong>Panozzo (1983, 1984) ans was for the first time used in micro<strong>structural</strong> styudy <strong>of</strong> deformedmarbes by Ulrich 92. However, we have quickly recognized the potential <strong>of</strong> thetechnique <strong>and</strong> I have implemented a range <strong>of</strong> statistical methods to quantify mineralmicrostructures <strong>of</strong> polyphase <strong>and</strong> monophase recrystallized aggregates. It is in particularthe analysis <strong>of</strong> grain sizes combined with CSD approach which allows determiningsome kinetic characteristics <strong>of</strong> grain nucleation <strong>and</strong> growth. The second approach implementedroutinely in the PolyLX code is the analysis <strong>of</strong> grain boundary frequencieswhich allow evaluating the relative importance <strong>of</strong> surface energy <strong>of</strong> mineral aggregate inquantifying the degree <strong>of</strong> deviation <strong>of</strong> grain contact frequencies from r<strong>and</strong>om distribution.The textural analysis is a powerful, but underused tool <strong>of</strong> petro<strong>structural</strong> analysis.Except acquirement <strong>of</strong> common statistical parameters, this technique can significantlyimprove underst<strong>and</strong>ing <strong>of</strong> processes <strong>of</strong> grain nucleation <strong>and</strong> grain growth,can bring insights on the role <strong>of</strong> surface energies or quantify duration <strong>of</strong> metamorphic<strong>and</strong> magmatic cooling events as long as appropriate thermodynamical data forstudied minerals exists. This technique also allows systematic evaluation <strong>of</strong> degree <strong>of</strong>29


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 30preferred orientations <strong>of</strong> grain boundaries in conjunction with their frequencies. Thismay help to better underst<strong>and</strong> the mobility <strong>of</strong> grain boundaries <strong>and</strong> precipitationsor removal <strong>of</strong> different mineral phases.The new open platform, object-oriented MATLAB ® toolbox PolyLX developedby myself, provide several core routines for data exchange, visualization <strong>and</strong> analysis<strong>of</strong> micro<strong>structural</strong> data, which can be run on any platform supported by MATLAB ® .Detailed descriptions <strong>of</strong> toolbox routines <strong>and</strong> methods <strong>of</strong> implementation <strong>of</strong> newtechniques could be find on dedicated webpage http://petrol.natur.cuni.cz/ ondro.The technique was first used in Lexa et al. (2005) where we examined the problem<strong>of</strong> micro<strong>structural</strong> maturity related to the duration <strong>of</strong> two metamorphic events whichoccurred at similar pressure <strong>and</strong> temperature conditions. We analysed samples <strong>of</strong> amphibolites<strong>and</strong> tonalitic gneisses that were affected by ductile deformation associatedwith emplacement <strong>of</strong> thin Carboniferous granodiorite sill. This event was controlledby intense solid state recrystallization due to intense deformation triggered by heatingeffect <strong>of</strong> adjacent granodiorite sill. On the other h<strong>and</strong> we studied similar rocks (amphibolites<strong>and</strong> tonalitic orthogneiss) that recrystallized during long lasting evolution <strong>of</strong>Cambro-Ordovician intracontinental rift. This event reveals much longer thermal timescale compared to the Carboniferous event for surprisingly similar PT conditions. The


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 31Figure 3.1: N 0 -Gt plot showing CSD evolution <strong>of</strong> plagioclase <strong>and</strong> hornblende. Thisfigure showing two contrasting positions for metamorphic events with different durations.Long-lived mature microstructures are grouped along so-called fabric attractor.micro<strong>structural</strong> analysis covered digitization <strong>of</strong> large number <strong>of</strong> samples which weresubsequently h<strong>and</strong>led using PolyLX s<strong>of</strong>tware. As a result <strong>of</strong> this analysis we discoveredthat the Carboniferous event was associated with grain size distribution whichin CSD space shown dominance <strong>of</strong> nucleation process with respect to slow growth rate.Increase <strong>of</strong> ambient temperature correlates with decrease in nucleation density accompaniedwith increase growth (Fig. 3.1). In contrast the Cambro-Ordovician rocks revealedpre-dominance <strong>of</strong> growth process over the nucleation. Our data plot on continuous curvein the CSD space which clearly indicates absence <strong>of</strong> Ostwald ripening <strong>and</strong> dominance <strong>of</strong>textural coarsening by DeH<strong>of</strong>f’s communicating neighbours theory, which governs thechange in shape <strong>of</strong> CSD curve. The analysis show that the time matters in mineral microstructuresleading to development <strong>of</strong> grain size distribution which can be consideredas a grain size attractor (Fig. 3.1).This study had shown for the first time the power <strong>of</strong> grain contact frequency analysis.The rocks deformed by short lived Carboniferous pulse reveal aggregate distribution<strong>of</strong> both mafic <strong>and</strong> felsic phases. However, the similar lithologies experiencing long lastingCambro-Ordovician event recorded evolution driving contact frequencies towardsthe field <strong>of</strong> regular distribution. The process <strong>of</strong> increase <strong>of</strong> unlike contacts in rocks is


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 32interpreted as mechanism <strong>of</strong> decrease <strong>of</strong> interfacial surface energy which is not only governedby mineral growth <strong>and</strong> development <strong>of</strong> triple point network as suggested by otherauthors. These differences were also underlined by strong shape preferred orientationcoupled with high aspect ratios for short lived deformation event while the protractedheating was characterized by almost r<strong>and</strong>om microstructure in terms <strong>of</strong> shapes <strong>of</strong> grain<strong>and</strong> their orientation. Moreover accompanied systematic micro<strong>structural</strong> analysis <strong>of</strong>gabbros deformed at different thermal conditions <strong>of</strong> 600°C <strong>and</strong> 700°C (Baratoux et al.,2005b) shown that identical lithologies reveal different CSD patterns, shape preferredorientations <strong>and</strong> grain contact frequencies which can be attributed to contrasting deformationmechanisms governing rock deformation. It was shown that the temperature isa leading factor controlling processes like mechanical mixing <strong>of</strong> phases, growth versusnucleation rates <strong>and</strong> internal strain partitioning within rock. We had shown in this workthat the mineral microstructure has a potential to reflec thermal conditions better thanany other thermometric method.Finally, we examined the micro<strong>structural</strong> characteristic <strong>of</strong> amphibolites deformedat different metamorphic grade <strong>and</strong> subsequently folded by independent deformationevent (Baratoux et al., 2005a). This study showed the power <strong>of</strong> PolyLX method inevaluating the degree <strong>of</strong> mechanical anisotropy <strong>of</strong> rocks in terms <strong>of</strong> mineral elongation,degree <strong>of</strong> preferred orientation <strong>of</strong> grain boundaries <strong>and</strong> their spatial distribution. Itwas demonstrated that the b<strong>and</strong>ed amphibolites with high degree <strong>of</strong> shape <strong>and</strong> grainboundary preferred orientation, important aggregate distribution <strong>and</strong> highly contrastinggrain size reveal high mechanical anisotropy while material that shows low aspect ratio,shape <strong>and</strong> grain boundary preferred orientation in conjunction with regular grain contactdistribution <strong>and</strong> similar grain size distributions <strong>of</strong> phases <strong>of</strong> contrasting rheologies revealsvery low mechanical anisotropy. These two contrasting mineral architectures than governfurther behaviour <strong>of</strong> rock when external stress is applied leading to almost ideal bucklingin the first case <strong>and</strong> passive flow folding in the latter.The PolyLX was successfully applied to determine deformation mechanisms <strong>of</strong> quartzin study <strong>of</strong> Jeřábek et al. (2007) where we used the technique for paleostress analysis<strong>of</strong> recrystallized quartz for determination <strong>of</strong> flow stress in the Vepor orthogneiss. Themethod revealed high reproducibility <strong>of</strong> datasets <strong>and</strong> precision in grain size determinationleading to highly precise stress estimates which are impossible to be achieved byst<strong>and</strong>ard stereometric measurements. This is because the PolyLX involves the solution<strong>of</strong> 3D sectional problem which is commonly ignored in similar studies. In combinationwith temperature estimates we were able to produce a stress-temperature deformationmap, which shown positive correlation <strong>of</strong> stress <strong>and</strong> temperature for relatively high


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 33Figure 3.2: Digitized microstructure images <strong>of</strong> highly anisotropic low temperatureamphibolite fabric <strong>and</strong> weakly anisotropic high temperature amphibolites.grade samples that was explained in term <strong>of</strong> progressive burial <strong>and</strong> temperature increase.Negative correlation <strong>of</strong> stress <strong>and</strong> temperature was found for low grade samples<strong>and</strong> was explained by stress/strain rate partitioning due to weak matrix effect.During subsequent years we concentrated our efforts on studies <strong>of</strong> partially moltenrocks especially orthogneisses. The PolyLX method was used by Závada et al. (2007)to study the microstructure <strong>and</strong> topology <strong>of</strong> melt seams related to mylonitic flow <strong>of</strong>orthogneiss deformed at high melt/fluid pressure. We have shown that at high meltpressure <strong>and</strong> low differential stress conditions the orthogneiss deformed at high temperatureconditions can yield in brittle manner involving so called cavitation process. ThePolyLX method allowed quantifying the topology <strong>of</strong> intragranular melt filled fracturesthereby providing decisive argument for unusual cavitation dominated grain boundarysliding. The diffusion creep accommodated by cavitation fracturing <strong>of</strong> feldspar aggregateproduces disproportionally lower ductility <strong>of</strong> feldspars compared to quartz, whichmodify our view on rheology <strong>of</strong> the orogenic lower crust.Our melt related study culminated in two papers where the PolyLX s<strong>of</strong>tware playedthe decisive role. In partially molten orthogneiss from the Kutná Hora crystalline complexSchulmann et al. (2008b) shown that with increasing melt proportion the crystalsize distribution <strong>of</strong> main phases show protracted growth in comparison to reduced nucleationrate, while the interstitial phases revealed increasing degree <strong>of</strong> nucleation rate


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 34Figure 3.3: Rf/phi plot quantifying the orientation <strong>and</strong> shape <strong>of</strong> intragranular meltfilledvoids in K-fedspar aggregate <strong>and</strong> 3d image <strong>of</strong> melt distribution in monomineralicaggregate.Figure 3.4: Grain contact frequency vs. brain boundary preferred orientation showingthe evolution <strong>of</strong> microstructure from highly aggregate to r<strong>and</strong>om <strong>and</strong> regular distributionswith increasing deformation.(Fig. 3.4). At the same time, the rock reveals first increase <strong>of</strong> aggregate distributionrelated with development <strong>of</strong> b<strong>and</strong>ed mylonite structure followed by complete mixing <strong>of</strong>all phases related to influx <strong>of</strong> melt <strong>and</strong> the rheological collapse <strong>of</strong> the rock. We correlatedthese stages <strong>of</strong> orthogneiss deformation with increasing amount <strong>of</strong> melt pressure whichreduced effective stress at constant (<strong>and</strong> low) differential stress level conditions. In fact,the amount <strong>of</strong> very small melt proportion controls not only the topology <strong>of</strong> melt seamsbut also dramatically weaken rock at melt fraction for about 5%Hasalová et al. (2008a) used the PolyLX method to evaluate the role <strong>of</strong> melt infiltrationinto orthogneiss microstructure. This work shows surprising CSD evolution from


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 35Figure 3.5: N 0 -Gt plots showing evolution <strong>of</strong> CSD from growth dominated to nucleationdominated micro<strong>structural</strong> stages within crustal channel. This evolution isinterpreted as result <strong>of</strong> syntectonic melt infiltration.originally high Gt/N 0 values towards low values in conjunction with increasing amount<strong>of</strong> mineral resorption <strong>and</strong> mineral overgrowth features. This particular evolution canbe only explained by increasing importance <strong>of</strong> nucleation rate which is incompatiblewith process <strong>of</strong> in-situ melting but reflect crystallization <strong>of</strong> melt in rock pores. TheseCSD plots were therefore used as a major argument for reactive porous flow in felsicrocks, a concept which was for the first time demonstrated in the continental crust. TheCSD results are supported by evolution in grain contact frequency plot which revealssystematically increasing importance <strong>of</strong> regular grain distributions which is driven tounexpected values. The regular grain distribution simply indicates progressive growth<strong>of</strong> new phases in the rock aggregate as the melt crystallizes in the rock. Finally, theprogressive melt infiltration is related with loss <strong>of</strong> grain shape preferred orientation coupledwith loss <strong>of</strong> aspect ratios, which in turn is connected with increasing development<strong>of</strong> preferred orientation <strong>of</strong> certain unlike boundaries. All that indicates that the porousflow was a dynamic process related to dynamic dilation <strong>of</strong> intragranular pores whichmaintained the porosity <strong>of</strong> rock at high level as long as the reactive porous flow wasactive.Our last studies are focused on quantification <strong>of</strong> micro<strong>structural</strong> evolution <strong>of</strong> granulitetextures (Franěk et al., 2011b). We discovered precursor rock <strong>of</strong> felsic granuliteswhich is interpreted as coarse grained orthogneiss, that was converted to granulite duringCarboniferous thickening <strong>and</strong> collision. These studies show that the hypersolvusalkaline feldspar is first decomposed into thick perthites that are rapidly replaced bypolycrystalline, equidimensional aggregate <strong>of</strong> pure K feldspar <strong>and</strong> plagioclase (Fig. 3.6).This decomposition is driven by heterogeneous nucleation process resulting from stored


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 36Figure 3.6: Example <strong>of</strong> digitized microstructures originated from heterogenous decomposition<strong>of</strong> alkaline feldspar due to increase <strong>of</strong> temperature.strain energy in the aggregate. Such a strain energy is driven by decompression associatedwith slight cooling <strong>of</strong> perthitic aggregate <strong>and</strong> leads to development <strong>of</strong> particularmicrostructure which reduces both strain <strong>and</strong> surface energies <strong>of</strong> the aggregate. ThePolyLX played decisive role in the characterizing the process <strong>and</strong> it shows that the microstructurereveals highest nucleation density compared to growth in conjunction withhighest possible regular grain distribution which was recorded in metamorphic rocks s<strong>of</strong>ar. This microstructure showed to be a precursor for granulite mylonitic fabric. ThePolyLX allows tracking the microstructure evolution <strong>of</strong> the rock <strong>and</strong> helps to define firstgrain boundary sliding diffusional process replaced subsequently by dislocation creepflow at lower temperatures.3.1 Perspectives for <strong>numerical</strong> modeling <strong>of</strong> deformationmicrostructures <strong>and</strong> their quantificationPolyLX method revealed high applicability in micro<strong>structural</strong> analysis thanks to its versatile<strong>and</strong> modular character allowing large amount <strong>of</strong> applications in paleopiezometry,crystal size distribution, grain contact frequencies <strong>and</strong> deformation mechanisms studiesaccompanied with texture <strong>analyses</strong> (Fig. 3.7). The potential is therefore in further use<strong>of</strong> the method in a variety <strong>of</strong> micro<strong>structural</strong> studies which can be combined with kineticgrow/nucleation models if the thermal history is known. For that purpose a coupling


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 37Figure 3.7: Example <strong>of</strong> textural data obtained from EBSD analysed by MTEX <strong>and</strong>subsequently imported into PolyLX to obtain micro<strong>structural</strong> quantitative datawith ELLE code is required. In recent collaboration with leaders in ELLE programming(Mark Jessell - Toulouse, S<strong>and</strong>ra Piazollo - Stockholm <strong>and</strong> Paul D. Bons - Tubingen)we achieved a full compatibility between both codes so, that now the micro<strong>structural</strong>processes can be both simulated <strong>and</strong> quantified. In close future we plan to test simpledeformation - growth simulations for known thermal histories coupled with detailedstatistical analysis <strong>of</strong> rock microstructure. Our aim is to define parameters allowing toprecisely characterize the microstructure relationships to stress <strong>and</strong> temperature as it isdone in metallurgy.In nature factors controlling the active deformation mechanisms <strong>and</strong> hence the rheologyinclude those that are properties <strong>of</strong> the deforming material (e.g. composition <strong>and</strong>mineral assemblage, texture, grain size), <strong>and</strong> those that are imposed on the system fromoutside (confining pressure, differential stress, temperature <strong>and</strong> fluid pressure). Thereforewe plan coupling <strong>of</strong> precise identification <strong>of</strong> PT evolution <strong>of</strong> rock carried out throughThermocalc <strong>and</strong> Perple X codes in collaboration with leading petrologists in the field(Roger Powell (Melbourne), Richard White (Mainz) <strong>and</strong> Pavla Štípská (Strasbourg))in order to attribute to compositional characteristics <strong>of</strong> mineral microstructures moreor less precise PT conditions <strong>and</strong> PT trends. This approach will allow us to search formicrostructures that are unique for given PT conditions but also time <strong>of</strong> rock residencein given thermal conditions. The latter parameter will be quantified by <strong>analyses</strong> <strong>of</strong> interfacialangles that are functions <strong>of</strong> time <strong>of</strong> annealing, but also phase spatial distributions<strong>and</strong> grain size evolutions.


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 383.2 Accompanying publicationsLexa, O., Štípská, P., Schulmann, K., Baratoux, L., Kröner, A., 2005. Contrastingtextural record <strong>of</strong> two distinct metamorphic events <strong>of</strong> similar P-T conditions <strong>and</strong>different durations. Journal <strong>of</strong> Metamorphic Geology 23 (8), 649–666. 231Baratoux, L., Schulmann, K., Ulrich, S., Lexa, O., 2005b. Contrasting microstructures<strong>and</strong> deformation mechanisms in metagabbro mylonites contemporaneouslydeformed under different temperatures (c. 650°C <strong>and</strong> c. 750°C).; Deformationmechanisms, rheology <strong>and</strong> tectonics; from minerals to the lithosphere. Journal<strong>of</strong> Structural Geology 243 (4), 97–125. 249Závada, P., Schulmann, K., Konopásek, J., Ulrich, S., Lexa, O., 2007. Extremeductility <strong>of</strong> feldspar aggregates–melt-enhanced grain boundary sliding <strong>and</strong> creepfailure: rheological implications for felsic lower crust. Journal <strong>of</strong> Geophysical Research,B, Solid Earth <strong>and</strong> Planets 112, B10210, doi:10.1029/2006JB004820.279Schulmann, K., Martelat, J.-E., Ulrich, S., Lexa, O., Štípská, P., Becker, J. K., Oct.2008b. Evolution <strong>of</strong> microstructure <strong>and</strong> melt topology in partially molten graniticmylonite: Implications for rheology <strong>of</strong> felsic middle crust. Journal <strong>of</strong> GeophysicalResearch, B, Solid Earth <strong>and</strong> Planets 113, B10406, doi:10.1029/2007JB005508.295Hasalová, P., Schulmann, K., Lexa, O., Štípská, P., Hrouda, F., Ulrich, S., Haloda,J., Týcová, P., 2008a. Origin <strong>of</strong> migmatites by deformation-enhanced melt infiltration<strong>of</strong> orthogneiss: a new model based on quantitative micro<strong>structural</strong> analysis.Journal <strong>of</strong> Metamorphic Geology 26, 29–53. 315Franěk, J., Schulmann, K., Lexa, O., Ulrich, S., Štípská, P., Haloda, J., Týcová, P.,2011b. Origin <strong>of</strong> felsic granulite microstructure by heterogeneous decomposition <strong>of</strong>alkali feldspar <strong>and</strong> extreme weakening <strong>of</strong> orogenic lower crust during the Variscanorogeny. Journal <strong>of</strong> Metamorphic Geology 29 (1), 103–130. 3413.3 Related publicationsBarraud, J., 2006. The use <strong>of</strong> watershed segmentation <strong>and</strong> GIS s<strong>of</strong>tware for texturalanalysis <strong>of</strong> thin sections. Journal <strong>of</strong> Volcanology <strong>and</strong> Geothermal Research 154 (1-2), 17–33.


3. <strong>Quantitative</strong> <strong>analyses</strong> <strong>of</strong> metamorphic microstructures 39Machek, M., Špaček, P., Ulrich, S., Heidelbach, F., 2007. Origin <strong>and</strong> orientation<strong>of</strong> microporosity in eclogites <strong>of</strong> different microstructure studied by ultrasound <strong>and</strong>micr<strong>of</strong>abric analysis. Engineering Geology 89 (3-4), 266–277.Jeřábek, P., Stunitz, H., Heilbronner, R., Lexa, O., Schulmann, K., 2007. Micro<strong>structural</strong>- deformation record <strong>of</strong> an orogen-parallel extension in the Veporunit, West Carpathians. Journal <strong>of</strong> Structural Geology 29 (11), 1722–1743.Závada, P., Schulmann, K., Lexa, O., Hrouda, F., Haloda, J., Týcová, P., 2009. Themechanism <strong>of</strong> flow <strong>and</strong> fabric development in mechanically anisotropic trachytelava. Journal <strong>of</strong> Structural Geology 31 (11), 1295–1307.Skrzypek, E., Schulmann, K., Štípská, P., Chopin, F., Lehmann, J., Lexa, O.,Haloda, J., 2011a. Tectono-metamorphic history recorded in garnet porphyroblasts:insights from thermodynamic <strong>modelling</strong> <strong>and</strong> electron backscatter diffractionanalysis <strong>of</strong> inclusion trails. Journal <strong>of</strong> Metamorphic Geology 29 (4), 473–496.Chopin, F., Schulmann, K., Štípská, P., Martelat, J., Pitra, P., Lexa, O., Petri,B., 2011b. Micro<strong>structural</strong> <strong>and</strong> petrological evolution <strong>of</strong> a high pressure graniticorthogneiss during continental subduction (Orlica-Śnieżnik dome, NE BohemianMassif). Journal <strong>of</strong> Metamorphic Geology, submitted.Oliot, E., Lexa, O., Schulmann, K., Goncalves, P., Marquer, D., 2011. Mid-crustalstrain localization in granitic rocks: constraints from quantitative textural <strong>and</strong>crystallographic preferred orientations <strong>analyses</strong>. Tectonophysics, submitted.Martelat, J.-E., Malamoud, K., Cordier, B., R<strong>and</strong>rianasolo, B., Schulmann, K.,Lardeaux, J.-M., 2011. Garnet crystal plasticity in the continental crust, new examplefrom south Madagascar. Journal <strong>of</strong> Metamorphic Geology, submitted.


BibliographyAerden, D., 1998. Tectonic evolution <strong>of</strong> the Montagne Noire <strong>and</strong> a possible orogenicmodel for syncollisional exhumation <strong>of</strong> deep rocks, Variscan belt, France. Tectonics17 (1), 62–79.Baratoux, L., Lexa, O., Cosgrove, J. W., Schulmann, K., 2005a. The quantitative linkbetween fold geometry, mineral fabric <strong>and</strong> mechanical anisotropy; as exemplified bythe deformation <strong>of</strong> amphibolites across a regional metamorphic gradient. Journal <strong>of</strong>Structural Geology 27 (4), 707–730.Baratoux, L., Schulmann, K., Ulrich, S., Lexa, O., 2005b. Contrasting microstructures<strong>and</strong> deformation mechanisms in metagabbro mylonites contemporaneously deformedunder different temperatures (c. 650°C <strong>and</strong> c. 750°C).; Deformation mechanisms, rheology<strong>and</strong> tectonics; from minerals to the lithosphere. Journal <strong>of</strong> Structural Geology243 (4), 97–125.Barraud, J., 2006. The use <strong>of</strong> watershed segmentation <strong>and</strong> GIS s<strong>of</strong>tware for texturalanalysis <strong>of</strong> thin sections. Journal <strong>of</strong> Volcanology <strong>and</strong> Geothermal Research 154 (1-2),17–33.Behrmann, J., 1988. Crustal scale extension in a convergent orogen: the Sterzing-Steinach mylonite zone in the Eastern Alps. Geodynamica Acta 2, 63–73.Burg, J.-P., 1999. Ductile structures <strong>and</strong> instabilities: their implication for Variscantectonics in the Ardennes. Tectonophysics 309 (1-4), 1–25.Chopin, F., Schulmann, K., Skrzypek, E., Lehmann, J., Dujardin, J., Martelat, J.-E., Lexa, O., Corsini, M., Edel, J.-B., Štípská, P., Pitra, P., 2011a. Crustal influx,indentation, ductile thinning <strong>and</strong> gravity redistribution in a continental wedge: keyprocesses in the evolution <strong>of</strong> a mantle gneiss dome (European Variscan belt). Tectonics,submitted.Chopin, F., Schulmann, K., Štípská, P., Martelat, J., Pitra, P., Lexa, O., Petri, B.,2011b. Micro<strong>structural</strong> <strong>and</strong> petrological evolution <strong>of</strong> a high pressure granitic orthogneissduring continental subduction (Orlica-Śnieżnik dome, NE Bohemian Massif).Journal <strong>of</strong> Metamorphic Geology, submitted.41


Bibliography 42Culshaw, N., Beaumont, C., Jamieson, R., 2006. The orogenic superstructureinfrastructureconcept: Revisited, quantified, <strong>and</strong> revived. Geology 34 (9), 733–736.Dewey, J. F., 1975. Finite plate evolution: some implications for the evolution <strong>of</strong> rockmasses at plate margins. American Journal <strong>of</strong> Science 275A, 260–284.Dumicz, M., 1979. Tectogenesis <strong>of</strong> the metamorphosed series <strong>of</strong> the Klodzko district: atentative explanation. Geologia Sudetica 14 (2), 29–46.Engl<strong>and</strong>, P., Houseman, G., 1986. Finite strain calculations <strong>of</strong> continental deformation2. Comparison with the India-Asia collision zone. Journal <strong>of</strong> Geophysical Research,B, Solid Earth <strong>and</strong> Planets 91 (3), 3664–3676.Franěk, J., Schulmann, K., Lexa, O., 2006. Kinematic <strong>and</strong> rheological model <strong>of</strong> exhumation<strong>of</strong> high pressure granulites in the Variscan orogenic root: Example <strong>of</strong> the Blanskýles granulite, Bohemian Massif, Czech Republic. Mineralogy <strong>and</strong> Petrology 86 (3-4),253–276.Franěk, J., Schulmann, K., Lexa, O., Tomek, v., Edel, J.-B., 2011a. Model <strong>of</strong> synconvergentextrusion <strong>of</strong> orogenic lower crust in the core <strong>of</strong> the Variscan belt: implicationsfor exhumation <strong>of</strong> high-pressure rocks in large hot orogens. Journal <strong>of</strong> MetamorphicGeology 29 (1), 53–78.Franěk, J., Schulmann, K., Lexa, O., Ulrich, S., Štípská, P., Haloda, J., Týcová, P.,2011b. Origin <strong>of</strong> felsic granulite microstructure by heterogeneous decomposition <strong>of</strong>alkali feldspar <strong>and</strong> extreme weakening <strong>of</strong> orogenic lower crust during the Variscanorogeny. Journal <strong>of</strong> Metamorphic Geology 29 (1), 103–130.Guy, A., Edel, J.-B., Schulmann, K., Tomek, v., Lexa, O., 2011. A geophysical model<strong>of</strong> the Variscan orogenic root (Bohemian Massif): Implications for modern collisionalorogens. Lithos 124 (1-2), 144–157.Harl<strong>and</strong>, W. B., 1971. Tectonic transpression in Caledonian Spitsbergen. GeologicalMagazine 108, 27–41.Hasalová, P., Schulmann, K., Lexa, O., Štípská, P., Hrouda, F., Ulrich, S., Haloda, J.,Týcová, P., 2008a. Origin <strong>of</strong> migmatites by deformation-enhanced melt infiltration <strong>of</strong>orthogneiss: a new model based on quantitative micro<strong>structural</strong> analysis. Journal <strong>of</strong>Metamorphic Geology 26, 29–53.Hasalová, P., Štípská, P., Powell, R., Schulmann, K., Janoušek, V., Lexa, O., 2008b.Transforming mylonitic metagranite by open-system interactions during melt flow.Journal <strong>of</strong> Metamorphic Geology 26, 55–80.Houseman, G., Engl<strong>and</strong>, P., 1986. Finite strain calculations <strong>of</strong> continental deformation1. Method <strong>and</strong> general results for convergent zones. Journal <strong>of</strong> Geophysical Research,B, Solid Earth <strong>and</strong> Planets 91 (3), 3651–3663.


Bibliography 43Jeřábek, P., Faryad, W. S., Schulmann, K., Lexa, O., Tajčmanová, L., 2008. Alpineburial <strong>and</strong> heterogeneous exhumation <strong>of</strong> Variscan crust in the West Carpathians:insight from thermodynamic <strong>and</strong> argon diffusion <strong>modelling</strong>. Journal <strong>of</strong> the GeologicalSociety 165, 479–498.Jeřábek, P., Stunitz, H., Heilbronner, R., Lexa, O., Schulmann, K., 2007. Micro<strong>structural</strong>- deformation record <strong>of</strong> an orogen-parallel extension in the Vepor unit, WestCarpathians. Journal <strong>of</strong> Structural Geology 29 (11), 1722–1743.Konopásek, J., Schulmann, K., Lexa, O., 2001. Structural evolution <strong>of</strong> the central part <strong>of</strong>the Krušné Hory (Erzgebirge) Mountains in the Czech Republic; evidence for changingstress regime during Variscan compression. Journal <strong>of</strong> Structural Geology 23 (9), 1373–1392.Konopásek, J., Schulmann, K., Lexa, O., 2003. Reply to comments by A. Krohe <strong>and</strong> A.P.Willner on ”Structural evolution <strong>of</strong> the central part <strong>of</strong> the Kru né Hory (Erzgebirge)Mountains in the Czech Republic—evidence for changing stress regime during Variscancompression”. Journal <strong>of</strong> Structural Geology 25 (6), 1005–1007.Koyi, H., Ge<strong>of</strong>frey Milnes, A., Schmeling, H., Talbot, C., Juhlin, C., Zeyen, H., 1999.Numerical models <strong>of</strong> ductile rebound <strong>of</strong> crustal roots beneath mountain belts. GeophysicalJournal International 139 (2), 556–562.Kratinová, Z., Ježek, J., Schulmann, K., Hrouda, F., Shail, R. K., Lexa, O., 2010.Noncoaxial K-feldspar <strong>and</strong> AMS subfabrics in the L<strong>and</strong>’s End granite, Cornwall: Evidence<strong>of</strong> magmatic fabric decoupling during late deformation <strong>and</strong> matrix crystallization.Journal <strong>of</strong> Geophysical Research, B, Solid Earth <strong>and</strong> Planets 115, B09104,doi:10.1029/2009JB006714.Krohe, A., Willner, A. P., 2003. Discussion on: ”Structural evolution <strong>of</strong> the centralpart <strong>of</strong> the Krušné Hory (Erzgebirge) Mountains in the Czech Republic; evidence forchanging stress regime during Variscan compression” by J. Konopásek, K. Schulmann<strong>and</strong> O. Lexa. Journal <strong>of</strong> Structural Geology 25 (6), 1001–1004.Lehmann, J., Schulmann, K., Edel, J.-B., Hrouda, F., Lexa, O., Štípská, P., Haloda,J., 2011. AMS <strong>and</strong> <strong>structural</strong> record <strong>of</strong> granitoid sills emplacement during growth<strong>of</strong> continental gneiss dome: implications for exhumation <strong>of</strong> deep crust. Journal <strong>of</strong>Structural Geology, submitted.Lexa, O., Cosgrove, J., Schulmann, K., 2004. Apparent shear-b<strong>and</strong> geometry resultingfrom oblique fold sections. Journal <strong>of</strong> Structural Geology 26 (1), 155–161.Lexa, O., Schulmann, K., Hrouda, F., in prep. De-formation <strong>of</strong> magmatic bodies in hotorogenic channel. Journal <strong>of</strong> Geophysical Research, B, Solid Earth <strong>and</strong> Planets.Lexa, O., Schulmann, K., Janoušek, V., Štípská, P., Guy, A., Racek, M., Jan. 2011. Heat


Bibliography 44sources <strong>and</strong> trigger mechanisms <strong>of</strong> exhumation <strong>of</strong> HP granulites in Variscan orogenicroot. Journal <strong>of</strong> Metamorphic Geology 29 (1), 79–102.Lexa, O., Schulmann, K., Ježek, J., 2003. Cretaceous collision <strong>and</strong> indentation in theWestern Carpathians: View based on <strong>structural</strong> analysis <strong>and</strong> <strong>numerical</strong> modeling.Tectonics 22 (6), 1066, doi:10.1029/2002TC001472.Lexa, O., Štípská, P., Schulmann, K., Baratoux, L., Kröner, A., 2005. Contrastingtextural record <strong>of</strong> two distinct metamorphic events <strong>of</strong> similar P-T conditions <strong>and</strong>different durations. Journal <strong>of</strong> Metamorphic Geology 23 (8), 649–666.Machek, M., Špaček, P., Ulrich, S., Heidelbach, F., 2007. Origin <strong>and</strong> orientation <strong>of</strong> microporosityin eclogites <strong>of</strong> different microstructure studied by ultrasound <strong>and</strong> micr<strong>of</strong>abricanalysis. Engineering Geology 89 (3-4), 266–277.Martelat, J.-E., Malamoud, K., Cordier, B., R<strong>and</strong>rianasolo, B., Schulmann, K.,Lardeaux, J.-M., 2011. Garnet crystal plasticity in the continental crust, new examplefrom south Madagascar. Journal <strong>of</strong> Metamorphic Geology, submitted.Matte, P., Lancelot, J., Mattauer, M., 1998. The Montagne Noire Hercynian AxialZone is not an extensional metamorphic core complex but a compresional post-nappeanticline with an anatectic core. Geodinamica Acta 11, 13–22.McKenzie, D. P., Parker, R. L., 1967. The North Pacific: an example <strong>of</strong> tectonics on asphere. Nature 216 (5122), 1276–1280.Neubauer, F., Genser, J., Kurz, W., Wang, X., 1999. Exhumation <strong>of</strong> the Tauern window,Eastern Alps. Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth, Part A: Solid Earth <strong>and</strong> Geodesy24 (8), 675–680.Oliot, E., Lexa, O., Schulmann, K., Goncalves, P., Marquer, D., 2011. Mid-crustalstrain localization in granitic rocks: constraints from quantitative textural <strong>and</strong> crystallographicpreferred orientations <strong>analyses</strong>. Tectonophysics, submitted.Panozzo, R., 1983. Two dimensional analysis <strong>of</strong> shape fabric using projections <strong>of</strong> linesin a plane. Tectonophysics 95, 279–294.Panozzo, R., 1984. Two dimensional strain from the orientation <strong>of</strong> lines in a plane.Journal <strong>of</strong> Structural Geology 6, 215–221.Platt, J. P., Vissers, R. L., 1989. Extensional collapse <strong>of</strong> thickened continental lithosphere:A working hypothesis for the Alboran Sea <strong>and</strong> Gibraltar arc. Geology 17 (6),540–543.Racek, M., Štípská, P., Pitra, P., Schulmann, K., Lexa, O., 2006. Metamorphic record<strong>of</strong> burial <strong>and</strong> exhumation <strong>of</strong> orogenic lower <strong>and</strong> middle crust: a new tectonothermalmodel for the Drosendorf window (Bohemian Massif, Austria). Mineralogy <strong>and</strong>


Bibliography 45Petrology 86 (3-4), 221–251.Ratschbacher, L., Frisch, W., Neubauer, F., Schmid, S. M., Neugebauer, J., 1989. Extensionin compressional orogenic belts: The eastern Alps. Geology 17 (5), 404–407.Rey, P., V<strong>and</strong>erhaeghe, O., Teyssier, C., 2001. Gravitational collapse <strong>of</strong> the continentalcrust: definition, regimes <strong>and</strong> modes. Tectonophysics 342 (3-4), 435–449.Rosenberg, C., Brun, J.-P., Gapais, D., November, 2004. Indentation model <strong>of</strong> the EasternAlps <strong>and</strong> the origin <strong>of</strong> the Tauern Window. Geology 32 (11), 997–1000.S<strong>and</strong>erson, D. J., Marchini, W., 1984. Transpression. Journal <strong>of</strong> Structural Geology6 (5), 449–458.Schulmann, K., Edel, J.-B., Hasalová, P., Cosgrove, J., Ježek, J., Lexa, O., 2009a. Influence<strong>of</strong> melt induced mechanical anisotropy on the magnetic fabrics <strong>and</strong> rheology <strong>of</strong>deforming migmatites, Central Vosges, France. Journal <strong>of</strong> Structural Geology 31 (10),1223–1237.Schulmann, K., Konopásek, J., Janoušek, V., Lexa, O., Lardeaux, J.-M., Edel, J.-B.,Štípská, P., Ulrich, S., 2009b. An Andean type Palaeozoic convergence in the BohemianMassif. Comptes Rendus Geosciences 341 (2-3), 266–286.Schulmann, K., Kröner, A., Hegner, E., Wendt, I., Konopásek, J., Lexa, O., Štípská, P.,2005. Chronological constraints on the pre-orogenic history, burial <strong>and</strong> exhumation <strong>of</strong>deep-seated rocks along the eastern margin <strong>of</strong> the Variscan orogen, Bohemian Massif,Czech Republic. American Journal <strong>of</strong> Science 305 (5), 407–448.Schulmann, K., Lexa, O., Štípská, P., Racek, M., Tajčmanová, L., Konopásek, J., Edel,J.-B., Peschler, A., Lehmann, J., 2008a. Vertical extrusion <strong>and</strong> horizontal channel flow<strong>of</strong> orogenic lower crust: key exhumation mechanisms in large hot orogens? Journal <strong>of</strong>Metamorphic Geology 26 (2), 273–297.Schulmann, K., Martelat, J.-E., Ulrich, S., Lexa, O., Štípská, P., Becker, J. K., Oct.2008b. Evolution <strong>of</strong> microstructure <strong>and</strong> melt topology in partially molten graniticmylonite: Implications for rheology <strong>of</strong> felsic middle crust. Journal <strong>of</strong> GeophysicalResearch, B, Solid Earth <strong>and</strong> Planets 113, B10406, doi:10.1029/2007JB005508.Schulmann, K., Thompson, A. B., Lexa, O., Ježek, J., 2003. Strain distribution<strong>and</strong> fabric development modeled in active <strong>and</strong> ancient transpressive zones. Journal<strong>of</strong> Geophysical Research, B, Solid Earth <strong>and</strong> Planets 108(B1), 2023, doi:10.1029/2001JB000632.Skrzypek, E., Schulmann, K., Štípská, P., Chopin, F., Lehmann, J., Lexa, O., Haloda, J.,2011a. Tectono-metamorphic history recorded in garnet porphyroblasts: insights fromthermodynamic <strong>modelling</strong> <strong>and</strong> electron backscatter diffraction analysis <strong>of</strong> inclusiontrails. Journal <strong>of</strong> Metamorphic Geology 29 (4), 473–496.


Bibliography 46Skrzypek, E., Štípská, P., Schulmann, K., Lexa, O., Lexová, M., 2011b. Prograde <strong>and</strong>retrograde metamorphic fabrics - a key for underst<strong>and</strong>ing burial <strong>and</strong> exhumation inorogens (Bohemian Massif). Journal <strong>of</strong> Metamorphic Geology 29 (4), 451–472.Štípská, P., Chopin, F., Skrzypek, E., Schulmann, K., Lexa, O., Pitra, P., Martelat,J., Bollinger, C., 2011. Juxtaposition <strong>of</strong> eclogite-facies <strong>and</strong> mid-crustal rocks duringexhumation: relative role <strong>of</strong> erosion <strong>and</strong> crustal scale folding in a continental wedge(Orlica-Śnieżnik dome, Bohemian Massif). Journal <strong>of</strong> Metamorphic Geology, submitted.Štípská, P., Schulmann, K., Kröner, A., 2004. Vertical extrusion <strong>and</strong> middle crustalspreading <strong>of</strong> omphacite granulite: a model <strong>of</strong> syn-convergent exhumation (BohemianMassif, Czech Republic). Journal <strong>of</strong> Metamorphic Geology 22 (3), 179–198.Štípská, P., Schulmann, K., Powell, R., 2008. Contrasting metamorphic histories <strong>of</strong> lenses<strong>of</strong> high-pressure rocks <strong>and</strong> host migmatites with a flat orogenic fabric (BohemianMassif, Czech Republic): a result <strong>of</strong> tectonic mixing within horizontal crustal flow?Journal <strong>of</strong> Metamorphic Geology 26 (6), 623–646.Van den Driessche, J., Brun, J.-P., 1992. Tectonic evolution <strong>of</strong> the Montagne Noire(French Massif Central): a model <strong>of</strong> extensional gneiss dome. Geodinamica Acta 5,85–99.V<strong>and</strong>erhaeghe, O., Teyssier, C., 2001. Partial melting <strong>and</strong> flow <strong>of</strong> orogens. Tectonophysics342 (3-4), 451–472.Závada, P., Schulmann, K., Konopásek, J., Ulrich, S., Lexa, O., 2007. Extreme ductility<strong>of</strong> feldspar aggregates–melt-enhanced grain boundary sliding <strong>and</strong> creep failure: rheologicalimplications for felsic lower crust. Journal <strong>of</strong> Geophysical Research, B, SolidEarth <strong>and</strong> Planets 112, B10210, doi:10.1029/2006JB004820.Závada, P., Schulmann, K., Lexa, O., Hrouda, F., Haloda, J., Týcová, P., 2009. Themechanism <strong>of</strong> flow <strong>and</strong> fabric development in mechanically anisotropic trachyte lava.Journal <strong>of</strong> Structural Geology 31 (11), 1295–1307.


Accompanying publications47


Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392www.elsevier.com/locate/jstrugeoStructural evolution <strong>of</strong> the central part <strong>of</strong> the KrusÏne hory Erzgebirge)Mountains in the Czech RepublicÐevidence for changing stress regimeduring Variscan compressionJirÏõ KonopaÂsek a,b, *, Karel Schulmann a , Ondrej Lexa aa Charles University, Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Faculty <strong>of</strong> Science, Albertov 6, 128 43, Praha 2, Czech Republicb Geophysical Institute, Czech Academy <strong>of</strong> Science, BocÏnõ II/1401, 141 31, Prague 4, Czech RepublicReceived 12 July 2000; revised 17 December 2000; accepted 19 December 2000AbstractIn the central part <strong>of</strong> the KrusÏne hory Erzgebirge) Mountains, the parautochthonous metasedimentary basement has been overthrust by acrustal nappe <strong>of</strong> ®ne- <strong>and</strong> coarse-grained orthogneisses. The thrust boundary is de®ned by the presence <strong>of</strong> ma®c eclogites with preservedsubduction-related fabric. Westward thrusting <strong>of</strong> an allochthonous unit is associated with the development <strong>of</strong> the main metamorphic foliation<strong>and</strong> lineation in non-eclogitic lithologies. Buttressing <strong>of</strong> the allochthonous body fromthe west is responsible for the development <strong>of</strong> late,large-scale folds with N±S trending hinges <strong>and</strong> vertical axial planes. Subsequent N±S compression leads to large-scale folding <strong>of</strong> both theparautochthonous <strong>and</strong> allochthonous units. This deformation produces km-scale antiforms with hinges plunging to the west <strong>and</strong> is associatedwith the development <strong>of</strong> the E±W stretching lineation as a result <strong>of</strong> complete reworking <strong>of</strong> earlier fabric in the limb zones. N±S shortening isalso associated with the development <strong>of</strong> small-scale folds <strong>and</strong> brittle-ductile kink b<strong>and</strong>s suggesting a decrease in temperature, <strong>and</strong>, thus, uplift<strong>of</strong> the whole studied area during this event. The last stage <strong>of</strong> deformation is characterised by the development <strong>of</strong> kink-b<strong>and</strong> folds <strong>and</strong> acrenulation cleavage. These structures suggest a sub-vertical direction <strong>of</strong> principal compression, developed exclusively in those parts <strong>of</strong> thearea in which the N±S compression produced steep planar fabric. q 2001 Elsevier Science Ltd. All rights reserved.1. IntroductionIn the western part <strong>of</strong> the Saxothuringian domain <strong>of</strong> theBohemian Massif, the identi®cation <strong>of</strong> the major tectonicunits is relatively easy because <strong>of</strong> considerable metamorphiccontrast between high-grade crystalline rocks <strong>and</strong>adjacent low-grade metasediments. Here, allochthonoushigh-grade crystalline units represented by the MuÈnchberg,Frankenberg <strong>and</strong> Wildenfels klippens are thrust over lowgradeCarboniferous sediments Franke et al., 1995).Another important <strong>structural</strong> pattern is associated withexhumation <strong>of</strong> the Saxonian granulites. The emplacement<strong>of</strong> this body into the upper crust is interpreted in terms <strong>of</strong> theextension associated with the development <strong>of</strong> a metamorphiccore complex Franke, 1993).In the eastern part <strong>of</strong> the Saxothuringian domain central<strong>and</strong> eastern KrusÏne hory Mountains), the identi®cation <strong>of</strong>the major <strong>structural</strong> units is more dif®cult because <strong>of</strong> themedium- to high-grade metamorphism affecting both theallochthonous <strong>and</strong> autochthonous units. Moreover, at least* Corresponding author. Fax: 1420-2-2195-2238.E-mail address: kony@natur.cuni.cz J. KonopaÂsek).two periods <strong>of</strong> high-grade metamorphism have affected thepre-Palaeozoic basement in this area, <strong>and</strong> it is dif®cult, if notimpossible, to separate each other in the ®eld MlcÏoch <strong>and</strong>Schulmann, 1992; KroÈner et al., 1995). The tectonics <strong>of</strong> theKrusÏne hory Mountains is commonly interpreted as a result<strong>of</strong> a large-scale westward oriented continental collisionMatte et al., 1990). This thrusting event is documentedby the presence <strong>of</strong> a ¯at foliation, an E±W trending lineation<strong>and</strong> commonly observed westward directed kinematicsRajlich, 1987; Matte et al., 1990; MlcÏoch <strong>and</strong> Schulmann,1992). Recently published studies show that this thrustingwas associated with the emplacement <strong>of</strong> a large-scalecrustal nappe over the Saxothuringian parautochthonKlaÂpova et al., 1998; Krohe, 1998). The main argumentfor the presence <strong>of</strong> an allochthonous unit is the widespreadoccurrence <strong>of</strong> ma®c eclogites surrounded by non-eclogiticrock assemblages KonopaÂsek, 1998; RoÈtzler et al., 1998;KlaÂpova et al., 1998; SchmaÈdicke et al., 1992). However,the exact boundary between the parautochthonous <strong>and</strong>allochthonous units is poorly documented. Closerexamination <strong>of</strong> mapped geological structures Hoth et al.,1994), as well as a detailed ®eld <strong>structural</strong> survey, showsthat the E±W thrusting is not the only Variscan tectonic0191-8141/01/$ - see front matter q 2001 Elsevier Science Ltd. All rights reserved.PII: S0191-814101)00003-749


1374J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392Fig. 1. a) Location <strong>of</strong> the KrusÏne hory Mountains in the European Variscides black square in the northern part <strong>of</strong> the Bohemian Massif). b) Simpli®edgeological map <strong>of</strong> the central part <strong>of</strong> the KrusÏne hory Mountains with the studied area outlined by a black line.event responsible for the building <strong>of</strong> the ®nal geologicalpattern.In this paper we document that the <strong>structural</strong> fabric associatedwith emplacement <strong>of</strong> a crustal nappe is later affectedby an important N±S compression, which is responsible forregional-scale refolding <strong>of</strong> a previously developed ¯at lyingfoliation. We discuss the succession <strong>of</strong> deformational eventsin conjunction with the thermal evolution <strong>and</strong> changingstress regimes through time. We also correlate small-scalestructures <strong>and</strong> observed ®nite strain ellipsoids with stressesoperating during the crustal-scale folding associated withexhumation <strong>of</strong> the middle crust.2. Geological settingThe Bohemian Massif represents the easternmost crystallinecomplex <strong>of</strong> the European Variscides. Its western part iscomposed <strong>of</strong> the Saxothuringian domain, which has beenalready recognised by Kossmat 1927) as a unit with adistinct lithological <strong>and</strong> stratigraphic evolution with respectto the more easterly lying unitsÐthe Tepla±Barr<strong>and</strong>ian <strong>and</strong>the Moldanubian domains.The area studied is situated in the central part <strong>of</strong> theKrusÏne hory Mountains Erzgebirge), which representthe easternmost termination <strong>of</strong> the Saxothuringian domain.50


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 1375Fig. 2. Synoptic geological map <strong>of</strong> the studied area simpli®ed after Sattran 1967) <strong>and</strong> Hoth et al. 1994)). The inset represents lithotectonic column. Solid black lines show the position <strong>of</strong> cross-sectionspresented in Fig. 3.51


1376J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392Fig. 3. a)±d) Cross-sections over the studied area with D2±D3 <strong>structural</strong> data represented by the lower-hemisphere equal area projection. Data are contoured at 1 £ uniformdistribution. See Fig. 2 for theposition <strong>of</strong> cross-sections.52


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 1377Fig. 3. continued)The KrusÏne hory Mountains forman antiform-like structurewith Proterozoic metamorphic rocks in its core <strong>and</strong> with aPalaeozoic metasedimentary cover. The age <strong>of</strong> metamorphismwas thought to be dominantly Variscan withdecreasing metamorphic conditions from the east to thewest Kossmat, 1925) Fig. 1).The core <strong>of</strong> this structure is represented mainly by amonotonous complex <strong>of</strong> micaschists with subordinate metagreywackesthe Osterzgebirge complex) <strong>and</strong> bymetapelites, orthogneisses <strong>and</strong> numerous bodies <strong>of</strong> eclogitesin the uppermost part the PrÏõÂsecÏnice complex) H<strong>of</strong>mann etal., 1988). The metasedimentary cover consists <strong>of</strong> quartzites,micaschists <strong>and</strong> phyllites <strong>of</strong> Lower Paleozoic ageHoth et al., 1979). This subdivision is based exclusivelyon lithological arguments, however, <strong>and</strong> no <strong>structural</strong> <strong>and</strong>metamorphic features <strong>of</strong> the area were considered.MlcÏoch <strong>and</strong> Schulmann 1992) suggested a Cadomianage <strong>of</strong> metamorphism for anatectic orthogneisses in thecentral part <strong>of</strong> the KrusÏne hory Mountains, which wereintruded by a large late-Cadomian porphyritic granite. ANeo-proterozoic age for this pluton 554 ^ 10 MaÐU/Pbzircon method), as well as that <strong>of</strong> surrounding migmatites,was obtained by KroÈner et al. 1995). The Cadomian structurewas subsequently reworked by the Variscan deformation<strong>and</strong> metamorphism that dominates in the entire areae.g. SchmaÈdicke et al., 1995; Kotkova et al., 1996; Werneret al., 1997; KroÈner <strong>and</strong> Willner, 1998).PT data fromall major lithologies suggest a high dP/dTgradient during Variscan metamorphism SchmaÈdicke et al.,1992; Holub <strong>and</strong> SoucÏek, 1994; KlaÂpova et al., 1998;KonopaÂsek, 1998; RoÈtzler et al., 1998). In the Germanpart <strong>of</strong> the KrusÏne hory Mountains, Willner et al. 1994)have proposed three major high-pressure HP) units, withdecreasing metamorphic grade from the lowermost to theuppermost unit. Krohe 1996) interpreted this feature as aresult <strong>of</strong> ductile extension during exhumation <strong>of</strong> thethickened Saxothuringian domain. All these units containma®c eclogites <strong>and</strong> some <strong>of</strong> them HP granulites Willner53


1378J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392et al., 1997) with peak PT conditions differing considerablyfrom those <strong>of</strong> the adjacent metasediments RoÈtzler et al.,1998).Metamorphic conditions in the study area are similar tothose de®ned in Germany. A high dP/dT gradient wasobserved in metasedimentary rocks KonopaÂsek, 1998,2001) which host ma®c eclogites. The peak PT conditions<strong>of</strong> the eclogites are not consistent with those <strong>of</strong> the adjacentparautochthonous metasediments, however, suggesting thatthey have been juxtaposed tectonically KlaÂpova et al.,1998).3. Lithological zonation <strong>of</strong> the studied areaThe area studied involves four main large-scale structuresexposed in the vicinity <strong>of</strong> the village <strong>of</strong> MeÏdeÏnecÐtheMeÏdeÏnec synform, the Oberwiesenthal structure, theMeÏdeÏnec antiform<strong>and</strong> the KlõÂnovec antiformSÏkvor,1975) Fig. 2). These large structures are composed <strong>of</strong> thefollowing rocks:1. Plagioclase schists are described by KonopaÂsek 1998)as characteristic metasediments forming the Saxothuringianparautochthon. Typical samples containreversely zoned plagioclase porphyroblasts envelopingnumerous garnet inclusions. As these rocks show a polyphasemetamorphic history, they can be sampled atdifferent stages <strong>of</strong> their evolution KonopaÂsek, 1998).These rocks are intercalated with metagreywackes'Dichte Gneisse' <strong>of</strong> Pietzsch 1914)) <strong>and</strong> metaconglomeratesMehnert, 1939; Sattran, 1963).2. Orthogneisses appear either as ®ne-grained equigranularvarieties or as deformed porphyritic granite 'RoteGneisse' <strong>of</strong> Scheumann 1935)). Porphyriticorthogneisses are heterogeneously deformed <strong>and</strong> occurin all deformation stages ranging from protomyloniticmetagranites up to b<strong>and</strong>ed ultramylonites.3. Garnetiferous micaschists are characterised by theappearance <strong>of</strong> numerous garnet porphyroblasts, sometimesup to 2 cm in diameter, surrounded by a whitemica <strong>and</strong> quartz matrix. In the southern limb <strong>of</strong> theKlõÂnovec antiform, garnetiferous micaschists are associatedwith a layer <strong>of</strong> amphibolites Fig. 2).4. Ma®c eclogites bear a typical HP mineral assemblageconsisting <strong>of</strong> omphacite 1 garnet 1 zoisite 1amphibole 1 rutile 1 quartz <strong>and</strong> h<strong>and</strong> specimens arecharacterised by well-developed planar <strong>and</strong> linear fabricKlaÂpova et al., 1998).The rock-types described above occur in all the largescalestructures, but the lithological zonation within thesestructures is not uniform. In the MeÏdeÏnec synform, plagioclaseschists with metagreywackes represent <strong>structural</strong>ly thedeepest level. These schists pass upward into the layer <strong>of</strong>garnetiferous micaschists associated with the ma®ceclogites. In the hanging wall <strong>of</strong> garnetiferous micaschist,there occurs a large slab <strong>of</strong> orthogneisses with a ®ne-grainedvariety at the bottom<strong>and</strong> a porphyritic type at the top Fig.3aÐnorthern part <strong>of</strong> the pro®le 1±1 0 ). Locally, fragments<strong>of</strong> the allochthonous orthogneiss slab can be observed alsowithin the parautochthonous micaschists Fig. 3b). Thelithological zonation <strong>of</strong> the KlõÂnovec antiformis morecomplicated. In its eastern part, the zonation <strong>of</strong> the southernlimb is the same as that in the MeÏdeÏnec synformFig. 3cÐsouthern part <strong>of</strong> the pro®le 3±3 0 ). However, farther to thewest, the lithotectonic zonation becomes progressivelyreversed with orthogneisses in the lowermost positionpassing in the garnetiferous micaschists <strong>and</strong> with plagioclaseschists in the uppermost position Fig. 3dÐsouthernpart <strong>of</strong> the pro®le 4±4 0 ). Moreover, in the core <strong>of</strong> theKlõÂnovec antiform, large bodies <strong>of</strong> eclogites are situated inthe centre <strong>of</strong> the orthogneiss body Fig. 2).Field observations show that ma®c eclogites, togetherwith a layer <strong>of</strong> garnetiferous micaschists, are in mostcases exposed along the boundary between the orthogneisses<strong>and</strong> plagioclase schists. The location <strong>of</strong> these unitsalong this boundary is critical for underst<strong>and</strong>ing the tectonicevolution <strong>of</strong> the studied area.4. PT evolution <strong>of</strong> the studied area <strong>and</strong> the de®nition <strong>of</strong>the allochthonous <strong>and</strong> parautochthonous domainsThe PT conditions <strong>of</strong> orthogneiss formation are notknown, but the stability <strong>of</strong> plagioclase in all the studiedsamples excludes the possibility <strong>of</strong> them being deformedunder eclogite facies conditions. The peak pressure conditions<strong>of</strong> the plagioclase schists were established to be 13±15 kbar at a temperature <strong>of</strong> 580±6308C KonopaÂsek, 1998).These pressure conditions are in contrast with those reportedby KlaÂpova et al. 1998) fromthe ma®c eclogites wherepeak pressures were estimated to be 26 kbar at 650±7008C. The PT conditions <strong>of</strong> the associated garnetiferousmicaschists approach those <strong>of</strong> the eclogites 6408C,22 kbarÐ KonopaÂsek, 2001). The difference between PTestimates from parautochthonous metasediments <strong>and</strong> eclogiteswas explained in terms <strong>of</strong> emplacement <strong>of</strong> previouslysubducted oceanic crust into continental rocks during theearly stages <strong>of</strong> collision KlaÂpova et al., 1998).As noted earlier, the ma®c eclogite bodies associated withthe garnetiferous micaschists occur systematically at theboundary between plagioclase schists <strong>and</strong> orthogneisses.This observation suggests that this boundary represents amajor crustal boundary along which ma®c eclogies wereexhumed <strong>and</strong> incorporated into the middle crust. Accordingto the interpretation <strong>of</strong> KonopaÂsek 1998) <strong>and</strong> KlaÂpova et al.1998), the plagioclase schists represent a Saxothuringianparautochthon, which was overthrust by middle-crustalorthogneisses derived froman orogenic root domain,today exposed in the eastern Moldanubian zone. Based onthis interpretation, these allochthonous orthogneisses,54


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 1379together with ma®c eclogites <strong>and</strong> garnetiferous micaschistswill be referred to as the Lower Crystalline Nappe.5. The succession <strong>of</strong> deformation structures D1±D3)A polyphase tectonic evolution has resulted in four D1±D4) stages <strong>of</strong> deformation. The structures developed in theeclogites have been described by KlaÂpova et al. 1998) <strong>and</strong>are, therefore, not examined in detail here, except whennecessary for the interpretation <strong>of</strong> a particular deformationstage.5.1. D1 structuresThe D1 structures are present exclusively in the eclogites<strong>and</strong> consist <strong>of</strong> syn-metamorphic S 1 foliation <strong>and</strong> an L 1lineation. As described by KlaÂpova et al. 1998), the S 1metamorphic foliation is mainly the result <strong>of</strong> a planarorientation <strong>of</strong> omphacite <strong>and</strong> a metamorphic layeringcharacterised by an alternation <strong>of</strong> omphacite-rich layerswith layers rich in garnet. The L 1 stretching lineation ischaracterised by a shape-preferred orientation <strong>of</strong> omphacitecrystals. The orientation <strong>of</strong> D1 structures in eclogiticboudins varies according to their position within largescalestructures over the studied area see KlaÂpova et al.1998) for details).5.2. D2 structuresThe D2 deformation in eclogites is marked by thedevelopment <strong>of</strong> asymmetric internal foliation boudinagewith neck zones ®lled with quartz, rutile, amphibole,paragonite <strong>and</strong> zoisite. Brittle cracks up to several meterslong have developed; they are either closed or ®lled with thesame assemblage as the neck zones KlaÂpova et al., 1998).The D2 structures in non-eclogitic lithologies can be bestobserved in the northern part <strong>of</strong> the studied area in theOberwiesenthal <strong>and</strong> MeÏdeÏnec synforms) where the rocksare only slightly affected by the D3 deformation. Largedomains <strong>of</strong> D2 fabric only slightly affected by D3 deformationalso appear in the basement metasediments exposedsouth <strong>of</strong> the KlõÂnovec antiform.In the orthogneisses, the S 2 foliation is characterised bythe alternation <strong>of</strong> recrystallised quartz <strong>and</strong> feldspar ribbonswith phyllosilicate rich domains, <strong>and</strong> in metasedimentsmainly by preferred orientation <strong>of</strong> micas <strong>and</strong> by alternation<strong>of</strong> mica-rich <strong>and</strong> quartz-rich layers. The orientation <strong>of</strong> the S 2foliation in the MeÏdeÏnec synformis variable but it generallystrikes N±S <strong>and</strong> dips gently between 10 <strong>and</strong> 208. In theOberwiesenthal structure, the S 2 foliation is ¯at lying inits northern part <strong>and</strong> becomes steeply inclined as onemoves south. The L 2 stretching <strong>and</strong> mineral lineation isde®ned by an alignment <strong>of</strong> quartz <strong>and</strong> feldspar aggregates<strong>and</strong> by a stretching <strong>of</strong> quartz-rich aggregates in theorthogneisses, <strong>and</strong> as a preferred arrangement <strong>of</strong> micas inthe metasediments. The orientation <strong>of</strong> the L 2 lineation isgenerally ESE±WNW Fig. 4).Numerous kinematic indicators such as sigmoidalK-feldspar porphyroclasts <strong>and</strong> S±C fabrics in porphyriticorthogneiss Fig. 5a) <strong>and</strong> shear-b<strong>and</strong>s in metasedimentsare consistent with top-to-the-west oriented shearing. Inseveral places, early F 2 isoclinal, recumbent folds wereobserved in basement rocks <strong>of</strong> semipelitic compositionFig. 5b). These early folds refold the primary compositionalb<strong>and</strong>ing S 0 <strong>of</strong> the metasediments leading to itscomplete transposition into a penetrative metamorphicfabric. The orientation <strong>of</strong> early F 2 fold hinges is E±W,being parallel to the L 2 mineral lineation Fig. 4).During late stages <strong>of</strong> the D2 deformation, the wholeassembled sequence was folded into large-scale periclinalstructures further described as the late F 2 folds) withsubvertical, N±S trending axial planes. These late F 2periclines formthe present day MeÏdeÏnec synform<strong>and</strong> theOberwiesenthal structure.5.3. D3 structuresIn the southern termination <strong>of</strong> the MeÏdeÏnec synform, theS 2 fabric dips gently to the NW. As one moves south, theorientation <strong>of</strong> the metamorphic fabric <strong>and</strong> the lithologicalbodies rotate so that the foliation dips steeply to the southFigs. 3a <strong>and</strong> 4). This geometry suggests that the gentlydipping planar fabric <strong>of</strong> the MeÏdeÏnec synformis refoldedby the large MeÏdeÏnec antiformwith the hinge zone plungingto the west at a shallow angle. This fold is asymmetricalwith a steeply dipping southern limb, a gently dippingnorthern limb <strong>and</strong> a hinge zone plunging gently to the west.A similar pattern is developed in the Oberwiesenthalstructure <strong>and</strong> the KlõÂnovec antiform. The Oberwiesenthalstructure has, in its eastern part, a similar geometry to theMeÏdeÏnec synform<strong>and</strong> shows progressive decrease <strong>of</strong> theinterlimb angle to the south Fig. 4). The connectionbetween the Oberwiesenthal synform<strong>and</strong> the KlõÂnovec antiformiseroded <strong>and</strong> the later structure represents, as in thecase <strong>of</strong> the MeÏdeÏnec structures, a former N±S trendingsynclinal structure rotated into an E±W direction Fig. 2).In the E±W trending limb, the foliation is subverticalFig. 5c) <strong>and</strong> the hinge <strong>of</strong> this large fold is steeply plungingto the SW. This is documented by the linear fabric <strong>of</strong> ma®ceclogites, which plunges steeply in the same direction. Theappearance <strong>of</strong> eclogite bodies in the centre <strong>of</strong> the KlõÂnovecantiform, as well as double thickness <strong>of</strong> the ®ne-grainedorthogneisses in the northern zone <strong>of</strong> the KlõÂnovec antiformFig. 2) suggest that the northern part <strong>of</strong> this E±W trendinglimb was thickened during the the D2 thrusting episode.The mineral lineation in the Oberwiesenthal structure, theMeÏdeÏnec synform<strong>and</strong> the MeÏdeÏnec antiformshows auniformWNW±ESE orientation Fig. 4), consistent withother parts <strong>of</strong> the KrusÏne hory Mountains parautochthone.g. MlcÏoch <strong>and</strong> Schulmann, 1992). In the KlõÂnovecantiform, the mineral <strong>and</strong> stretching lineation <strong>of</strong> the55


1380J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392Fig. 4. Structural map <strong>of</strong> the studied area showing trends <strong>of</strong> D1, D2 <strong>and</strong> D3 planar <strong>and</strong> linear structures.56


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 1381Fig. 5. a) S±C fabric in porphyritic orthogneiss <strong>of</strong> the Lower Crystalline nappe. This fabric originated during D2 <strong>and</strong> suggests top-to-the-west oriented sense<strong>of</strong> movement. b) Isoclinal F 2 fold in metagreywackes <strong>of</strong> the parautochthonous unit. c) Vertical S 2 foliation in the southern limb <strong>of</strong> the KlõÂnovec antiform. Thisvertical fabric resulted fromrefolding <strong>of</strong> originally ¯at S 2 foliation during D3. d) F 3 fold with steep axial plane in the autochthonous plagioclase schists south<strong>of</strong> the KlõÂnovec antiform. e) Conjugate system <strong>of</strong> D3 kink b<strong>and</strong>s affecting the S 2 foliation in the hinge zone <strong>of</strong> the MeÏdeÏnec antiform. f) Late F 4 folds withsubhorizontal axial planes in amphibolite <strong>of</strong> the southern limb <strong>of</strong> the KlõÂnovec antiform.orthogneisses is systematically plunging gently to the WFig. 4), whereas in the eclogites, the omphacite L 1 lineationis gently plunging to the W±WSW in the central part <strong>of</strong> theKlõÂnovec antiform<strong>and</strong> steeply to the SW in the hinge zone.Numerous D3 fold structures up to several metres in sizewith E±W trending subhorizontal hinge zones <strong>and</strong> steep tointermediate, mostly northward dipping axial planes occurin the parautochthonous plagioclase schists south <strong>of</strong> theKlõÂnovec antiformFigs. 4 <strong>and</strong> 5d).The D3 deformation is most intense in the parautochthonousplagioclase schists in the area between the northernlimb <strong>of</strong> the KlõÂnovec antiform<strong>and</strong> the southern limb <strong>of</strong>57


1382J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392Fig. 6. Schematic block-diagram <strong>of</strong> the studied area. Flat lying S2 foliation can be observed in the northern part <strong>of</strong> the studied area in the MeÏdeÏnec synform. Going to the south, the early D2 fabric is completelyreworked by D3 deformation. The Oberwiesenthal structure is not shown.58


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 1383the MeÏdeÏnec antiform. In this area, the S 2 fabric is completelyreworked into a steep to subvertical E±W trendingS 3 foliation. Here, a strong sub-horizontal stretching <strong>of</strong>quartz aggregates can be observed in several places. In thehinge zone <strong>of</strong> the MeÏdeÏnec antiform, monoclinal or conjugatekink b<strong>and</strong>s are developed affecting the early S 2 metamorphicfoliation <strong>of</strong> the garnetiferous micaschist <strong>and</strong>orthogneisses. These kink b<strong>and</strong>s exhibit E±W trendingaxes <strong>and</strong> kink planes, bimodally distributed with respectto ¯at-lying S 2 foliation Fig. 5e). The geometry <strong>of</strong> thesekink b<strong>and</strong>s suggests that they are genetically linked to theMeÏdeÏnec antiformal D3 structure but the mechanism <strong>and</strong> thekinematics <strong>of</strong> kinking will be discussed in a separate sectionbelow.The general structure <strong>of</strong> the studied domain is shown inblock-diagram Fig. 6), summarising lithological <strong>and</strong> <strong>structural</strong>observations discussed above as an effect <strong>of</strong> the D3folding superimposed on the D2 fabrics.6. Problem <strong>of</strong> the L 2 <strong>and</strong> L 3 lineation solved usingaggregate-shape analysisAn important feature is the parallelism <strong>of</strong> the L 2 minerallineation in the MeÏdeÏnec synforma D2 structure) with thelineation observed in micaschists <strong>and</strong> orthogneisses <strong>of</strong> boththe MeÏdeÏnec <strong>and</strong> the KlõÂnovec antiforms. If the WNW±ESEtrending L 2 lineation in the MeÏdeÏnec synformwere rotatedby late F 2 folding, the originally EW trending, gentlyplunging L 2 lineation would become subvertical Fig. 7).This can be observed in N±S trending, steep S 2 foliationin the southern part <strong>of</strong> the Oberwiesenthal structure. Subsequentrotation <strong>of</strong> the S 2 foliation by F 3 folds with a steephinge zone will keep the orientation <strong>of</strong> the L 2 in a subverticalposition Fig. 7) . The lineation in the steep limbs<strong>of</strong> the KlõÂnovec <strong>and</strong> the MeÏdeÏnec antiforms is always foundto plunge gently to the west, however, indicating that it wasformed during the D3 deformation L 3 lineation). Consequently,the change <strong>of</strong> the L 2 into an L 3 lineation must beassociated with complete modi®cation <strong>of</strong> the D2 fabricduring the D3 deformation. In order to investigate theproposed changes to the D2 fabric ellipsoid during the D3deformation, we have carried out a shape analysis <strong>of</strong> mineralaggregates in the orthogneisses fromthe MeÏdeÏnec synform<strong>and</strong> the MeÏdeÏnec <strong>and</strong> the KlõÂnovec antiforms.In order to determine the role <strong>of</strong> the F 3 folding inpossible fabric modi®cations, an aggregate-shape analysis<strong>of</strong> K-feldspar <strong>and</strong> quartz was carried out on the porphyriticorthogneisses. Selected porphyritic augen orthogneisseswere sampled in the hinge zones <strong>and</strong> limbs <strong>of</strong> both theMeÏdeÏnec <strong>and</strong> the KlõÂnovec antiforms, as well as in theMeÏdeÏnec synform<strong>and</strong> the Oberwiesenthal structure. Twosections in each sample were studied: a) perpendicular tometamorphic foliation <strong>and</strong> parallel to mineral lineation theXZ section <strong>of</strong> the ®nite strain ellipsoid), <strong>and</strong> b) perpendicularto both foliation <strong>and</strong> lineation the YZ section<strong>of</strong> the ®nite strain ellipsoid). The shapes <strong>of</strong> the K-feldspar<strong>and</strong> quartz aggregates were traced on transparent sheet,digitised <strong>and</strong> then analysed. Because <strong>of</strong> the large size <strong>of</strong>measured aggregates, only a limited number <strong>of</strong> measurementscould be carried out on each sample Table 1). Finitestrain ratios were calculated fromthe two principal strainellipsoid planes using a harmonic mean method Lisle,1977; Ramsay <strong>and</strong> Huber, 1983, p. 80). These ratios wereFig. 7. Succession <strong>of</strong> the lower-hemisphere equal area projections shows schematically expected evolution <strong>of</strong> the orientation <strong>of</strong> the L 2 lineation during D2 <strong>and</strong>D3 folding. Early L 2 lineation is trending E±W being subhorizontal. During late F 2 folding, the L 2 lineation does not change the orientation, but becomessteeper. The successive F 3 folding changes the orientation <strong>of</strong> steep S 2 foliation; the L 2 lineation should rotate into N±S direction maintaining its steep plunge.In the D3 structures the MeÏdeÏnec <strong>and</strong> the KlõÂnovec antiforms), however, the lineation is always E±W oriented <strong>and</strong> mostly subhorizontal.59


1384J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392Table 1Harmonic means <strong>of</strong> measured X/Z <strong>and</strong> Y/Z ratios <strong>and</strong> corresponding calculated X/Y ratios <strong>and</strong> k <strong>and</strong> d values for orthogneiss samples from the D2 domains theMeÏdeÏnec synform<strong>and</strong> the Oberwiesenthal structure) <strong>and</strong> the D3 domains the KlõÂnovec <strong>and</strong> MeÏdeÏnec antiforms). See Fig. 8 for corresponding Flinn's graph<strong>and</strong> the d/k plotSample Locality Object X/Z Y/Z X/Y k d Number <strong>of</strong> objects measured:in XZin YZD2MR10 MeÏdeÏnec synformQtz 5.907 2.26 2.614 1.28 2.047 18 25Kf 3.438 1.372 2.507 4.055 1.552 4 7MR7 MeÏdeÏnec synformQtz 4.914 1.865 2.634 1.888 1.849 24 35Kf 2.525 1.831 1.379 0.456 0.914 8 16MR8 MeÏdeÏnec synformQtz 4.594 2.34 1.963 0.719 1.65 22 21Kf 1.752 1.285 1.364 1.276 0.462 7 4OW407 Oberwiesenthal str. Qtz 3.614 3.033 1.192 0.094 2.042 11 9Kf 3.794 2.234 1.699 0.566 1.418 12 14OW417 Oberwiesenthal str. Kf 3.891 2.372 1.64 0.466 1.514 13 13D3MR115 KlõÂnovec antiformQtz 5.271 2.955 1.783 0.401 2.106 10 9Kf 3.171 1.913 1.657 0.72 1.125 10 8MR11 MeÏdeÏnec antiformKf 4.209 2.488 1.692 0.465 1.641 21 49MR12 MeÏdeÏnec antiformKf 10.45 9.001 1.161 0.02 8.002 20 23MR3 KlõÂnovec antiformQtz 10.26 6.301 1.628 0.118 5.338 23 13Kf 5.487 3.194 1.718 0.327 2.309 15 11MR401 KlõÂnovec antiformQtz 12.33 8.385 1.47 0.064 7.399 10 14Kf 6.961 5.168 1.347 0.083 4.182 12 13used to construct the k-parameter <strong>of</strong> Flinn 1962) <strong>and</strong>d-parameter e.g. Ramsay <strong>and</strong> Huber, 1983, p. 202) toestimate the shape <strong>of</strong> the strain ellipsoid <strong>and</strong> strain intensity,respectively.The ®nite strain analysis shows that the augenorthogneisses in the MeÏdeÏnec synform, as well as those inthe hinge zone <strong>of</strong> the MeÏdeÏnec antiform, exhibit ellipsoidsranging fromconstrictional to plane-strain or slightly oblatesymmetry, k-parameters vary from 0.72 to 1.89 for quartz<strong>and</strong> 0.46 to 4.06 for feldsparsÐFig. 8 <strong>and</strong> Table 1). Exceptionallyoblate fabric was observed for quartz aggregates inthe sample OW407 k ˆ 0.094). Quartz generally showshigher strain intensity d-parameters from 1.65 to 2) thanfeldspar d-parameters from 0.9 to 1.55). In studied samples,the orthogneisses have a character <strong>of</strong> augen orthogneiss withwell preserved K-feldspar porphyroclasts. In the southern,steep limb <strong>of</strong> the MeÏdeÏnec antiform, <strong>and</strong> in the orthogneissbodies <strong>of</strong> the KlõÂnovec antiform, the shapes <strong>of</strong> strainellipsoids are oblate k ˆ 0.4±0.06 for quartz <strong>and</strong>k ˆ 0.72±0.02 for K-feldsparÐFig. 8 <strong>and</strong> Table 1). As inthe previous case, quartz generally shows higher strainintensity d-parameter is from 2.1 to 7.4) than feldspard-parameter from 1.1 to 8). The increase <strong>of</strong> deformationintensity is connected with the disappearance <strong>of</strong> augenstructure <strong>and</strong> development <strong>of</strong> b<strong>and</strong>ed orthogneiss. It isnoted that samples with weak oblate symmetry show welldevelopedL 3 lineation characterised by stretching <strong>of</strong> feldspar<strong>and</strong> quartz aggregates. In contrast, samples with strongoblate symmetry do not show any stretching lineation <strong>and</strong>the L 3 fabric is characterised only by alignment <strong>of</strong> whitemica in the foliation plane.We suggest in agreement with the mesoscopic <strong>structural</strong>data) that the symmetry <strong>of</strong> the D2 deformation is representedby a plane strain to constrictional or slightly oblatestrain ellipsoid resulting fromwestward directed noncoaxialshearing. This symmetry is preserved in both theMeÏdeÏnec synform<strong>and</strong> the Oberwiesenthal structure <strong>and</strong>probably also in the hinge zone <strong>of</strong> the MeÏdeÏnec antiform.In contrast, the shapes <strong>of</strong> K-feldspar <strong>and</strong> quartz aggregatesin the limbs <strong>of</strong> the KlõÂnovec antiformshow oblate geometry,high strain intensity <strong>and</strong> X-axes oriented E±W. Asmentioned above, this type <strong>of</strong> geometry cannot be producedby passive rotation <strong>of</strong> the D2 fabric during the F 3 folding <strong>and</strong>we propose that the large-scale F 3 folding is responsible forthe modi®cation <strong>of</strong> the shapes <strong>of</strong> the D2 strain ellipsoid. Theaggregate shape analysis also shows that the large scale F 3folding occurred under still elevated thermal conditions,which allowed ductile deformation <strong>of</strong> feldspar clasts.7. Brittle-ductile structures late D3 <strong>and</strong> D4)As reported by KlaÂpova et al. 1998), the D4 deformationin the eclogites resulted in the development <strong>of</strong> two main sets<strong>of</strong> structures. Most frequently, the D4 deformation produceslate S 4 retrograde shear zones, which crosscut the S 1eclogitic foliation at high angles. In the thick eclogiteboudins, the D4 deformation has resulted in the refolding<strong>of</strong> the early S 1 foliation. There, D4 structures are open toclosed recumbent, asymmetrical F 4 folds. The D4 deformationis best recorded in amphibolites, garnetiferousmicaschists <strong>and</strong> basement plagioclase schists, <strong>and</strong> is60


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 13853X/Y2D2K=1D2D2D3Quartz - D2Quartz - D3K-feldspar - D2K-feldspar - D3D3D3D3dD211 2 3 4 5 6 7 8 910864Y/ZQuartz - D2Quartz - D3K-feldspar - D2K-feldspar - D3200 1 2 3 4 5kFig. 8. Flinn's graph <strong>and</strong> d/k plot obtained from measurements <strong>of</strong> K-feldspar <strong>and</strong> quartz aggregates in coarse-grained orthogneisses <strong>of</strong> the MeÏdeÏnec nappe. Inthe MeÏdeÏnec synform<strong>and</strong> northern part <strong>of</strong> the MeÏdeÏnec antiform D2 structures), both K-feldspar <strong>and</strong> quartz aggregates show plane strain to prolate symmetryat relatively low strain intensities. On the other h<strong>and</strong>, the same aggregates in the southern limb <strong>of</strong> the MeÏdeÏnec antiform<strong>and</strong> in the KlõÂnovec antiformD3structures) show oblate symmetry <strong>and</strong> higher strain intensities see text for methods <strong>and</strong> details). Grey areas represent starting <strong>and</strong> ending positions <strong>of</strong> the D2ellipsoids used for <strong>modelling</strong> <strong>of</strong> the D3 fabric. Results <strong>of</strong> this <strong>modelling</strong> are presented in Fig. 10.represented by late F 4 folds <strong>and</strong> kink b<strong>and</strong>s Fig. 5f). Thelatest deformation in the metasediments is represented bysemi-brittle normal faults dipping to the SW. These occur onthe southern limb <strong>of</strong> the KlõÂnovec antiform. Evidence <strong>of</strong> theD4 deformation in orthogneisses is rare. The D4 structuresappear mostly in the southern limb <strong>of</strong> the KlõÂnovec antiformas discrete sets <strong>of</strong> brittle-ductile S 4 cleavage with homogeneousspacing obliquely oriented to the main S 2 foliationFig. 9).7.1. The D3 <strong>and</strong> D4kink-b<strong>and</strong> folds in metasediments <strong>and</strong>orthogneissesIn several places, conjugate arrays <strong>of</strong> kink-b<strong>and</strong> foldswere observed, but in most cases only one single set isdeveloped monoclinal kink-b<strong>and</strong> folds after Ramsay <strong>and</strong>Huber 1987), p. 427). All these kink-b<strong>and</strong> folds are <strong>of</strong>contractional type Ramsay <strong>and</strong> Huber, 1987, p. 427) <strong>and</strong>their common feature is an E±W orientation <strong>of</strong> their hinges.In the northern limb <strong>of</strong> the MeÏdeÏnec antiformclose to thehinge zone, Fig. 9) several localities with well-developedkink-b<strong>and</strong> folds occur. These appear in both the metasediments<strong>and</strong> the orthogneisses in zones where the S 2 fabricdips gently Loc. 130, 130.V, 6, 138ÐFig. 9). The sametype <strong>of</strong> kink-b<strong>and</strong> folds can be locally observed in zones <strong>of</strong>¯at lying S 2 metamorphic fabric south <strong>of</strong> the KlõÂnovec antiformLoc.307). These kink-b<strong>and</strong> folds, which are eitherconjugate or forma single set with a uniformorientation <strong>of</strong>the kink-plane, are related to the D3 deformation.The D4 kink-b<strong>and</strong> folds are developed exclusively inzones <strong>of</strong> steep S 2 ±S 3 fabric in both the plagioclase schists<strong>and</strong> the garnetiferous micaschists, <strong>and</strong> are mostfrequently observed in close proximity to the allochthonousorthogneiss body. Three groups <strong>of</strong> D4 kink-b<strong>and</strong> folds arerecognised <strong>and</strong> these are shown in Fig. 9. A group a) ismade up from a set <strong>of</strong> monoclinal kink-b<strong>and</strong> folds, whichaffect foliations dipping steeply to the north. They indicatetop-to-the-north normal movement. Group b) consists <strong>of</strong>61


1386J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392Fig. 9. Structural map <strong>of</strong> the studied area shows D3 <strong>and</strong> D4 monoclinal <strong>and</strong> conjugate systems <strong>of</strong> kink b<strong>and</strong>s. Orientation <strong>of</strong> the kink b<strong>and</strong>s is presented in thelower-hemisphere equal area projection together with the orientation <strong>of</strong> S 2 ±S 3 foliation see lower left part <strong>of</strong> the ®gure for principal fabric elements). For eachlocality numbered inside or above each projection), all measured kink b<strong>and</strong>s <strong>and</strong> foliations were averaged <strong>and</strong> presented as a mean value. Estimated s 1 <strong>and</strong> s 3directions are shown at localities with developed conjugate kink b<strong>and</strong>s. The equal area projection on the right h<strong>and</strong> side <strong>of</strong> the ®gure represents poles to brittleductile,late D4 shear b<strong>and</strong>s in orthogneisses <strong>and</strong> micaschists. Three groups <strong>of</strong> F 4 compressional kink b<strong>and</strong>s observed in the studied area are shown in the lowerpart <strong>of</strong> the ®gure. a) Kink b<strong>and</strong>s showing northward normal kinematics affecting northward dipping steep foliation. b) Conjugate set <strong>of</strong> kink b<strong>and</strong>s affectingsubvertical foliation. c) Kink b<strong>and</strong>s showing southward normal kinematics affecting southward dipping steep foliation.symmetrically developed conjugate kink-b<strong>and</strong> folds <strong>and</strong>these are developed in a vertical foliation. Group c)consists <strong>of</strong> monoclinal kink-b<strong>and</strong> folds which deform asteep, south-dipping foliations <strong>and</strong> indicates top-to-thesouthnormal kinematics.In the southern limb <strong>of</strong> the KlõÂnovec antiform, most <strong>of</strong> thekink-b<strong>and</strong> folds belong to group 3, with groups 2 <strong>and</strong> 1observed less frequently. On the southern limb <strong>of</strong> theMeÏdeÏnec antiform, the three groups <strong>of</strong> kink-b<strong>and</strong> folds aremore equally developed.62


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 13877.2. Analysis <strong>of</strong> kink-b<strong>and</strong> foldsThe geometry <strong>of</strong> symmetrical kink-b<strong>and</strong> folds was usedto determine the orientation <strong>of</strong> stress axes according to themethod <strong>of</strong> Ramsay 1962a). The orientation <strong>of</strong> s 1 is perpendicularto the kink-b<strong>and</strong> fold axis <strong>and</strong> lies in the plane thatbisects the obtuse angle between the conjugate kink-b<strong>and</strong>folds. If only one set <strong>of</strong> the conjugate kink-b<strong>and</strong> foldsdevelops, it implies that the maximum principal compressivestress s 1 was oblique to the foliation. The approximateorientation can be deduced but, as the angle betweens 1 <strong>and</strong> the kink-b<strong>and</strong> fold is determined by the value <strong>of</strong> themechanical anisotropy Cobbold et al., 1971) <strong>and</strong> because <strong>of</strong>the phenomenon <strong>of</strong> stress de¯ection, it is not possible todetermine the exact orientation <strong>of</strong> s 1 .The sub-horizontal orientation <strong>of</strong> s 1 , deduced fromthegeometry <strong>and</strong> orientation <strong>of</strong> the D3 kink-b<strong>and</strong> folds, isconsistent with earlier deduced N±S compression associatedwith the D3 deformation. Some <strong>of</strong> them show signs<strong>of</strong> later rotation, which probably occurred during theprogressive development <strong>of</strong> the D3 folding e.g. Locality6 in Fig. 9).The uniformorientation <strong>of</strong> the D4 fold axes <strong>and</strong> theessentially bimodal distribution <strong>of</strong> kink-b<strong>and</strong> folds in thesouthern limb <strong>of</strong> the KlõÂnovec antiformFig. 9) indicatethat they probably represent a conjugate set groups a)<strong>and</strong> c), Fig. 9). This conjugate set is developed throughoutthe whole study area. Locally, only one set <strong>of</strong> the kink-b<strong>and</strong>folds is found to have developed <strong>and</strong>, as discussed above,this can be interpreted as a result <strong>of</strong> variations <strong>of</strong> the S 2 ±S 3orientation with respect to principal compression. The stressanalysis <strong>of</strong> both the monoclinic <strong>and</strong> conjugate kink-b<strong>and</strong>folds shows that the principal compression s 1 during theD4 deformation was subvertical.8. Discussion8.1. Kinematic interpretation <strong>of</strong> the D1±D2 <strong>structural</strong>successionÐbuilding <strong>of</strong> the nappe pileThe D1 fabric in eclogites has developed during a pressure<strong>and</strong> temperature peak in spatially different setting thanthe present-day country rocks. After their emplacement inthe crust, the eclogites were passively transported at the base<strong>of</strong> the Lower Crystalline Nappe during the D2 thrusting <strong>and</strong>behaved as rigid inclusions in a weak matrix <strong>of</strong> micaschists.This kind <strong>of</strong> behaviour resulted in the non-uniformorientation<strong>of</strong> the D1 structures within the isolated eclogiticboudins which were developed mainly during the largescaleD3 folding KlaÂpova et al., 1998).The D2 structures in the parautochthonous plagioclaseschists, <strong>and</strong> in the allochthonous orthogneisses, are associatedwith the main metamorphic event. They representsyn-metamorphic structures developed during thewestward thrusting <strong>of</strong> the Lower Crystalline Nappe overthe metasedimentary Saxothuringian parautochthon, asinferred from numerous kinematic indicators. KlaÂpova etal. 1998) suggested that the D2 structures in the eclogitesdeveloped in the same kinematic regime.During the ®nal stages <strong>of</strong> D2, the principal compressionthat was acting E±W caused large-scale buckling <strong>of</strong> thewhole nappe sequence <strong>and</strong> <strong>of</strong> the parautochthonous domain.These buckle folds have vertical axial planes <strong>and</strong> foldhinges perpendicular to the L 2 stretching lineation. Weinterpret these structures as a result <strong>of</strong> the buckling <strong>of</strong> thestacked nappe sequence at the end <strong>of</strong> a D2 thrusting episodecaused by the buttressing effect <strong>of</strong> a rigid autochthon. TheMeÏdeÏnec synform<strong>and</strong> the Oberwiesenthal structureoriginated in the same manner.8.2. F3 refolding <strong>and</strong> regional interference patternTwo large-scale F 3 antiforms later refolded the wholenappe system, together with the parautochthonoussequence. These antiforms have E±W striking steep axialplanes <strong>and</strong> westward plunging hinge zones. In the KlõÂnovecantiform, the D3 deformation is associated with the rotation<strong>of</strong> the omphacite L 1 lineation in the eclogites as a result <strong>of</strong>the active rotation <strong>of</strong> eclogitic boudins within weakerorthogneisses <strong>and</strong> metasediments.The increase in the intensity <strong>of</strong> the D3 folding in the south<strong>of</strong> the study area may be the result <strong>of</strong> the original geometry<strong>of</strong> the D2 synforms. If the late F 2 folds have a large wavelength<strong>and</strong> small amplitude as, for example, the MeÏdeÏnecsynform), later compression parallel to their axes will resultin the development <strong>of</strong> dome <strong>and</strong> basin structures interferencepattern type I <strong>of</strong> Ramsay 1962b)). The observed<strong>structural</strong> pattern indicates, however, that late F 2 folds areoverprinted by F 3 folds with moderate to steeply plunginghinges. The development <strong>of</strong> this kind <strong>of</strong> F 3 fold suggests thatlate F 2 folds were non-cylindrical domes <strong>and</strong> basinselongated in a N±S direction with increasing amplitude tothe south. Such periclinal basins <strong>and</strong> domes are commonlydescribed fromthe Zagros Mountains Iran) or the JuraMountains Price <strong>and</strong> Cosgrove, 1990, pp. 262±263). TheN±S compression <strong>of</strong> such a structure may produce theobserved large folds with steeply plunging hinges <strong>and</strong>E±W vertical axial plane. In this manner, an interferencepattern <strong>of</strong> type II originates Ramsay, 1962b). The D3 foldingintensity is thus controlled by D2 fold shape with basementdepth increasing towards the south <strong>and</strong> interlimb angleincreasing towards the north.8.3. Strain pattern in the F 3 fold limbs <strong>and</strong> F 3 fold mechanicsThe strain ellipsoid in the limb zones <strong>of</strong> both the KlõÂnovec<strong>and</strong> MeÏdeÏnec antiforms is characterised by oblate shapessuggesting ¯attening perpendicular to the F 3 fold axialplane Fig. 8). Moreover, the X-axis <strong>of</strong> ®nite strain inthese zones is sub-horizontal <strong>and</strong> E±W in direction as aresult <strong>of</strong> the D3 deformation. We suppose that in E±W63


1388J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392Fig. 10. Calculated evolution <strong>of</strong> the D3 fabric as a result <strong>of</strong> homogeneous isovolumic pure-shear deformation applied to the D2 strain ellipsoid with N±Strending Z-axis <strong>and</strong> vertical X-axis. Black <strong>and</strong> grey lines show the evolution <strong>of</strong> the strain ellipsoid with 5% step <strong>of</strong> strain black dots) being labeled by percent<strong>of</strong> strain accommodated by the X-axis <strong>of</strong> the D2 strain ellipsoid negative values represent shortening). a) Results <strong>of</strong> calculation starting fromthe D2 ellipsoidwith the plane strain symmetry <strong>and</strong> low strain intensity grey ®eld labeled D2). Provided that almost the entire D3 deformation is accommodated by the E±Wtrending Y-axis <strong>of</strong> the D2 ellipsoid, after 40±50% <strong>of</strong> shortening by pure-shear, the resulting fabric will show E±W trending long axis, oblate symmetry, <strong>and</strong>still relatively low intensity <strong>of</strong> deformation grey ®eld labeled D3). b) Results <strong>of</strong> calculation starting from the D2 ellipsoid with prolate symmetry <strong>and</strong> mediumintensity <strong>of</strong> deformation grey ®eld labelled D2). Provided that almost the entire D3 deformation is accommodated by the E±W trending Y-axis <strong>of</strong> the D2ellipsoid, after 40±60% <strong>of</strong> shortening by pure-shear, the resulting fabric will show weak vertical elongation, oblate symmetry <strong>and</strong> high strain intensity grey®eld labelled D3).oriented limbs, the ductile shortening was superimposed onthe original D2 plane strain fabric.We have simulated superposition <strong>of</strong> homogeneouscoaxial deformation on a plane-strain ellipsoid with steepX-axis <strong>and</strong> vertical E±W trending XY plane D2 fabric afterF 3 passive foldingÐFig. 7) <strong>and</strong> the results are shown inFig. 10. Fromthe above presented geological arguments,it is assumed that the D2 <strong>and</strong> D3 Z-axes were parallelafter rotation <strong>of</strong> S 2 foliation during large scale F 3 folding.Thus, considered isovolumic changes <strong>of</strong> the shape <strong>of</strong>fabricellipsoidareinducedbyshortening<strong>of</strong>theformerZ-axis. This isovolumetric strain is accommodated bydifferential elongation <strong>of</strong> former Y- <strong>and</strong> X-axes <strong>of</strong> theD2 ellipsoid. Paths in Fig. 10 represent differentamounts <strong>of</strong> elongation <strong>of</strong> the former X-axis negativevalues represent shortening). Points on the strain paths64


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 1389aNS 2L 2L 2bD 3 kink-b<strong>and</strong>sNL 1L 2L 3D 4 kink-b<strong>and</strong>sFig. 11. Schematic sketches showing the evolution <strong>of</strong> large-scale structures in time. a) Situation after the late D2 deformation is interpreted as a result <strong>of</strong>buttressing <strong>of</strong> the allochthonous unit from the west. Continuous E±W compression leads to the development <strong>of</strong> large-scale periclinal structures the MeÏdeÏnecsynform <strong>and</strong> Oberwiesenthal structure). b) Subsequent N±S compression leads to refolding <strong>of</strong> the late D2 structures by km-scale antiforms the MeÏdeÏnec <strong>and</strong>KlõÂnovec antiforms). Brittle-ductile structures in the ¯at lying S 2 foliation appear during the late D3 deformation. Disappearance <strong>of</strong> the N±S oriented D3compressive stress leads to strengthening <strong>of</strong> the role <strong>of</strong> overburden. Vertical D4 compression produces D4 kink-b<strong>and</strong> folds to kink-b<strong>and</strong>s.show 5% increments <strong>of</strong> strain imposed on a former D2fabric ellipsoid.In those areas that were not affected by the D3 deformation,we have identi®ed three types <strong>of</strong> the D2 strains symmetriesFig. 8) <strong>and</strong> these approximately represent startingpoints <strong>of</strong> trajectories in the Fig. 10. An attempt was made to®nd those ®nite strains observed in natural samples from themegafold limbs. Moreover, different ®nite D3 fabrics mustbe produced after an equal number <strong>of</strong> increments <strong>of</strong> the D3deformation. Our analysis demonstrates Fig. 10) that thiscase is reached when most <strong>of</strong> the imposed shortening isaccommodated by elongation <strong>of</strong> former Y-axis <strong>and</strong> smallchange <strong>of</strong> former X-axis. For plane-strain to constrictionalvertical D2 fabric with higher strain intensity Fig. 10a),65


1390J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392>>1σ 1 (N-S)σ 1 (vertical)temperature (oC)400 500 6002a1132b011depth (km)30 50Fig. 12. Diagram<strong>of</strong> the orientation <strong>of</strong> s 1 versus depth <strong>and</strong> temperature during D3 <strong>and</strong> D4 shows two possible scenarios <strong>of</strong> such an evolution. Path 1±2a±3shows N±S compression during exhumation <strong>of</strong> the studied area, which is demonstrated by progressive development <strong>of</strong> brittle-ductile D3 structures. After thedecreasing role <strong>of</strong> the N±S compression at shallow crustal levels <strong>and</strong> low temperature, the prevailing vertical compression will produce the D4 kink-b<strong>and</strong>s.This scenario is consistent with ®eld observations. The path 1±2b shows the diminishing role <strong>of</strong> the N±S compressive stress <strong>and</strong> increase in the importance <strong>of</strong>vertical compression caused by the overburden at high temperatures in depth. This evolution would result in refolding <strong>of</strong> the F 3 folds by ductile F 4 folds withsubhorizontal axial planes. The PT path adopted fromKonopaÂsek 2001).superimposed D3 deformation will produce strong oblatefabrics without macroscopically visible aggregate lineation.In the case <strong>of</strong> low D2 strain intensities Fig. 10b), the resultingfabric is marked by L±S symmetry with intermediateintensities <strong>and</strong> visible horizontal L 3 aggregate stretchinglineation former Y-axis).The ®eld observations suggest that the F 3 folds developedby mechanisms <strong>of</strong> buckling <strong>of</strong> a single layer <strong>of</strong> orthogneisssurrounded by micaschist. Strong ¯attening in the limbareas <strong>of</strong> the KlõÂnovec antiformalso indicates conditions <strong>of</strong>moderate viscosity contrast between strong orthogneisses<strong>and</strong> weak micaschists.8.4. The origin <strong>of</strong> the D3±D4 kink b<strong>and</strong>s <strong>and</strong> other D4brittle-ductile structuresA question arises: were the above-described D4 kinkb<strong>and</strong>sdeveloped as post-D3 structures, or were they createdduring the early D3 deformation <strong>and</strong> then re-oriented duringthe D3 folding? As described above, the steep metamorphicfabric in metasediments between the northern limb <strong>of</strong> theKlõÂnovec antiform<strong>and</strong> the MeÏdeÏnec antiform, as well as thezone <strong>of</strong> metasediments in the southern limb <strong>of</strong> the KlõÂnovecantiform, are believed to represent the S 3 cleavage. Therefore,the kink-b<strong>and</strong>s affecting the S 3 cleavage must be post-D3. This is the case <strong>of</strong> those kink-folds developed in thesteep southern limb <strong>of</strong> the MeÏdeÏnec antiform<strong>and</strong> in thesteep zones south <strong>of</strong> the KlõÂnovec antiform. On the otherh<strong>and</strong>, the kink-b<strong>and</strong>s developed in the hinge zone <strong>of</strong> theMeÏdeÏnec antiformactually correspond to the axial cleavage<strong>of</strong> this structure <strong>and</strong> are, therefore, developed in the earlystages <strong>of</strong> D3 folding.If we accept a hypothesis that the kink-b<strong>and</strong>s in metasedimentsclose to the KlõÂnovec <strong>and</strong> the MeÏdeÏnec antiformswere not generated in the same stress ®eld as the D3 largescale folds, another local source <strong>of</strong> stress must have existedto generate these structures. A possible source <strong>of</strong> stressduring the D4 is the weight <strong>of</strong> the overburden after therelaxation <strong>of</strong> the D3 stresses. This will produce subverticals 1 , which has been documented in the southern part <strong>of</strong> theKlõÂnovec antiform. Moreover, the expected deformationassociated with the relaxation <strong>of</strong> the D3 stresses is ratherlow. This is consistent with the development <strong>of</strong> kink b<strong>and</strong>s<strong>and</strong> brittle-ductile clevage in orthogneisses <strong>and</strong> metasediments.The amount <strong>of</strong> strain achieved during theirdevelopment is probably more than two orders <strong>of</strong> magnitudeless than the deformation associated with the D1±D3 deformation<strong>and</strong>, thus, rather insigni®cant at the scale <strong>of</strong> thestudied area.8.5. Variations <strong>of</strong> principal compression direction throughtimeThe analysis <strong>of</strong> the D2, D3 <strong>and</strong> D4 structures shows66


J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392 1391complex stress±temperature±time evolution Fig. 11). Asdiscussed above, the D2 structures are associated with westwardthrusting <strong>of</strong> the Lower Crystalline nappe over theSaxothuringian basement suggesting non-coaxial deformationwith E±W oriented s 1 direction.Strain analysis within D3 fold limbs indicates a N±Soriented horizontal compression s 1 <strong>and</strong> important verticalstress s 2 inhibiting elongation in the vertical direction.Then, the only elongation accommodating N±S horizontalcompression s 1 is possible in the horizontal E±W direction.This indicates an important role <strong>of</strong> rigid overburden forstresses acting in a low viscosity layer at depth. SubhorizontalN±S compression leads to lateral ¯ow <strong>of</strong> materialwithout the vertical component in the early stages <strong>of</strong>D3 deformation. The same stress regime, however, isresponsible for the formation <strong>of</strong> kink b<strong>and</strong> structures inthe hinge <strong>of</strong> the MeÏdeÏnec antiform. Kink b<strong>and</strong> structuresrepresent a brittle-ductile regime Dewey, 1965, 1969)<strong>and</strong>, thus, indicate a decrease in temperature under thesame orientation <strong>of</strong> the D3 stress ®eld. It is, therefore,suggested that the horizontal N±S compression operatedduring an uplift <strong>of</strong> the whole region fromthe deep to thesupracrustal level.The presence <strong>of</strong> D4 kink b<strong>and</strong>s must be associated withthe disappearance <strong>of</strong> horizontal stresses <strong>and</strong> increase insubvertical s 1 compression. If this change <strong>of</strong> stress regimewould occur at high temperature conditions, then the F 3folds would be refolded by large-scale ductile F 4 foldswith subhorizontal axial planes. Fig. 12 documents thatthe F 3 folding started at peak temperature conditions ca.6008C) enabling viscous buckling <strong>of</strong> the nappe sequencecaused by shortening in the N±S direction. This deformationphase continued during a temperature decreaseassociated with uplift, resulting in the development <strong>of</strong> D 3kink-b<strong>and</strong>s <strong>and</strong> a crenulation cleavage in the hinge zones <strong>of</strong>large-scale anticlines. The latest F 4 phase occurred underrelatively low temperature conditions <strong>and</strong> indicates thedisappearance <strong>of</strong> horizontal stress, allowing vertical shortening<strong>of</strong> steep D 3 fabric. These changes in stress can beinterpreted as an increase <strong>of</strong> the role <strong>of</strong> overburden aftertermination <strong>of</strong> the D3 compressive stress.8.6. Exhumation <strong>of</strong> eclogites <strong>and</strong> the importance <strong>of</strong>extension in the Czech part <strong>of</strong> the KrusÏne hory Erzgebirge)MountainsIn contrast with the western Saxothuringian domain,where the boundary between eclogites-bearing nappes <strong>and</strong>supracrustal autochthon can be easily identi®ed, a similarlimit is dif®cult to establish in the eastern KrusÏne horyMountains. Using <strong>structural</strong> <strong>and</strong> petrological criteria, wehave de®ned the boundary between the parautochthonousSaxothuringian metasediments <strong>and</strong> allochthonous eclogitesbearingnappe. Thrusting-associated structures developed inboth the parautocthonous <strong>and</strong> allochthonous units documentthat the nappe emplacement occurred in the middle crust at adepth corresponding to 13±15 kbar. Field observations,however, are not able to provide any information aboutthe mechanism <strong>of</strong> emplacement <strong>of</strong> eclogites from a depthcorresponding to 26 kbar to the base <strong>of</strong> non-eclogiticorthogneiss nappe. This work shows mechanical behaviour<strong>of</strong> the crust during <strong>and</strong> after the nappe emplacement witheclogites as a part <strong>of</strong> lithological assemblage. The exhumation<strong>of</strong> assembled parautochthonous <strong>and</strong> allochthonous unitsto supracrustal levels is associated with complex <strong>structural</strong>reworking <strong>of</strong> originally simple fabric during the subsequentN±S shortening, which is responsible for the ®nal pattern <strong>of</strong>the central part <strong>of</strong> the KrusÏne hory Mountains in the CzechRepublic.There is a range <strong>of</strong> publications emphasising the role <strong>of</strong>the late Variscan extensional deformation for the ®naltectonic <strong>and</strong> metamorphic pattern <strong>of</strong> the German part <strong>of</strong>the eastern Saxothuringian domain Willner et al., 1994;Krohe, 1996, 1998; RoÈtzler et al., 1998). These interpretationsare based on regional distribution <strong>of</strong> metamorphicunits <strong>and</strong> orientation <strong>of</strong> shear structures in different areas<strong>of</strong> the Erzgebirge. In this study, we have examined a particularlysuitable area with steeply developed anisotropyaffected by vertical shortening <strong>and</strong> demonstrated that itachieves no more than ®rst percents <strong>of</strong> the bulk strain duringthis event. If the extensional tectonics would be a keyregime responsible for ®nal geometry <strong>of</strong> studied crystallinecomplexes then signi®cantly more important vertical shorteningshould be expected ®rst in areas with subverticallydeveloped anisotropy. Therefore, we suggest that the <strong>structural</strong>pattern <strong>of</strong> the whole Saxothuringian domain should bere-evaluated in terms <strong>of</strong> detailed <strong>structural</strong> analysis todemonstrate the real signi®cance <strong>of</strong> late orogenic extensionin this part <strong>of</strong> the Bohemian Massif.AcknowledgementsWe are very grateful to John Cosgrove for helpfulcomments on an early version <strong>of</strong> the manuscript. Thepaper has also bene®ted from the comments <strong>of</strong> G. Oliver<strong>and</strong> W. Franke. We also thank Jaroslav Synek for drawingFig. 6. This work was funded by the Grant Agency <strong>of</strong> theCzech Republic, grant no. 205/96/0279.ReferencesCobbold, P.R., Cosgrove, J.W., Summers, J.M., 1971. Development <strong>of</strong>internal structures in deformed anisotropic rocks. Tectonophysics 121), 23±53.Dewey, J.F., 1965. Nature <strong>and</strong> origin <strong>of</strong> kink-b<strong>and</strong>s. Tectonophysics 1,459±494.Dewey, J.F., 1969. The origin <strong>and</strong> development <strong>of</strong> kink b<strong>and</strong>s in a foliatedbody. Geological Journal 6, 193±216.Flinn, D., 1962. On folding during three dimensional progressive deformation.Quarterly Journal <strong>of</strong> the Geological Society <strong>of</strong> London 118, 385±428.Franke, W., 1993. The Saxonian Granulites: a metamorphic core complex?Geologische Rundschau 82, 505±515.67


1392J. KonopaÂsek et al. / Journal <strong>of</strong> Structural Geology 23 2001) 1373±1392Franke, W., Dallmeyer, D., Weber, K., 1995. Geodynamic evolution <strong>of</strong> theVariscan Belt). In: Dallmeyer, D., Franke, W., Weber, K. Eds.). Pre-Permian Geology <strong>of</strong> Central <strong>and</strong> Eastern Europe. Springer-Verlag, NewYork, pp. 579±593.H<strong>of</strong>mann, J., MatheÂ, G., Werner, C.D., 1988. Saxothuringian zone <strong>and</strong> thecentral German crystalline zone in the German Democratic Republik.In: Zoubek, V., CogneÂ, J., Kozhoukharov, D., KraÈutner, H.G. Eds.).Precambrian in Younger Fold Belts. Wiley <strong>and</strong> Sons, Chicester, p. 885.Holub, F.V., SoucÏek, J., 1994. Blueschist-greenschist metamorphism <strong>of</strong>metabasites in the western KrusÏne hory Erzgebirge) Mts. ZentralblattfuÈr Geologie und PalaÈontologie. Teil I 1992 7/8), 815±826.Hoth, K., Lorenz, W., Hirschmann, G., Berger, H.-J., 1979. LithostratigraphischegliederungsmoÈglichkeiten des regionalmetamorphen Jungproterosoikumsam Beispiel des Erzgebirges. Zeitschrift fuÈrGeologische Wissenschaften 7, 397±404.Hoth, K., Wasternack, J., Berger, H.-J., Breiter, K., MlcÏoch, B., SchovaÂnek,P. 1994. Geologische Karte Erzgebirge/Vogtl<strong>and</strong>, 1:100,000.SaÈchsisches L<strong>and</strong>esamt fuÈr Umwelt und Geologie.KlaÂpovaÂ, H., KonopaÂsek, J., Schulmann, K., 1998. Eclogites from theCzech part <strong>of</strong> the Erzgebirge: multi-stage metamorphic <strong>and</strong> <strong>structural</strong>evolution. Journal <strong>of</strong> the Geological Society <strong>of</strong> London 155, 567±583.KonopaÂsek, J., 1998. Formation <strong>and</strong> destabilization <strong>of</strong> the high pressureassemblage garnet±phengite±paragonite KrusÏne hory Mountains,Bohemian Massif): the signi®cance <strong>of</strong> the Tschermak substitutioninthe metamorphism <strong>of</strong> pelitic rocks. Lithos 42, 269±284.KonopaÂsek, J., 2001. Eclogitic micaschists in the central part <strong>of</strong> the KrusÏneÂhory Mountains Bohemian Massif). European Journal <strong>of</strong> Mineralogy13, 1.Kossmat, F., 1925. UÈ bersicht der Geologie von Sachsen. G.A. Kaufmann,Leipzig.Kossmat, F., 1927. Gliederung des varistischen Gebirgsbaues. Abh<strong>and</strong>lungendes Sachsischen Geologischen L<strong>and</strong>esamtes 1, 39.KotkovaÂ, J., KroÈner, A., Todt, W., Fiala, J., 1996. Zircon dating <strong>of</strong> NorthBohemian granulites, Czech Republic: further evidence for the LowerCarboniferous high-pressure event in the Bohemian Massif.Geologische Rundschau 85, 154±161.Krohe, A., 1996. Variscan tectonics <strong>of</strong> Central Europe: postaccretionaryintraplate deformation <strong>of</strong> weak continental lithosphere. Tectonics 15,1364±1388.Krohe, A., 1998. Extending a thickened crustal bulge: toward a new geodynamicevolution model <strong>of</strong> the paleozoic NW Bohemian Massif,German Continental Deep Drilling site SE Germany). Earth-ScienceReviews 44, 95±145.KroÈner, A., Willner, A.P., Hegner, E., Frischbutter, A., H<strong>of</strong>mann, J.,Bergner, R., 1995. Latest Precambrian zircon ages, Nd isotopicsystematics <strong>and</strong> PT evolution <strong>of</strong> granitoid orthogneisses <strong>of</strong> theErzgebirge, Saxony <strong>and</strong> Czech Republic. Geologische Rundschau 84,437±456.KroÈner, A., Willner, A.P., 1998. Time <strong>of</strong> formation <strong>and</strong> peak <strong>of</strong> VariscanHP±HT metamorphism <strong>of</strong> quartz-feldspar rocks in the centralErzgebirge, Saxony, Germany. Contributions to Mineralogy <strong>and</strong>Petrology 132, 1±20.Lisle, R.J., 1977. Estimation <strong>of</strong> tectonic strain ratio from the mean shape <strong>of</strong>deformed elliptical markers. Geologie en Mijnbow 56, 140±144.Matte, P., Maluski, H., Rajlich, P., Franke, W., 1990. Terrane boundaries inthe Bohemian Massif: result <strong>of</strong> large-scale Variscan shearing. Tectonophysics177, 151±170.Mehnert, K.R., 1939. Die Meta-Konglomerate des WiesenthalerGneiszuges imsaÈchsischen Erzgebirge. Mineralogische und PetrographischeMitteilungen 50, 194±272.MlcÏoch, B., Schulmann, K., 1992. Superposition <strong>of</strong> Variscan ductile sheardeformation on pre-Variscan mantled gneiss structure Catherine dome,Erzgebirge, Bohemian Massif). Geologische Rundschau 81 2), 501±513.Pietzsch, K. 1914. UÈ ber das Alter der dichten Gneise des saÈchsischenErzgebirges. Zentralblatt fuÈr Mineralogie 1914, 202±211, 225±241.Price, N.J., Cosgrove, J.W., 1990. Analysis <strong>of</strong> Geological Structures..Cambridge University Press, Cambridge Reprinted 1994).Rajlich, P., 1987. Hercynian ductile tectonics in the Bohemian Massif.Geologische Rundschau 76 3), 755±786.Ramsay, J.G., 1962a. The geometry <strong>of</strong> conjugate fold systems. GeologicalMagazine 99, 516±526.Ramsay, J.G., 1962b. Interference patterns produced by the superposition<strong>of</strong> folds <strong>of</strong> similar types. Journal <strong>of</strong> Geology 70, 466±481.Ramsay, J.G., Huber, M.I., 1983. The Techniques <strong>of</strong> Modern StructuralGeology. Volume 1, Strain Analysis. Academic Press, London.Ramsay, J.G., Huber, M.I., 1987. The Techniques <strong>of</strong> Modern StructuralGeology. Volume 2, Folds <strong>and</strong> Fractures. Academic Press, LondonFourth printing 1993).RoÈtzler, K., Schumacher, R., Maresch, W., Willner, A., 1998. Characterization<strong>and</strong> geodynamic implications <strong>of</strong> contrasting metamorphicevolution in juxtaposed high-pressure units <strong>of</strong> the Western ErzgebirgeSaxony, Germany). European Journal <strong>of</strong> Mineralogy 10, 261±280.Sattran, V., 1963. Metaconglomerates from the KrusÏe hory Mountains <strong>and</strong>their genetic signi®cance. SbornõÂk GeologickuÂch VeÏd, Geologie 2, 41±62 in Czech).Sattran, V. 1967. Geological map <strong>of</strong> the central part <strong>of</strong> the KrusÏne horyMountains between KlaÂsÏterec nad OhrÏõ <strong>and</strong> Vejprty 1:25,000). CzechGeological Survey. Unpublished.Scheumann, K.H., 1935. Die Rotgneise der Glimmerschieferdecke dessaÈchsischen Granulitgebirges. Berichten mathematische-physikalischeKlasse der saÈchsischen Akademie der Wissenschaften 87, 251±286.SchmaÈdicke, E., Okrusch, M., Schmidt, W., 1992. Eclogite-facies rocks inthe Saxonian Erzgebirge, Germany: high pressure metamorphism undercontrasting P±T conditions. Contributions to Mineralogy <strong>and</strong> Petrology110, 226±241.SchmaÈdicke, E., Mezger, K., Cosca, M.A., Okrusch, M., 1995. VariscanSm±Nd <strong>and</strong> Ar±Ar ages <strong>of</strong> eclogite facies rocks from the Erzgebirge,Bohemian Massif. Journal <strong>of</strong> Metamorphic Geology 13, 537±552.SÏkvor, V. 1975. Geology <strong>of</strong> the Czech part <strong>of</strong> the KrusÏne hory Mts. <strong>and</strong>SmrcÏiny. Knihovna U U G 48, 120pp. [in Czech].Werner, O., Lippolt, H.J., Hess, J.C., 1997. Isotopische R<strong>and</strong>bedingungenzur Bestimmung der Alter tektonischer Prozesse im variscischenErzgebirge. Terra Nostra 97/5, 199±204.Willner, A.P., RoÈtzler, K., Krohe, A., Maresch, W., Schumacher, R., 1994.Druck-Temperatur-Deformations-Entwicklung verschiedener KrustengesteineimErzgebirge: Eine Modellregion fuÈr die Exhumierung vonKrustengesteinen. Terra Nostra 3/94, 104±106.Willner, A.P., RoÈtzler, K., Maresch, W.V., 1997. Pressure±temperature <strong>and</strong>¯uid evolution <strong>of</strong> quartzo±feldspathic metamorphic rocks with a relichigh-pressure, granulit-facies history fromthe Central ErzgebirgeSaxony, Germany). Journal <strong>of</strong> Petrology 38, 307±336.68


JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. B1, 2023, doi:10.1029/2001JB000632, 2003Strain distribution <strong>and</strong> fabric development modeledin active <strong>and</strong> ancient transpressive zonesKarel Schulmann, 1 Alan Bruce Thompson, 2 Ondrej Lexa, 1 <strong>and</strong> Josef Ježek 3Received 23 May 2001; revised 26 April 2002; accepted 20 May 2002; published 16 January 2003.[1] A model based on kinematics <strong>of</strong> transpression allows the measurable internalparameters (simultaneous strain, fabric, <strong>and</strong> vertical elevation) to be simulated in relationto the macroscopically determined external parameters <strong>of</strong> transpressive zones (i.e., zonewidth, velocity, angle <strong>of</strong> convergence, <strong>and</strong> zone depth). We present a diagram <strong>of</strong>convergence angle against time for homogeneous transpression, where isolines <strong>of</strong> strainintensity D, strain symmetry K, <strong>and</strong> vertical elevation <strong>of</strong> rock samples are superposed.However, the distribution <strong>of</strong> internal strain parameters is sensitive to three types <strong>of</strong> strainpartitioning: (1) Discrete partitioning results in general decrease <strong>of</strong> finite strainaccumulations <strong>and</strong> in increase <strong>of</strong> pure shear component, (2) ductile partitioning splits thetranspressional domain into a pure shear zone where strain accumulations decreases <strong>and</strong> ina wrench-dominated zone where strain accumulations increases, <strong>and</strong> (3) viscositypartitioning is marked by different strain rates in zones <strong>of</strong> different viscosity <strong>and</strong> therefore bydifferent strain parameters. INDEX TERMS: 8025 Structural Geology: Mesoscopic fabrics; 8110Tectonophysics: Continental tectonics—general (0905); 0905 Exploration Geophysics: Continental structures(8109, 8110); KEYWORDS: transpression, strain partitioning, exhumation, oblique convergenceCitation: Schulmann, K., A. B. Thompson, O. Lexa, <strong>and</strong> J. Ježek, Strain distribution <strong>and</strong> fabric development modeled in active <strong>and</strong>ancient transpressive zones, J. Geophys. Res., 108(B1), 2023, doi:10.1029/2001JB000632, 2003.1. Introduction[2] Modern tectonic studies face the problem <strong>of</strong> underst<strong>and</strong>ingthe relationships between small-scale structures<strong>and</strong> large-scale geometry in orogenic zones. The linksbetween the external <strong>and</strong> internal parameters governingthe problem are particularly important. These factors maybe viewed as far-field causes related to local effects. Weexamine here these relationships in ancient <strong>and</strong> activetranspressive zones. External <strong>structural</strong> parameters are representedby the geometry <strong>of</strong> orogenic zones (i.e., zone width(distance between colliding plates), obliquity (angle <strong>of</strong>convergence) <strong>and</strong> depth <strong>of</strong> the transpressive zone. Irregularities<strong>of</strong> plate boundaries (shape <strong>of</strong> indenting block), <strong>and</strong>the velocity <strong>and</strong> duration <strong>of</strong> plate convergence) may alsoplay an important role. Internal (local) <strong>structural</strong> parametersthat can be examined are fabric <strong>and</strong> strain intensity, symmetry<strong>of</strong> fabrics, orientations <strong>of</strong> strain axes, pressure (depth)memory, <strong>and</strong> metamorphic facies <strong>of</strong> the rocks. Theseinternal <strong>structural</strong> parameters are determined by the interactions<strong>of</strong> the external forces with local lithological heterogeneities<strong>and</strong> rheologies in the transpressive zone.[3] Relative motion <strong>of</strong> lithospheric plates on a sphericalsurface is such that the plate convergence vectors are <strong>of</strong>ten1 Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Faculty <strong>of</strong> Science,Charles University, Prague, Czech Republic.2 Department Erdwissenschaften, ETH Zurich, Switzerl<strong>and</strong>.3 Institute <strong>of</strong> Applied Mathematics <strong>and</strong> Computer Science, Faculty <strong>of</strong>Science, Charles University, Prague, Czech Republic.Copyright 2003 by the American Geophysical Union.0148-0227/03/2001JB000632$09.00not orthogonal to plate boundaries [McKenzie <strong>and</strong> Parker,1967; Dewey, 1975]. These plate boundaries experiencecombined transcurrent <strong>and</strong> convergent displacements associatedwith development <strong>of</strong> deformation zones <strong>of</strong> differentsize. Within continental blocks, the deformation is not onlyrestricted to active plate boundaries but occurs within zones<strong>of</strong> weakness inside rigid continental domains [e.g., Tommasi<strong>and</strong> Vauchez, 1997] <strong>and</strong> can be approximately described asa deformation <strong>of</strong> a weak zone bounded by rigid blocks withsteep parallel walls. All the mentioned types <strong>of</strong> deformationzones can be more or less described by a model called‘‘transpression,’’ introduced first by Harl<strong>and</strong> [1971], developedby S<strong>and</strong>erson <strong>and</strong> Marchini [1984], <strong>and</strong> elaborated bymany others. Despite its simplicity, it seems that the modelstill has much to contribute toward our underst<strong>and</strong>ing <strong>of</strong> thenature <strong>of</strong> convergent orogeny. In this work we develop themodel to quantify the effects <strong>of</strong> external (macroscopic)parameters on temporal development <strong>of</strong> internal strainparameters in transpressive (obliquely convergent) weakzones <strong>of</strong> finite width.2. Structural Definition <strong>of</strong> Transpression <strong>and</strong>Problems to Be Solved[4] The classical zone <strong>of</strong> transpressive deformation is atabular weak region subjected between its steep walls to asimultaneous pure shear <strong>and</strong> simple shear. In the model <strong>of</strong>S<strong>and</strong>erson <strong>and</strong> Marchini [1984] the material is able to slipfreely upward (vertically extruded) along the walls <strong>of</strong> thetranspression zone. The transpressive deformation zonedefined by S<strong>and</strong>erson <strong>and</strong> Marchini [1984] was also limiteddownward by a rigid horizontal plate (like a rigid floor),ETG 6 - 169


ETG 6 - 2 SCHULMANN ET AL.: STRAIN DISTRIBUTIONthus allowing extrusion only in the vertical direction.Fossen <strong>and</strong> Tik<strong>of</strong>f [1998] suggested boundary conditionswhere zones <strong>of</strong> transpressive deformation are able to growthor shrink vertically as well as horizontally along strike.[5] Fossen <strong>and</strong> Tik<strong>of</strong>f [1993] defined two types <strong>of</strong> transpressionon the basis <strong>of</strong> the orientation <strong>of</strong> the instantaneousstretching axes: (1) pure-shear-dominated transpression inwhich v 1 <strong>and</strong> v 3 (eigenvectors <strong>of</strong> instantaneous straincorresponding to maximum <strong>and</strong> minimum elongation)define a vertical plane <strong>and</strong> the v 1 (lineation) is vertical,<strong>and</strong> (2) wrench-dominated transpression in which v 1 <strong>and</strong> v 3are horizontal, while the v 2 (intermediate eigenvector) isvertical. Pure-shear-dominated transpression operates whenthe angle between relative plate convergence direction <strong>and</strong>plate boundary is greater than 20° [Pinet <strong>and</strong> Cobbold,1992], <strong>and</strong> the main axis <strong>of</strong> finite strain is always vertical. Inwrench-dominated transpression acting at an angle <strong>of</strong> convergence(a) <strong>of</strong>


SCHULMANN ET AL.: STRAIN DISTRIBUTION ETG 6 - 3Figure 1. Block diagrams <strong>and</strong> principal strain axes orientations for transpressive systems with externalparameters: (a) plate velocity (v = 1 to 10 cm yr 1 ) <strong>and</strong> initial width (d = 50 to 300 km), which give theratio R vd from 2 to 0.03, respectively, <strong>and</strong> rigid floor depth (RFD = 70, 40 km). As internal parameters,the instantaneous orientation <strong>of</strong> foliation plane (XY ) <strong>and</strong> lineation (X ) are related to angle <strong>of</strong> convergence(a) <strong>of</strong>(c)a =60° <strong>and</strong> (d) a


ETG 6 - 4 SCHULMANN ET AL.: STRAIN DISTRIBUTIONFigure 2. (a) Diagram showing relation between strain intensity (D) <strong>and</strong> time (expressed by timeparameter, k t , where t = k t /R vd ). The dot-dashed curves show time evolution <strong>of</strong> D for different angle <strong>of</strong>convergence (a). The distribution curve at the left shows the envelope <strong>of</strong> natural strains summarized byPfiffner <strong>and</strong> Ramsay [1982], Hrouda [1993], <strong>and</strong> others. The horizontal line through the maximum (at D =1.2) shows the strain observable by field measurements. (b) The relationship between strain intensity (D)<strong>and</strong> angle <strong>of</strong> convergence (a) through three time sections (5, 10, <strong>and</strong> 15 Myr). The contours show variations<strong>of</strong> R vd values, which distinguish narrow <strong>and</strong> fast converging transpressive orogens (R vd > 0.5) from wide<strong>and</strong> slowly converging ones (R vd < 0.5). Vertical lines at convergence angles 30° <strong>and</strong> 50° show strainaccumulations for R vd values <strong>of</strong> active transpressive zones in New Zeal<strong>and</strong> <strong>and</strong> Sumatra, respectively.suggests that the base <strong>of</strong> lithospheric transpressional zonesis not always present as a rigid floor. In such a case, verticalmovements within the system are controlled by isostaticalresponse <strong>and</strong> can thus lead also to downward motion <strong>of</strong>material early in the shortening history.4. Modeling <strong>of</strong> Internal Parameters <strong>of</strong>Transpressional Systems[16] We calculate the internal strain parameters <strong>of</strong> transpressionalsystems (strain rate, finite strain intensity, <strong>and</strong>symmetry <strong>and</strong> orientation <strong>of</strong> finite strain axes) in terms <strong>of</strong>external parameters defined above. The description <strong>of</strong>method <strong>of</strong> calculations <strong>and</strong> derivation <strong>of</strong> equations are givenin Appendices A–D.4.1. Strain Rates <strong>and</strong> Finite Strain Intensities[17] The first result <strong>of</strong> our model is that in the range <strong>of</strong>assumed R vd , a strain rate interval from 10 14 s 1 to 10 16s 1 is obtained. This is within the range <strong>of</strong> generallyassumed strain rates extrapolated from experimental laboratorydata [Carter <strong>and</strong> Tsenn, 1987] <strong>and</strong> corresponds wellto that deduced for natural orogens by Pfiffner <strong>and</strong> Ramsay[1982]. Theoretical curves <strong>of</strong> finite strain accumulation fordifferent obliquities are presented in Figure 2b. In Figure 2bthe strain intensity parameter D (see Appendix A) is plottedon the vertical axis against the time parameter, k t , whichrelates the time <strong>of</strong> deformation with R vd , the ratio <strong>of</strong>convergence velocity <strong>and</strong> initial zone widthk t ¼ tR vd :The introduction <strong>of</strong> such a time parameter follows from thedefinition <strong>of</strong> the transpression model <strong>and</strong> allows us to graphthe temporal developments corresponding to zones <strong>of</strong>different width <strong>and</strong> convergence velocities (Figure 2a). Ifthe ratio R vd = 1, the scale <strong>of</strong> the k t axis represents timedirectly in million years. For other R vd values we obtain thecorresponding time from equation (1).[18] In order to compare Figure 2a with those on Figure 7<strong>of</strong> Pfiffner <strong>and</strong> Ramsay [1982] each <strong>of</strong> the curves calculatedfor convergence angles varying from a =0° to 90° is labeledwith its strain rate value. Because <strong>of</strong> the triaxial character <strong>of</strong>the deformation in transpressive zones we use a strain ratecalculated as rate <strong>of</strong> change <strong>of</strong> the square root <strong>of</strong> theminimum eigenvalue corresponding to short axis <strong>of</strong> instantaneousstrain tensor, <strong>and</strong> instead <strong>of</strong> R = X/Z, which is a goodcharacteristic for plane strain we use the D parameter.ð1Þ72


SCHULMANN ET AL.: STRAIN DISTRIBUTION ETG 6 - 5[19] The strain rates are expressed in terms <strong>of</strong> R vd . Thefinite strain development is more rapid for pure frontalconvergence than for a transpressional zone <strong>of</strong> any obliquity.The convergence with a =90° <strong>and</strong> a =0° end-membercurves represent pure shear <strong>and</strong> simple shear strain ratepaths in Figure 7 <strong>of</strong> Pfiffner <strong>and</strong> Ramsay [1982]. Thedensity distribution curve on the left side <strong>of</strong> Figure 2aindicates the distribution <strong>of</strong> finite strains plotted fromPfiffner <strong>and</strong> Ramsay [1982] <strong>and</strong> Hrouda [1993] <strong>and</strong> completedwith finite strain data <strong>of</strong> Rajlich et al. [1988],Schultz-Ela <strong>and</strong> Hudleston [1991], Schulmann et al.[1994], <strong>and</strong> Kirkwood [1995].[20] The strain intensity D is strongly dependent on boththe angle <strong>of</strong> convergence a, <strong>and</strong> on the ratio R vd .Itistherefore highest for frontal collision at high R vd . Withdecreasing R vd <strong>and</strong> decreasing convergence angle, the strainintensity decreases rapidly (Figure 2b). Starting from zonesmarked by the value R vd = 0.1 the variations in strainintensity are negligible with time. However, zones withR vd values higher than 0.2 show rapid increase <strong>of</strong> strainintensity after only a short duration <strong>of</strong> convergence. Thehighest gradient <strong>of</strong> strain intensity increase occurs for lowconvergence angles (wrench-dominated transpression). Forhigher convergence angles, the strain intensity increasessmoothly with increasing convergence angle. The strainintensity value also increases steadily with time, which isconsistent with progressive accumulation <strong>of</strong> finite strainduring shortening <strong>of</strong> a collisional belt.4.2. Temporal Evolution <strong>of</strong> Foliation <strong>and</strong> Lineation[21] In our model, in pure-shear-dominated transpression(a >20°) the mineral growth lineation is vertical for anywidth <strong>of</strong> the belt <strong>and</strong> any angle <strong>of</strong> convergence. In wrenchdominatedtranspression (a < 20°), the orientation <strong>of</strong>lineation varies for different angle <strong>of</strong> convergence a <strong>and</strong>for different belt width d (Figure 3a). The orientation <strong>of</strong> thelongest strain axis, s 1 , starts at a maximum angle <strong>of</strong> 45° (fora simple shear zone) with respect to zone margin <strong>and</strong> tendstoward parallelism with simple shear direction (a = 0)represented by zone boundaries with increasing amount <strong>of</strong>deformation. However, the rate <strong>of</strong> lineation reorientation isvery rapid in the case <strong>of</strong> high R vd . For instance for R vd =2<strong>and</strong> convergence at an angle <strong>of</strong> 10°, the lineation rotatesfrom an initial angle <strong>of</strong> 40°–5° in 3 Myr. For R vd = 0.03 itwill take 200 Myr. An important conclusion is that in widebelts <strong>and</strong> slow highly oblique convergence (e.g., R vd = 0.03)the stretching lineation should be oriented for long times ata high angle (20°–35°) with respect to plate boundariesafter a long period <strong>of</strong> convergence (20–50 Myr, Figure 3a).In contrast, narrow oblique belts <strong>of</strong> rapid convergence (e.g.,R vd = 2) should have stretching lineations which are subparallel(


ETG 6 - 6 SCHULMANN ET AL.: STRAIN DISTRIBUTIONFigure 3. (a) Diagram showing orientation <strong>of</strong> mineral lineation (x axis) through time (parameter k t )forwrench-dominated transpression (a


SCHULMANN ET AL.: STRAIN DISTRIBUTION ETG 6 - 7Figure 4. (a) Diagram showing the development <strong>of</strong> strain symmetry (K ) in time (parameter k t )forvarious convergence angles (a). (b) Diagram showing the relationship between strain symmetry (K ) <strong>and</strong>angle <strong>of</strong> convergence (a) through three time sections (5, 10, <strong>and</strong> 15 Myr). The curves show only slightdependence <strong>of</strong> K on the R vd values <strong>and</strong> on time.[31] However, the RFD is not easily defined in ancienttranspressional zones, <strong>and</strong> we can make only rough estimatesbased on petrological data <strong>and</strong> observed depth <strong>of</strong>Moho. Such analysis shows that in the case <strong>of</strong> shear zones<strong>of</strong> southern Madagascar the depth in which the rocks wereoriginally located was around 60–70 km [Martelat et al.,1997; Pili et al., 1997]. Martelat et al. <strong>and</strong> Pili et al. alsosuggested that the lithospheric shear zones are coupled withunderlying mantle so that the RFD may be located evendeeper in the mantle lithosphere [Teyssier <strong>and</strong> Tik<strong>of</strong>f, 1997;Vauchez et al., 1998].6. Superposition <strong>of</strong> Strain Parameters inTranspressive Belts[32] The preceding considerations <strong>and</strong> results enable us tocreate a type <strong>of</strong> map <strong>of</strong> strain parameters registered in rockselevated to the surface. Figure 6 shows such a map whereisolines <strong>of</strong> strain intensities D <strong>and</strong> isolines <strong>of</strong> strain symmetryK are superposed on a diagram <strong>of</strong> convergence angleagainst time parameter (a - k t space). We have addedcontours <strong>of</strong> RFD/z 0 to this (dashed curves in Figure 6).These shows the times when the samples are elevated closeto the surface. The ratio RFD/z 0 also expresses the acrosswidthshortening <strong>of</strong> the zone. The shaded area in a -] k tspace shows the range <strong>of</strong> naturally observed strain intensities.For these D values (strain intensities) the strainsymmetry parameter K corresponds well with high or lowa (angle <strong>of</strong> convergence), respectively, for vertical <strong>and</strong>horizontal orientation <strong>of</strong> lineation. On the other h<strong>and</strong> theintensity <strong>of</strong> strain (D) is not very sensitive to the angle <strong>of</strong>convergence greater than 20° (pure-shear-dominated transpression)but is strongly dependent on time. For the case <strong>of</strong>very obliquely convergent zones (low a) the time needed toaccumulate observed strain intensities are almost twice aslong as for high convergence angles. The maximum strainsat the right side <strong>of</strong> the shaded area in Figure 6 may beattributed to rock samples elevated from thickened midcrustaldepths (40 km) <strong>of</strong> transpressional zones. However,this is only valid for zones with convergence angles >50°.For more oblique zones, only very shallow samples (uppermost25% <strong>of</strong> the zone) can be exhumed. This also meansthat horizontal stretching lineations may be exhumed fromvery shallow depths in s<strong>of</strong>t transpressional zones.7. Effects <strong>of</strong> Strain Partitioning on TemporalStrain Parameter Development[33] Next we examine the effects <strong>of</strong> three different types<strong>of</strong> strain partitioning on the temporal development <strong>of</strong> finitestrain parameters.[34] Discrete displacement partitioning is modeled usingan approach <strong>of</strong> Teyssier et al. [1995]. In their model,variable amounts <strong>of</strong> total lateral displacement can be consideredto be consumed by discrete faulting. This isexpressed by a ratio p 1 between fault-accommodated lateraldisplacement <strong>and</strong> total lateral displacement. The evolution<strong>of</strong> strain parameters in the viscous domain <strong>of</strong> homogeneous75


ETG 6 - 8 SCHULMANN ET AL.: STRAIN DISTRIBUTIONFigure 5. Vertical elevation rate for transpression expressed in terms <strong>of</strong> angle <strong>of</strong> convergence (a) <strong>and</strong>time parameter (k t ) for different base depths, RFD = 40, 70 km, <strong>and</strong> different original sample depths z 0 =30, 60 km. The curves show elevation achieved by these samples in a given time. For example, in Figure5a (RFD = 40 km, z 0 = 30 km) for convergence angle a =80° sample is elevated to depth 10 km after k t =1.2. For the case <strong>of</strong> R vd = 0.1 the time <strong>of</strong> elevation corresponds to 12 Myr.deformation, where the rest <strong>of</strong> lateral displacement <strong>and</strong> thewhole across strike shortening are accommodated, can bededuced using our hypothetical strain map (Figure 6). Toevaluate the influence <strong>of</strong> partitioning on strain parameters,recalculation <strong>of</strong> values R vd <strong>and</strong> a are made using theequations (B1) <strong>and</strong> (B2) (in Appendix B) or they can betaken from Figure 7. The effect <strong>of</strong> discrete partitioning onfinite strain parameters is expressed by a virtual increase <strong>of</strong>convergence angle <strong>and</strong> decrease <strong>of</strong> R vd (decrease <strong>of</strong> plateconvergence velocity or increase <strong>of</strong> weak zone width). Thusfor a transpression zone with an angle <strong>of</strong> convergence <strong>of</strong>45°, R vd equal 0.1, <strong>and</strong> with 50% <strong>of</strong> lateral displacementconsumed by discrete faults, we use values <strong>of</strong> a 0 = 63.43°0<strong>and</strong> R vd = 0.079 (Figure 7). For 10 Myr <strong>of</strong> convergencewithout discrete partitioning, finite strain parameters are D =1.64 <strong>and</strong> K = 0.58. When discrete partitioning accommodates50% <strong>of</strong> the lateral displacement, the finite strainparameters are D = 0.9 <strong>and</strong> K = 0.8.[35] Ductile partitioning splits the deformed domain intoa pure shear zone (PSZ) <strong>and</strong> a wrench-dominated zone(WDZ). We assume that the pure shear-across-strike shortening<strong>and</strong> elevation are homogeneously distributed acrossthe whole system, while simple shear-lateral displacement isaccommodated only in the WDZ. We examine development<strong>of</strong> strain parameters for different widths <strong>of</strong> the WDZ,expressed as ratio p 2 <strong>of</strong> the width <strong>of</strong> WDZ <strong>and</strong> the width<strong>of</strong> the whole transpressional zone. Such a partitioning <strong>of</strong>pure shear <strong>and</strong> simple shear within transpressional zones isresponsible for decomposition <strong>of</strong> the velocity gradienttensor into two separate tensors according to equations(C1) <strong>and</strong> (C2) (in Appendix C).[36] While the evolution <strong>of</strong> strain parameters in the PSZmay be obvious, the evolution <strong>of</strong> strain parameters in theWDZ zones is less so. The results <strong>of</strong> calculations for WDZzones <strong>of</strong> different width are shown in Figure 8. Fifty percent<strong>of</strong> ductile partitioning (Figure 8a) exhibits a similar pattern<strong>of</strong> strain parameters as a nonpartitioned system. Nevertheless,the domain with horizontal lineation <strong>and</strong> the domain<strong>of</strong> oblate symmetry are both slightly enlarged. This isrelated to the shift <strong>of</strong> the lineation switch, which occurs76


SCHULMANN ET AL.: STRAIN DISTRIBUTION ETG 6 - 9Figure 6. Superposition <strong>of</strong> strain intensity (D), strain symmetry (K ), <strong>and</strong> sample elevation for variousvalues <strong>of</strong> RFD <strong>and</strong> initial sample depth z 0 , plotted in terms <strong>of</strong> angle <strong>of</strong> convergence (a) <strong>and</strong> time parameter(k t ). The values <strong>of</strong> D show a strong dependence on R vd <strong>and</strong> time (but not much on a)fora =20° to 90°;fora


ETG 6 - 10 SCHULMANN ET AL.: STRAIN DISTRIBUTIONFigure 7. Relation between amount <strong>of</strong> discrete partitioning (amount <strong>of</strong> lateral displacement0accommodated by fault) <strong>and</strong> change <strong>of</strong> angle <strong>of</strong> convergence a <strong>and</strong> R vd ratio in transpression zone.Knowing the angle <strong>of</strong> convergence <strong>and</strong> the amount <strong>of</strong> discrete partitioning, new angle <strong>of</strong> convergence a 00<strong>and</strong> new R vd values can be depicted <strong>and</strong> used for estimation <strong>of</strong> strain parameters in diagram Figure 6.Thick black lines indicate amount <strong>of</strong> discrete partitioning for different active transpressive zones.mulate strain during the entire exhumation path. A rock’scapacity to accumulate strain is determined by its thermal,micro<strong>structural</strong> <strong>and</strong> rheological development. The bestexamples are syntectonically emplaced granites in transpressionalshear zones [e.g., Melka et al., 1992; Parry et al.,1997; Brown <strong>and</strong> Solar, 1998]. There is only a very shorttime <strong>of</strong> granite intrusion when the magma has enoughcrystals to record the deformation. Further cooling <strong>of</strong> themagma is responsible for hardening <strong>of</strong> granites <strong>and</strong> freezing<strong>of</strong> the fabrics at certain depth levels.[39] Numerous regional studies <strong>of</strong> transpressional zonesshow that rock complexes have experienced ductile deformationfollowed by late folding <strong>and</strong> brittle fracturing in thesame tectonic regime. This means that the accumulation <strong>of</strong>ductile strain ends at some depth below the surface dependingupon the ambient thermal regime. It is also common thatsamples originally located deep in lithospheric transpressionalzones have had their mineralogy changed due toretrograde metamorphism during exhumation. This meansthat deformation registered in transpressional zones wasdeveloping only during a small time interval during theconvergent activity.[40] To be able to correlate the external <strong>and</strong> internalparameters using the transpressional model, we need tospecify the time span (or the duration <strong>of</strong> the verticalelevation path) in which the accumulated strain is attributed.This means that in Figure 6 the time axis together withelevation contours (ratios <strong>of</strong> R vd /z 0 ) needs to be rescaled sothat times corresponding to the finite strain parameters, D<strong>and</strong> K, are higher.9. Discussion[41] Transpressional models in general assume that aweak <strong>and</strong> deformable zone bounded by the rigid walls <strong>of</strong>adjacent lithosphere is progressively shortened in the course<strong>of</strong> convergence. We can put forward a question as towhether the measured internal (microscopic) parameters <strong>of</strong>ancient transpressional zones (lineation, foliation, K <strong>and</strong> Dvalues) may be used to estimate the initial external (macroscopic)parameters (R vd , RFD, <strong>and</strong> a)?9.1. Ancient Zones[42] The strain symmetry parameter K is sensitive to theangle <strong>of</strong> convergence (a), whereas the strain intensityparameter D is well correlated with across-width shortening<strong>of</strong> a transpressional zone. On the basis <strong>of</strong> Figure 6 <strong>and</strong> usingmeasured natural finite strains <strong>and</strong> orientation <strong>of</strong> lineation in78


SCHULMANN ET AL.: STRAIN DISTRIBUTION ETG 6 - 11Figure 8. Superposition <strong>of</strong> strain intensity (D) <strong>and</strong> strain symmetry (K ) in terms <strong>of</strong> angle <strong>of</strong>convergence (a) for different widths <strong>of</strong> wrench-dominated zone: (a) 50%, (b) 20%, (c) 10%, <strong>and</strong> (d) 5%.For 100%, see Figure 6. These diagrams may correspond to different depth levels through thetranspressional zone with increasing width <strong>of</strong> wrench-dominated zone with depth in agreement with themodel <strong>of</strong> Pinet <strong>and</strong> Cobbold [1992].the rock (vertical for a >20°) we can conclude that in theframework <strong>of</strong> our model the maximum initial width <strong>of</strong>ancient transpressional zones should not exceed doubletheir actual width. Transpressive zones with horizontallineation corresponding to highly oblique convergence(low a) would have been 1.4 times wider originally. Thereforethe maximum initial width <strong>of</strong> ancient transpressionalzones could vary from 20 to 100 km. This width isapparently less than those <strong>of</strong> recently active transpressionalsystems.[43] Provided the average range <strong>of</strong> plate velocities <strong>of</strong>continental convergence did not change through geologicalhistory, from upper Proterozoic to Recent, we can estimatethe range <strong>of</strong> lifetimes <strong>of</strong> ancient transpressional zones. Asexamples <strong>of</strong> highly obliquely convergent transpressionalzones with large amounts <strong>of</strong> finite strain data, we canconsider the Central Bohemian Shear Zone (horizontallineation, K =0.8 1, vertical foliation, D =0.5 1.5[Rajlich et al., 1988]), <strong>and</strong> the transpressional zone developedin the eastern Variscan Culm Basin (horizontal lineation,K = 0.6–1.2, vertical foliation, D = 0.4 1.5[Rajlich, 1990]). These transpressional zones are associatedwith homogeneous ductile deformation over a width <strong>of</strong> upto 15 km. Because these transpressional zones are associatedwith low- to very low grade metamorphism, weconsider also an example <strong>of</strong> finite strain analysis from ahighly oblique transpressional zone developed at high-grademetamorphic conditions (kyanite-sillimanite zone) in theEastern Bohemian Massif which show horizontal lineationwith strain symmetry ranging from to oblate to plane strain<strong>and</strong> D from 0.6 to 1.2 [Schulmann, 1990]. As an example <strong>of</strong>transpression with a supposed high angle <strong>of</strong> convergence<strong>and</strong> with a large amount <strong>of</strong> data on finite strain, we considera transpressional zone in the Archean greenstone belt <strong>of</strong>Minnesota (vertical lineation <strong>and</strong> steep foliation, K variesaround plane strain, D ranges from 1.5 to 2.5 [Schultz-Ela<strong>and</strong> Hudleston, 1991]), where the deformed zone reaches anapproximate width <strong>of</strong> 30 km. Assuming plate velocities in79


ETG 6 - 12 SCHULMANN ET AL.: STRAIN DISTRIBUTIONthe range 1–5 cm yr 1 , the lifetimes <strong>of</strong> these shear zonesmight be in the range 0.5–4 Myr. These times are too shortfor an assumed duration <strong>of</strong> orogenic events. If we considerthat the duration <strong>of</strong> orogenic events is long (over an interval<strong>of</strong> several tens <strong>of</strong> million years), then the computed strainintensities for such given external parameters are unrealisticallyhigh (up to D = 100).9.2. Active Zones[44] Our calculations show that for a transpressional zonewith R vd < 0.2, the strain rate does not exceed a value <strong>of</strong> 6 10 15 s 1 for any angle <strong>of</strong> convergence (a). This R vd valuecorresponds to the active zones <strong>of</strong> Sumatra (d = 300 km, v =70 mm yr 1 , R vd = 0.23) <strong>and</strong> San Andreas (d = 200 km, v =50 mm yr 1 , R vd = 0.25). For a zone with value R vd = 0.48,corresponding to the Alpine Fault zone in New Zeal<strong>and</strong>,there will be maximum strain rate <strong>of</strong> 1.58 10 14 s 1distributed over a width <strong>of</strong> 100 km for a plate velocity <strong>of</strong> 48mm yr 1 . Zones with R vd >2 (small d, large v) show highstrain rate values <strong>of</strong> 3.2 10 14 s 1 <strong>and</strong> will apply fornarrow continental zones (10–20 km) with a mean platevelocity <strong>of</strong> 5 km yr 1 . Note that all <strong>of</strong> these values areconsistent with the strain rate estimates suggested by Carter<strong>and</strong> Tsenn [1987].[45] Assuming that the above mentioned strain valuesdevelop in large zones <strong>of</strong> homogeneous deformation, inFigure 6b we have plotted recent macroscopic convergenceparameters <strong>of</strong> well-known active transpressive zones. Wecan take the above listed macroscopic parameters <strong>of</strong> thesetranspressive zones <strong>and</strong> ask at what time are the averagestrain intensities developed? Figure 6b shows that theaverage strain <strong>of</strong> D = 1.2 would be produced in the Sumatrazone after 3.75 Myr, in San Andreas after 4.8 Myr, <strong>and</strong> inAlpine Fault Zone, New Zeal<strong>and</strong>, after 1.7 Myr.[46] The average strains, characteristic for most <strong>of</strong> thezones with dispersed deformation are achieved in 5 Myr forzones with R vd = 0.2 <strong>and</strong> after 10 Myr for zones with R vd =0.1. We note that the New Zeal<strong>and</strong> zone <strong>of</strong> distributedfaulting would have a strain intensity corresponding to D =13 after 5 Myr, while the San Andreas <strong>and</strong> Sumatra zoneswould remain within a realistic range <strong>of</strong> strain intensities (D<strong>of</strong> x to y). We note that for the entire lifetime <strong>of</strong> the SanAndreas system (20 Myr) only the last 5 Myr are attributedto dextral transpression [Walcott, 1993] which is caused byPacific plate rotation [Luyendyk et al., 1985]. Similarfeatures are reported from paleomagnetic investigations <strong>of</strong>the New Zeal<strong>and</strong> system [Walcott, 1987], <strong>and</strong> therefore anydirect application <strong>of</strong> the transpressional model to strainaccumulations in time is problematic.9.3. What Effects Could Explain These Discrepancies?[47] We are unable to correlate succinctly the external <strong>and</strong>internal parameters <strong>of</strong> homogeneous transpression. Thisinconsistency, apart from plate rotation, is well explainedby the three concepts <strong>of</strong> strain partitioning discussed above.Discrete partitioning results in general decrease <strong>of</strong> finitestrain accumulations <strong>and</strong> in increase <strong>of</strong> pure shear componentin deformed zone. The effects <strong>of</strong> ductile partitioningbecomes important for narrow wrench-dominated zone (lessthan 30% <strong>of</strong> the width <strong>of</strong> the whole transpressional zone)<strong>and</strong> is responsible for decrease <strong>of</strong> strain accumulation in thepure-shear-dominated zone <strong>and</strong> increase in the wrenchdominatedzone. We note that the pure-shear-dominatedzone exhibits all <strong>structural</strong> characteristics <strong>of</strong> frontal shortening.Viscosity partitioning is marked by different strainrates in domains <strong>of</strong> different viscosity leading to differentstrain accumulations. In strongly oblique zones (a


SCHULMANN ET AL.: STRAIN DISTRIBUTION ETG 6 - 13logical data allowing estimate <strong>of</strong> depth changes throughtime exist but so far are related to intrusions <strong>of</strong> magmas intranspressional regimes [e.g., Melka et al., 1992; Parry etal., 1997]. Future work should certainly concentrate onacquisition <strong>of</strong> fabric data together with detailed petrologicalinvestigations, to delimit the depth-temperature regime, <strong>and</strong>with geochronological investigations, to underst<strong>and</strong> theduration <strong>of</strong> transpressional orogeny.9.4. Importance <strong>of</strong> the Switch <strong>of</strong> Lineation[52] As an additional result <strong>of</strong> our modeling, we haveobtained information about the switch <strong>of</strong> lineation thatdeserves to be mentioned in conclusion. Analysis <strong>of</strong> thelineation produced by the classical transpression modelshows that the switch <strong>of</strong> lineation during progressive shortening<strong>of</strong> the zone is theoretically possible. However, thecorresponding strain intensities are very high <strong>and</strong> the degree<strong>of</strong> oblateness does not permit meaningful measurement <strong>of</strong> thechange in the linear fabric. An interesting observation is thatoblate fabrics are measured only exceptionally, more <strong>of</strong>tenwe have to deal with lineation (vertical or horizontal <strong>and</strong>close to plain strain symmetries). In contrast to the model, innature horizontal <strong>and</strong> vertical lineations are <strong>of</strong>ten observedsimultaneously in transpressional zones [Hudleston et al.,1988; Schultz-Ela <strong>and</strong> Hudleston, 1991; Melka et al., 1992;Tik<strong>of</strong>f <strong>and</strong> Greene, 1997]. This fact underlines the ideas <strong>of</strong>partitioning <strong>of</strong> the strain in transpressional zones [Tik<strong>of</strong>f <strong>and</strong>Greene, 1997] or perhaps to later superposition <strong>of</strong> simpleshear zones on already existing fabrics [Jones <strong>and</strong> Tanner,1995].Appendix A: Strain Parameters[53] In the transpressional model, computation <strong>of</strong> thethree-dimensional motion path <strong>of</strong> rock samples <strong>and</strong> strainparameters in the transpressive belt are based on thevelocity gradient tensor:0 10 _g 0L ¼ @ 0 _e 0 A: ðA1Þ0 0 _eThe pure shear strain rate component _e (corresponding tovertical extrusion <strong>and</strong> horizontal shortening) <strong>and</strong> simpleshear rate component _g (corresponding to lateral horizontaldisplacement) are calculated using the respective equations:_e ¼ n d sin a_g ¼ n cos a;d ðA3ÞðA2Þwhere v is the relative plate velocity, d is zone width, <strong>and</strong> ais the angle <strong>of</strong> convergence (Figure 1). In our model we willconsider an example <strong>of</strong> constant strain rate during temporaldevelopment <strong>of</strong> a transpressional zone, so for given periodfinite strain parameters can be obtained from followingfinite strain tensor:0_g1_ecos a ½ 1 exp ð _et ÞŠ01F ¼ @ 0 exp ð _et Þ 0 A: ðA4Þ0 0 exp ð_etÞDuring deformation we trace rock samples moving in thetranspressive zone (Figure 1b). At each position <strong>of</strong> a rocksample we calculate finite strain geometry, i.e., the principaldirections <strong>and</strong> the principal stretches by taking theeigenvectors <strong>and</strong> square roots <strong>of</strong> eigenvalues, respectively,<strong>of</strong> ‘‘Cauchy-Green tensor’’ FF T [Truesdell <strong>and</strong> Toupin,1960], which are further used to calculate the parameters K<strong>and</strong> D [Ramsay <strong>and</strong> Huber, 1983, pp. 201–202].[54] Strain intensity is expressed by the D value:qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiD ¼ R xy 1 2þ Ryz 1 2: ðA5ÞThe strain symmetry K value is calculated asK ¼ R xy 1R yz 1 ; ðA6Þwhere R xy = S 1 /S 2 <strong>and</strong> R yz = S 2 /S 3 represent elongationaspect ratios measured on orthogonal axes. Furthermore,when K = 0 a (S) planar fabric is indicated, <strong>and</strong> for K =1(LS) a plane strain fabric in which foliation is equally strongas lineation, <strong>and</strong> K = 1 for a linear fabric without foliation.[55] Vertical displacement produced by the pure shearcomponent can be expressed by a valuezt ðÞ¼z 0 exp ð_etÞ;ðA7Þwhere z 0 = z(t 0 ) is the initial vertical distance <strong>of</strong> a rocksample above a reference level (rigid floor depth) <strong>of</strong> zeroelevation.Appendix B: Discrete Partitioning[56] To evaluate strain parameters in a domain wherediscrete partitioning operate, we can use above definedapproach using following recalculated values <strong>of</strong> angle <strong>of</strong>convergence a <strong>and</strong> R vdtan a 0 ¼tan að1 p 1 ÞðB1ÞqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiR 0 vd ¼ R vd ð1 p 1 Þ 2 cos 2 a þ sin 2 a; ðB2Þwhere p 1 is ratio between fault-accommodated lateraldisplacement <strong>and</strong> total lateral displacement.Appendix C: Ductile Partitioning[57] Ductile partitioning in transpression zone results indecomposition <strong>of</strong> velocity gradient tensor into two componentsfor pure shear zone (L PSD ) <strong>and</strong> wrench-dominatedzone (L WDZ ) as follows:010 0 0L PSD ¼ @ 0 _e ð1 p 2 Þ 0 A0 0 _e ð1 p 2 ÞðC1Þ010 _g 0L WDZ ¼ @ 0 _ep 2 0 A; ðC2Þ0 0 _ep 281


ETG 6 - 14 SCHULMANN ET AL.: STRAIN DISTRIBUTIONwhere p 2 is ratio <strong>of</strong> the width <strong>of</strong> WDZ <strong>and</strong> the width <strong>of</strong> thewhole transpressional zone.Appendix D: Viscosity Partitioning[58] Viscosity partitioning assumes different strain ratesor different R vd , respectively, in two adjacent domains whilethe angle <strong>of</strong> convergence remains constant. The values <strong>of</strong>R vd areR m 2for low-viscosity domain <strong>and</strong>r mvd ¼ R vd1 þ p 3 r m 1 ðD1Þvd ¼ R 1vd1 þ p 3 r m 1 ðD2ÞR m 1for high-viscosity domain, whereis viscosity contrast.r m ¼ m 1; ðm m 1 > m 2 Þ2[59] Acknowledgments. The Czech National Foundation grant 205/98/K004, the Grant Agency <strong>of</strong> Charles University grant 296/1997/B GEO,the Schweizerische National Fonds <strong>and</strong> ETH Research credits are gratefullyacknowledged for financial support. We thank Basil Tik<strong>of</strong>f <strong>and</strong> an anonymousreviewer for very helpful reviews.ReferencesBrown, M., <strong>and</strong> G. S. Solar, Shear-zone systems <strong>and</strong> melts: Feedbackrelations <strong>and</strong> self- organization in orogenic belts, J. Struct. Geol., 20,211–227, 1998.Carter, N. L., <strong>and</strong> M. C. Tsenn, Flow properties <strong>of</strong> continental lithosphere,Tectonophysics, 136, 27–63, 1987.Cloetingh, S., J. D. Vanwees, P. A. V<strong>and</strong>erbeek, <strong>and</strong> G. Spadini, Role <strong>of</strong>pre-rift Rheology in kinematics <strong>of</strong> extensional basin formation—Constraintsfrom thermomechanical models <strong>of</strong> Mediterranean <strong>and</strong> intracratonicbasins, Mar. Pet. Geol., 12, 793–807, 1995.DeMets, C., R. G. Gordon, D. F. Argus, <strong>and</strong> S. Stein, Current plate motions,Geophys. J. Int., 101, 425–478, 1990.Dewey, J. F., Finite plate evolution: Some implications for the evolution <strong>of</strong>rock masses at plate margins, Am. J. Sci., 275-A, 260–284, 1975.Dutton, B. J., Finite strains in transpression zones with no boundary slip,J. Struct. Geol., 19, 1189–1200, 1997.Engl<strong>and</strong>, P., Comment on ‘‘Brittle failure in the upper mantle during extension<strong>of</strong> continental lithosphere’’ by Dale S. Sawyer, J. Geophys. Res.,91, 10,487–10,490, 1986.Fossen, H., <strong>and</strong> B. Tik<strong>of</strong>f, The deformation matrix for simultaneous simpleshearing, pure shearing <strong>and</strong> volume change, <strong>and</strong> its application to transpression-transtensiontectonics, J. Struct. Geol., 15, 413–422, 1993.Fossen, H., <strong>and</strong> B. Tik<strong>of</strong>f, Extended models <strong>of</strong> transpression <strong>and</strong> transtension,<strong>and</strong> application to tectonic setting, in Continental Transpressional<strong>and</strong> Transtensional Tectonics, edited by R. E. Holdsworth, R. A. Strachan,<strong>and</strong> J. F. Dewey, Geol. Soc. Spec. Publ., 135, 15–33, 1998.Harl<strong>and</strong>, W., Tectonic transpression in Caledonian Spitzbergen, Geol. Mag.,108, 27–42, 1971.Hrouda, F., Theoretical-models <strong>of</strong> magnetic-anisotropy to strain relationshiprevisited, Phys. Earth Planet. Inter., 77, 237–249, 1993.Hudleston, P. J., D. D. Schultz-Ela, <strong>and</strong> D. L. Southwick, Transpression inan Archean Greenstone-Belt, northern Minnesota, Can. J. Earth Sci., 25,1060–1068, 1988.Ježek, J., K. Schulmann, <strong>and</strong> A. B. Thompson, Modelling <strong>of</strong> strain partitioningin transpression zones, in Proceedings <strong>of</strong> the fourth annual conference<strong>of</strong> the International Association for Mathematical Geology, editedby A. Buccianti, G. Nardi, <strong>and</strong> R. Potenza, pp. 827 – 832, De Frede,Naples, Italy, 1998.Jones, R. R., <strong>and</strong> P. W. G. Tanner, Strain partitioning in transpression zones,J. Struct. Geol., 17, 793–802, 1995.Jones, R. R., R. E. Holdsworth, <strong>and</strong> W. Bailey, Lateral extrusion in transpressionzones: The importance <strong>of</strong> boundary conditions, J. Struct. Geol.,19, 1201–1217, 1997.Kirkwood, D., Strain partitioning <strong>and</strong> progressive deformation history in atranspressive belt, northern Appalachians, Tectonophysics, 241, 15–34,1995.Lisle, R., <strong>and</strong> J. Savage, Factors influencing Rock Competence - Data Froma Swedish Deformed Conglomerate, Geol. Foren. Stockholm Forh., 104,219–224, 1982.Luyendyk, B. P., M. J. Kamerling, R. R. Terres, <strong>and</strong> J. S. Hornafius, Simpleshear <strong>of</strong> southern California during Neogene time suggested by paleomagneticdeclinations, J. Geophys. Res., 90, 2454–2466, 1985.Martelat, J. E., C. Nicollet, J. M. Lardeaux, G. Vidal, <strong>and</strong> R. Rakotondrazafy,Lithospheric tectonic structures developed under high-grade metamorphismin the southern part <strong>of</strong> Madagascar, Geodin. Acta, 10, 94–114, 1997.Martelat, J. E., J. M. Lardeaux, C. Nicollet, <strong>and</strong> R. Rakotondrazafy, Strainpattern <strong>and</strong> late Precambrian deformation history in southern Madagascar,Precambrian Res., 102, 1–20, 2000.McKenzie, D., <strong>and</strong> R. L. Parker, The North Pacific: An example <strong>of</strong> tectonicson a sphere, Nature, 216, 1276–1280, 1967.Melka, R., K. Schulmann, B. Schulmannova, F. Hrouda, <strong>and</strong> M. Lobkowicz,The evolution <strong>of</strong> perpendicular linear fabrics in synkinematicallyemplaced tourmaline granite (central Moravia-Bohemian Massif ),J. Struct. Geol., 14, 605–620, 1992.Newman, J., W. M. Lamb, M. R. Drury, <strong>and</strong> R. L. M. Vissers, Deformationprocesses in a peridotite shear zone: Reaction-s<strong>of</strong>tening by an H 2 O-deficient,continuous net transfer reaction, Tectonophysics, 303, 193–222,1999.Nyman, M. W., R. D. Law, <strong>and</strong> S. S. Morgan, Conditions <strong>of</strong> contactmetamorphism,Papoose Flat Pluton, eastern California, USA-Implicationsfor cooling <strong>and</strong> strain histories, J. Metamorph. Geol., 13, 627–643, 1995.Parry, M., P. Stipska, K. Schulmann, F. Hrouda, J. Jezek, <strong>and</strong> A. Kroener,Tonalite sill emplacement at an oblique plate boundary; northeasternmargin <strong>of</strong> the Bohemian Massif, Tectonophysics, 280, 61–81, 1997.Pfiffner, O. A., <strong>and</strong> J. G. Ramsay, Constraints on geological strain rates:Arguments from finite strain states <strong>of</strong> naturally deformed rocks, J. Geophys.Res., 87, 311 – 321, 1982.Pili, E., Y. Ricard, J. M. Lardeaux, <strong>and</strong> S. M. F. Sheppard, Lithospheric shearzones <strong>and</strong> mantle-crust connections, Tectonophysics, 280, 15–29, 1997.Pinet, N., <strong>and</strong> P. R. Cobbold, Experimental insights into the partitioning <strong>of</strong>motion within zones <strong>of</strong> oblique subduction, Tectonophysics, 206, 371–388, 1992.Rajlich, P., Strain <strong>and</strong> tectonic styles related to Variscan transpression <strong>and</strong>transtension in the Moravo-Silesian Culmian basin Bohemian Massif,Czechoslovakia, Tectonophysics, 174, 351–367, 1990.Rajlich, P., K. Schulmann, <strong>and</strong> J. Synek, Strain analysis on conglomeratesfrom the central Bohemian shear zone, Krystalinikum, 19, 119–134,1988.Ramsay, J. G., <strong>and</strong> M. I. Huber, The Techniques <strong>of</strong> Modern StructuralGeology, vol. 1, Strain Analysis, Academic, San Diego, Calif., 1983.Ranalli, G., <strong>and</strong> D. C. Murphy, Rheological stratification <strong>of</strong> the lithosphere,Tectonophysics, 132, 281–295, 1987.Robin, P. Y., <strong>and</strong> A. R. Cruden, Strain <strong>and</strong> vorticity patterns in ideallyductile transpression zones, J. Struct. Geol., 16, 447–466, 1994.Royden, L., Coupling <strong>and</strong> decoupling <strong>of</strong> crust <strong>and</strong> mantle in convergentorogens: Implications for strain partitioning in the crust, J. Geophys. Res.,101, 17,679–17,705, 1996.S<strong>and</strong>erson, D. J., <strong>and</strong> W. R. D. Marchini, Transpression, J. Struct. Geol, 6,449–458, 1984.Schulmann, K., Fabric <strong>and</strong> kinematic study <strong>of</strong> the bites orthogneiss (southwesternMoravia)—Result <strong>of</strong> large-scale northeastward shearing parallelto the Moldanubian Moravian boundary, Tectonophysics, 177, 229–244,1990.Schulmann, K., R. Melka, M. Lobkowicz, P. Ledru, J. M. Lardeaux, <strong>and</strong>A. Autran, Contrasting styles <strong>of</strong> deformation during progressive nappestacking at the southeastern margin <strong>of</strong> the Bohemian Massif (ThayaDome), J. Struct. Geol., 16, 355–370, 1994.Schultz-Ela, D. D., <strong>and</strong> P. J. Hudleston, Strain in an Archean greenstonebelt<strong>of</strong> Minnesota, Tectonophysics, 190, 233–268, 1991.Stuwe, K., <strong>and</strong> T. D. Barr, On uplift <strong>and</strong> exhumation during convergence,Tectonics, 17, 80–88, 1998.Teyssier, C., <strong>and</strong> B. Tik<strong>of</strong>f, Study <strong>of</strong> strike-slip partitioning in Californiaresolves San Andreas discrepancy <strong>and</strong> constrains lithospheric structure,Geol. Soc. Am. Abstr. Programs, 29, 347, 1997.Teyssier, C., B. Tik<strong>of</strong>f, <strong>and</strong> M. Markley, Oblique plate motion <strong>and</strong> continentaltectonics, Geology, 23, 447–450, 1995.Thompson, A. B., K. Schulmann, J. Jezek, <strong>and</strong> V. Tolar, Thermally s<strong>of</strong>tenedcontinental extensional zones: Arcs <strong>and</strong> rifts as precursors to thickenedorogenic zones, Tectonophysics, 332, 115–141, 2001.82


SCHULMANN ET AL.: STRAIN DISTRIBUTION ETG 6 - 15Tik<strong>of</strong>f, B., <strong>and</strong> D. Greene, Stretching lineations in transpressional shearzones: An example from the Sierra Nevada Batholith California, J. Struct.Geol., 19, 29–39, 1997.Tik<strong>of</strong>f, B., <strong>and</strong> C. Teyssier, Strain modeling <strong>of</strong> displacement-field partitioningin transpression orogens, J. Struct. Geol., 16, 1575–1588, 1994.Tommasi, A., <strong>and</strong> A. Vauchez, Continental-scale rheological heterogeneities<strong>and</strong> complex intraplate tectono-metamorphic patterns: Insights froma case study <strong>and</strong> <strong>numerical</strong> models, Tectonophysics, 279, 327 – 350,1997.Truesdell, C. A., <strong>and</strong> R. A. Toupin, The classic field theory, in Encyclopedia<strong>of</strong> Physics, vol.III,Principles <strong>of</strong> Classical Mechanics <strong>and</strong> FieldTheory, edited by S. Flügge, pp. 226–793, Springer-Verlag, New York,1960.Vauchez, A., <strong>and</strong> A. Nicolas, Mountain building-Strike-parallel motion <strong>and</strong>mantle anisotropy, Tectonophysics, 185, 183–201, 1991.Vauchez, A., A. Tommasi, <strong>and</strong> G. Barruol, Rheological heterogeneity, mechanicalanisotropy <strong>and</strong> deformation <strong>of</strong> the continental lithosphere, Tectonophysics,296, 61–86, 1998.Walcott, D., Neogene tectonics <strong>and</strong> kinematics <strong>of</strong> western North America,Tectonics, 12, 326–333, 1993.Walcott, R. I., Geodetic strain <strong>and</strong> the deformational history <strong>of</strong> the NorthIsl<strong>and</strong> <strong>of</strong> New Zeal<strong>and</strong> during the Late Cainozoic, Philos. Trans. R. Soc.London, Ser. A, 321, 163–181, 1987.White, J. C., <strong>and</strong> C. K. Mawer, Deep-crustal deformation textures alongmegathrusts from Newfoundl<strong>and</strong> <strong>and</strong> Ontario—Implications for micro<strong>structural</strong>preservation, strain rates, <strong>and</strong> strength <strong>of</strong> the lithosphere, Can.J. Earth Sci., 29, 328–337, 1992.Wood, D. S., Current views <strong>of</strong> the development <strong>of</strong> slaty cleavage, Annu.Rev. Earth Planet. Sci., 2, 369–401, 1974.J. Ježek, Institute <strong>of</strong> Applied Mathematics <strong>and</strong> Computer Science,Faculty <strong>of</strong> Science, Charles University, Albertov 6, 128 43 Prague, CzechRepublic. ( jezek@natur.cuni.cz)O. Lexa <strong>and</strong> K. Schulmann, Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology,Faculty <strong>of</strong> Science, Charles University, Albertov 6, 128 43 Prague, CzechRepublic. (lexa@natur.cuni.cz; schulman@natur.cuni.cz)A. B. Thompson, Institut für Mineralogie und Petrographie, ETHZentrum, NO E 64, CH-8092 Zürich, Switzerl<strong>and</strong>. (thompson@erdw.ethz.ch)83


TECTONICS, VOL. 22, NO. 6, 1066, doi:10.1029/2002TC001472, 2003Cretaceous collision <strong>and</strong> indentation in the West Carpathians:View based on <strong>structural</strong> analysis <strong>and</strong> <strong>numerical</strong> modelingOndrej Lexa <strong>and</strong> Karel SchulmannInstitute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University, Prague, Czech RepublicJosef JežekInstitute <strong>of</strong> Applied Mathematics <strong>and</strong> Computer Science, Charles University, Prague, Czech RepublicReceived 29 October 2002; revised 13 March 2003; accepted 30 July 2003; published 21 November 2003.[1] A model <strong>of</strong> indentation <strong>of</strong> a rigid promontory intoweak metasedimentary rocks during Cretaceousconvergence is suggested for tectonic evolution <strong>of</strong>southern part <strong>of</strong> West Carpathians. An early arcuatecleavage fan has developed in front <strong>of</strong> northwardmoving southern rigid basement block. The interaction<strong>of</strong> the moving indenter with western stationarybasement promontory is responsible for development<strong>of</strong> the boundary-parallel shear zone along which themain southern indenter is shifted to the east. This resultsin development <strong>of</strong> a new steep transpressional cleavageoverprinting the early fabric. Eastward displacement <strong>of</strong>the southern indenter causes the development <strong>of</strong> a thrustzone parallel to the margin <strong>of</strong> the eastern stationarypromontory. A proposed <strong>numerical</strong> model <strong>of</strong> thedeformation <strong>of</strong> a thin viscous sheet in front <strong>of</strong> ovalrigid indenter reliably simulates the development <strong>of</strong> theobserved deformation pattern. Modeled discretepartitioning between the western promontory <strong>and</strong> theindenting block fully agrees with the observedsecondary cleavage associated with the transpressionalshear zone. Our <strong>numerical</strong> model interconnects thiscomplex kinematic frame with finite strain pattern,which was to date possible only for simple boundaryconditions. In addition, the model explains thepolyphase cleavage patterns in terms <strong>of</strong> complexshapes <strong>of</strong> promontories <strong>and</strong> changes in movements <strong>of</strong>indenting blocks. INDEX TERMS: 5475 Planetology: SolidSurface Planets: Tectonics (8149); 8005 Structural Geology: Folds<strong>and</strong> folding; 8020 Structural Geology: Mechanics; 8110Tectonophysics: Continental tectonics—general (0905);KEYWORDS: Cretaceous collision, indentation, west Carpathians,<strong>numerical</strong> model. Citation: Lexa, O., K. Schulmann, <strong>and</strong> J. Ježek,Cretaceous collision <strong>and</strong> indentation in the West Carpathians: Viewbased on <strong>structural</strong> analysis <strong>and</strong> <strong>numerical</strong> modeling, Tectonics,22(6), 1066, doi:10.1029/2002TC001472, 2003.1. Introduction[2] Existing interpretations <strong>of</strong> Mesozoic tectonic evolution<strong>of</strong> Alpine <strong>and</strong> Carpathian chains are based on broadCopyright 2003 by the American Geophysical Union.0278-7407/03/2002TC001472$12.005 - 1knowledge <strong>of</strong> metamorphic <strong>and</strong> <strong>structural</strong> data obtainedfrom studies <strong>of</strong> internal parts <strong>of</strong> these orogenic belts[Genser et al., 1996; Schmid et al., 1996; Trumpy, 1973].These data underline subduction-related tectonics associatedwith lithospheric-scale subduction <strong>of</strong> oceanic <strong>and</strong> Europeanlithosphere below the African indenter [Allem<strong>and</strong> <strong>and</strong>Lardeaux, 1997]. The Mesozoic evolution <strong>of</strong> the WestCarpathian belt is an integral part <strong>of</strong> the closing <strong>of</strong> theTethyan ocean, being characterized by Jurassic subduction<strong>and</strong> Late Jurassic exhumation <strong>of</strong> high pressure rocks[Faryad, 1995; Faryad <strong>and</strong> Henjes-Kunst, 1997; Maluskiet al., 1993; Mock <strong>and</strong> Reichwalder, 1992].[3] The West Carpathian structure is traditionally interpretedas a result <strong>of</strong> Cretaceous crustal shortening <strong>of</strong>Variscan basement crustal segments associated withdécollement <strong>of</strong> Mesozoic sedimentary sequences. TheMesozoic strata originally formed basins separating individualcrustal segments, which were inverted <strong>and</strong> passivelytransported toward the north in the form <strong>of</strong> twolarge-scale nappes (the lower Krížna <strong>and</strong> upper Chočnappes) [Andrusov, 1936, 1958; Biely et al., 1968;Plašienka, 1991; Plašienka et al., 1997]. The Cretaceoustectonic evolution in West Carpathians was mostly studiedin Mesozoic rocks with well-known stratigraphy[Biely <strong>and</strong> Bystrický, 1967; Bystrický, 1967; Vozár,1978]. Consequently, the Cretaceous stacking <strong>of</strong> Mesozoicsequences, the kinematics <strong>and</strong> displacement <strong>of</strong>Mesozoic nappes <strong>and</strong> the degree <strong>of</strong> their metamorphicreequilibration are well defined <strong>and</strong> understood [cf.Plašienka, 1995]. However, the Alpine <strong>structural</strong> <strong>and</strong>metamorphic evolution <strong>of</strong> pre-Mesozoic crystalline rocksis poorly described due to difficulties in distinguishingVariscan <strong>and</strong> Alpine fabrics.[4] This paper presents new information on <strong>structural</strong> <strong>and</strong>mechanical behavior <strong>of</strong> the southern part <strong>of</strong> West Carpathianpre-Mesozoic crystalline segments during Cretaceousconvergence. We have investigated the polyphase cleavagepatterns developed in low-grade metasedimentary EarlyPaleozoic sequences. It is generally accepted that suchcleavage patterns reflect the geometry <strong>and</strong> direction<strong>of</strong> movement <strong>of</strong> adjacent rigid blocks [Sintubin, 1999;Woodcock et al., 1988]. The detailed knowledge <strong>of</strong> shapes<strong>of</strong> rigid Variscan promontories <strong>and</strong> the geometry <strong>and</strong>superposition <strong>of</strong> cleavage patterns allow modeling <strong>of</strong> thecollisional process in time. We use the finite element<strong>numerical</strong> method introduced by Engl<strong>and</strong> et al. [1985] to85


5 - 2 LEXA ET AL.: COLLISION IN WEST CARPATHIANSmodel the finite strain pattern in time <strong>and</strong> space for complexboundary conditions.2. Geological Setting[5] The West Carpathians have been traditionally dividedinto the outer <strong>and</strong> inner <strong>structural</strong> zones. However, takinginto account the recent studies <strong>of</strong> Mesozoic evolution, thetriple division into Inner, Central <strong>and</strong> Outer West Carpathiansis commonly accepted (Figure 1) [see Plašienka etal., 1997]. The Outer West Carpathians (OWC) are representedby flysch dominated Paleogene formations <strong>and</strong>Mesozoic rocks lacking pre-Mesozoic basement. The Inner<strong>and</strong> Central West Carpathians (IWC <strong>and</strong> CWC), separatedby a Late Jurassic oceanic suture, are composed <strong>of</strong> a pre-Late Cretaceous imbricated nappe system comprisingcrystalline basement units with characteristic Late Paleozoic<strong>and</strong> Mesozoic sequences [Matějka <strong>and</strong> Andrusov, 1931] <strong>and</strong>postnappe Late Cretaceous to Neogene sedimentary <strong>and</strong>volcanic formations.2.1. Pre-Cretaceous Geology <strong>of</strong> the Studied Area[6] The studied area is located in the southern part <strong>of</strong> theCWC <strong>and</strong> comprises three major lithological <strong>and</strong> tectonometamorphicunits (Figure 1). From the north to the south<strong>and</strong> from the bottom to the top they are (1) Variscancrystalline basement (Vepor Unit) with Late Paleozoic <strong>and</strong>Mesozoic cover sequences, (2) Early to Late Paleozoic,basinal, mostly low grade turbiditic sequences (GemerUnit), <strong>and</strong> (3) Mesozoic accretionary wedge containingblueschist facies relics overlain by (4) flat, nonmetamorphosedSilica nappe.2.1.1. Variscan Crystalline Basement: The Vepor Unit[7] The crustal rocks <strong>of</strong> the Vepor basement are composed<strong>of</strong> two contrasting Variscan metamorphic domainsexhibiting pre-Alpine thrust tectonics. The <strong>structural</strong>lylower domain, generally dipping to the north, is composed<strong>of</strong> medium-grade schists exhibiting peak PT conditions inKy-St micaschists corresponding to maximum <strong>of</strong> 10 kbar<strong>and</strong> 550–600°C [Korikovskij et al., 1989; Méres <strong>and</strong>Hovorka, 1991]. The classic Barrow type metamorphiczonation is difficult to establish due to the Alpine greenschistfacies overprint but in the eastern part <strong>of</strong> the Veporbasement (Čierna Hora Mountains), a metamorphic zonationranging from the biotite zone in the south to thestaurolite zone in the north is documented [Jacko et al.,1990; Korikovskij et al., 1990]. The <strong>structural</strong>ly highercrystalline unit is represented by a domain <strong>of</strong> heterogeneouspara- <strong>and</strong> ortho-derived migmatites intruded byporphyritic to medium-grained peraluminous granites.The Variscan age <strong>of</strong> metamorphism is documented byTh-U-Pb dating <strong>of</strong> monazite (370–350 Ma [Janák et al.,2001a]) <strong>and</strong> by sporadically preserved 40 Ar/ 39 Ar coolingages ranging from 358 to 312 Ma [Dallmeyer et al.,1996]. High-grade fabrics represented by compositionallayering <strong>and</strong> stromatitic b<strong>and</strong>ing in migmatites are dippingeither to the NW or to the N under steep to mediumangles. Peak temperature conditions yield 680–730°C atpressures <strong>of</strong> 4–6 kbar (Figure 2) [Siman et al., 1996].These PT conditions most likely correspond to lateexhumation <strong>and</strong> decompression melting. This is deducedfrom other parts <strong>of</strong> West Carpathians where relics <strong>of</strong>high-pressure assemblages preserved in similar types <strong>of</strong>migmatites were found [Hovorka <strong>and</strong> Méres, 1989;Janák et al., 1996].[8] The southern Vepor Variscan crystalline basement isunconformably covered by Late Carboniferous (Stephanian)s<strong>and</strong>stones <strong>and</strong> shales [Pl<strong>and</strong>erová <strong>and</strong> Vozárová, 1978].These metasediments were intruded by granitoids that wereresponsible for contact metamorphism ranging from biotiteto cordierite zones. Contact metamorphic conditions <strong>of</strong>500°C <strong>and</strong> 2 kbar were established by Vozárová [1990].Vozárová also suggests that contact metamorphism overprintedregional greenschist facies assemblages <strong>of</strong> Carboniferousmetasediments. The Vepor crystalline complexes aswell as the Late Carboniferous sequences are overlain byPermian <strong>and</strong> Triassic clastics <strong>and</strong> locally <strong>of</strong> Middle <strong>and</strong> LateTriassic carbonates.2.1.2. Early to Late Paleozoic Gemer Unit[9] The Gemer Unit consists <strong>of</strong> three principal groups thatdiffer in lithology <strong>and</strong> metamorphic grade. The amphibolitefacies metamorphic conditions are reported from the KlátovGroup [Faryad, 1990], which tectonically overlies thegreenschist facies metabasites <strong>and</strong> phyllites <strong>of</strong> the RakovecGroup. The Variscan metamorphic conditions correspond to440–480°C at 6–8 kbar for the northern part <strong>and</strong> 350–430°C at 4–5 kbar (Figure 2) for the southern part <strong>of</strong> theRakovec Group [Faryad <strong>and</strong> Bernhardt, 1996; Vozárová,1993]. The metamorphic foliation is represented by compositionallayering in metabasites dipping to the NNW undershallow to intermediate angles. The tectonically lowermostGelnica Group builds the major part <strong>of</strong> the Gemer Unit. Thelower part <strong>of</strong> this unit consists <strong>of</strong> a thick turbiditic sequence<strong>of</strong> Ordovician age [Soták et al., 1999; Vozárová etal., 1999]gradually passing into volcanosedimentary sequencesaccompanied by products <strong>of</strong> massive rhyolite-dacite volcanism<strong>of</strong> probably Silurian-Devonian age [Cambel et al.,1990]. The top <strong>of</strong> the Gelnica Group is composed <strong>of</strong>phyllites <strong>and</strong> black shales containing sporadic carbonatelenses. The metamorphic conditions indicate temperatures<strong>of</strong> 350–400°C <strong>and</strong> pressures <strong>of</strong> 2.5–3.5 kbar [Faryad, 1992,1994] for regional metamorphism, which is overprintedby contact metamorphism (450–550°C <strong>and</strong> 1.5–2 kbar[Faryad, 1992]) associated with intrusion <strong>of</strong> granites incentral part <strong>of</strong> the Gelnica Group. Strong greenschist faciesFigure 1. (opposite) Geological <strong>and</strong> <strong>structural</strong> map <strong>of</strong> the studied area with trajectories <strong>of</strong> cleavage based on this work <strong>and</strong>that <strong>of</strong> Snopko <strong>and</strong> Reichwalder [1970]. The inset in the upper left corner shows the Carpathian arc with location <strong>of</strong> thestudied area.86


LEXA ET AL.: COLLISION IN WEST CARPATHIANS 5 - 387


5 - 4 LEXA ET AL.: COLLISION IN WEST CARPATHIANSFigure 2. Idealized E-W cross section showing pre-Cretaceous <strong>structural</strong> <strong>and</strong> metamorphic pattern <strong>of</strong>the studied area. Insets show Variscan (Gelnica, Rakovec, <strong>and</strong> Klátov Formations) <strong>and</strong> Jurassic (Bôrkanappe) metamorphic PT conditions. The eastern ‘‘unknown’’ block represents hypothetical active marginduring closure <strong>of</strong> Meliata ocean. The legend is same as for Figure 1.metamorphic foliation <strong>and</strong> relics <strong>of</strong> sedimentary bedding aredipping conformably to the NNW. The metamorphic grade<strong>and</strong> degree <strong>of</strong> <strong>structural</strong> transposition are gradually vanishingto the south.[10] Already consolidated Rakovec <strong>and</strong> Gelnica Groupsare discordantly covered by Early Carboniferous flysch <strong>and</strong>Permian clastics <strong>of</strong> ‘‘red bed’’ type. Sedimentation wasterminated by the development <strong>of</strong> Late Permian-EarlyTriassic evaporite formations. The metamorphism <strong>of</strong> Permianclastics yielded temperatures <strong>of</strong> 200–250°C [Šucha <strong>and</strong>Eberl, 1992].2.1.3. Mesozoic Meliata Accretionary Wedge <strong>and</strong>Superficial Triassic Nappe[11] Mesozoic sequences are located mainly along thesouthern margin <strong>of</strong> the Gemer Unit, locally rimming itsnorthern margin or forming klippens on the Gemer Unit.The bottom part <strong>of</strong> the Meliata accretionary wedge isformed by the sequence <strong>of</strong> thrust sheets (initially thesouthern slope <strong>of</strong> the European continent) consisting <strong>of</strong>thinned continental margin, Permian clastics, Triassic limestones<strong>and</strong> blueschist facies metabasalts (Figure 2) (380–460°C at 9–13 kbar [Faryad, 1995]).[12] These rocks are overlain by subduction-relatedmelange composed <strong>of</strong> deep-water oceanic sediments <strong>of</strong> theMeliata ocean <strong>and</strong> relics <strong>of</strong> oceanic mantle rocks. Theuppermost part <strong>of</strong> the accretionary wedge is composed <strong>of</strong>Triassic (Turňa) carbonates, shales <strong>and</strong> pelagic sediments <strong>of</strong>northern slope <strong>of</strong> the Apulian continent. The high pressurerocks <strong>of</strong> the Meliata accretionary wedge exhibit strongductile <strong>and</strong> polyphase deformation <strong>of</strong> all lithologies. Themylonitic foliation is generally dipping to the east, bears anintense southeast plunging stretching lineation <strong>and</strong> kinematicindicators, such as sigmoidal pebbles <strong>of</strong> metabasalts, whichsuggest top to the northwest transport. The Turňa shalesshow very low grade metamorphic overprint <strong>and</strong> locallydeveloped slaty cleavage subparallel to the bedding.[13] The accretionary wedge <strong>and</strong> the underlying GemerUnit are tectonically overlain by extensive horizontallylying Silica nappe derived from the Apulian shelf. The base<strong>of</strong> the nappe is composed <strong>of</strong> Late Permian-Early Triassicevaporites <strong>and</strong> shales followed by Middle <strong>and</strong> Late Triassiccarbonates, which were affected by brittle faulting [Hók etal., 1995].2.2. Summary <strong>of</strong> Pre-Cretaceous TectonometamorphicEvolution <strong>of</strong> the Studied Area[14] The metamorphic <strong>and</strong> <strong>structural</strong> patterns describedabove are consistent with north-south polarity <strong>of</strong> the Variscanorogenic evolution marked by southward thrusting <strong>of</strong>deep-seated lower crustal <strong>and</strong> middle crustal complexesover low-grade forel<strong>and</strong>. The Vepor Unit is regarded as arelic <strong>of</strong> the Variscan internal domain in which the anatecticlower crust is thrust over the middle crustal Barroviancomplex. The original Paleozoic relationship between Vepor<strong>and</strong> Gemer Units is largely obscured by later Mesozoicevolution. North dipping pre-Mesozoic fabrics <strong>and</strong> increasein metamorphic grade from the south to the north in both theGemer <strong>and</strong> Vepor Units suggest an existence <strong>of</strong> N-S polarity<strong>of</strong> Paleozoic orogeny [Faryad, 1990]. On the basis <strong>of</strong> thesedata, the Klátov <strong>and</strong> Rakovec groups may be interpreted asrelics <strong>of</strong> an Early Paleozoic basin underthrust beneath thehigh-grade rocks <strong>of</strong> the internal domain represented by theVepor Unit exposed in the north. In our model the wholenappe stack was further thrust over Early Paleozoic basinrepresented by the Gelnica Group, which is probablyunderlain by a Neo-Proterozoic basement (Figure 2). Thelack <strong>of</strong> thrust-related structures in Middle Carboniferousmetasediments indicates that the nappe stacking occurred inpre-Westfalian times.[15] The southward thrusting <strong>of</strong> high-grade gneisses overthe low-grade mostly metasedimentary forel<strong>and</strong> formed twopromontories separated by embayment <strong>of</strong> weakly metamorphosedsediments. This irregular geometric distribution <strong>of</strong>gneissic complexes <strong>and</strong> s<strong>of</strong>t sediments played a critical rolein the subsequent Mesozoic tectonic evolution.[16] The last major pre-Cretaceous tectonic event responsiblefor the final <strong>structural</strong> pattern was the Jurassic southeastwardsubduction (in recent coordinates) <strong>of</strong> the Meliataocean <strong>and</strong> the southern passive margin <strong>of</strong> the Europeanplatform. This process resulted in formation <strong>of</strong> an accretionarywedge <strong>and</strong> its northwestward obduction over both88


LEXA ET AL.: COLLISION IN WEST CARPATHIANS 5 - 5the Gemer <strong>and</strong> Vepor Units during the Late Jurassic (150–160 Ma [Dallmeyer et al., 1996; Faryad <strong>and</strong> Henjes-Kunst,1997; Maluski et al., 1993]) (Figure 2). In the southern part<strong>of</strong> the Gemer Unit, sedimentary bedding is well preserved<strong>and</strong> folded by large-scale open folds with N-S trendinghinges (Figure 3). This folding is connected with development<strong>of</strong> spaced cleavage steeply dipping to the east suggestingthat the thinned continental margin was intensivelyreworked during Jurassic subduction processes.3. Cretaceous Polyphase Structural Evolution:Collisional Stage[17] The Cretaceous collisional evolution <strong>of</strong> the Gemer<strong>and</strong> Vepor Units is marked by four major distinct tectonicevents: (1) formation <strong>of</strong> the Gemer Cleavage Fan (GCF)structure affecting the central part <strong>of</strong> the Gemer Unit,(2) extensional deformation <strong>of</strong> the western Vepor promontory,(3) transpressional shearing affecting the western Veporpromontory <strong>and</strong> development <strong>of</strong> the Trans-Gemer ShearZone (TGSZ), <strong>and</strong> (4) extrusion <strong>of</strong> the Gemer Unit over theeastern Vepor promontory along the Eastern Gemer Thrust(EGT).[18] In order to evaluate strain variations in conglomerateswe used published data [Németh et al., 1997], while forslates the degree <strong>of</strong> deformation associated with development<strong>of</strong> crenulation cleavage was evaluated using criteriaintroduced by Bell <strong>and</strong> Rubenach [1983]. In our work weused five stages <strong>of</strong> crenulation cleavage development[Passchier <strong>and</strong> Trouw, 1996, Figure 4.17ab] to semiquantitativelycharacterize the strain gradient in studied area.3.1. Early Cretaceous Deformation:Gemer Cleavage Fan[19] The Gemer Cleavage Fan represents the most spectacularstructure overprinting the pre-Mesozoic metamorphicfabric <strong>of</strong> the Gelnica <strong>and</strong> Rakovec Groups as well asJurassic fabric to the south. This cleavage forms an asymmetricpositive fan structure developed across the entirelength <strong>of</strong> the Gemer Unit (Figure 3). The intensity <strong>and</strong>metamorphic grade <strong>of</strong> the cleavage are highest in a 5 kmwide, E-W oriented axial zone <strong>of</strong> the fan structure. Here, theonly relics <strong>of</strong> pre-Mesozoic fabric in the form <strong>of</strong> rootlessfolds are preserved in intensively developed steep slatycleavage corresponding to stages 4 <strong>and</strong> 5 <strong>of</strong> Bell <strong>and</strong>Rubenach [1983] classification (Figure 3). Finite strainmeasurements <strong>of</strong> Németh et al. [1997] show oblate finitestrain symmetry <strong>and</strong> rather moderate strain intensities. Thiszone is characterized by the presence <strong>of</strong> numerous bodies <strong>of</strong>granites. These intrusions have been originally consideredCretaceous in age [Kantor, 1957; Máška, 1957; Vozár et al.,1996] <strong>and</strong> later, based on monazite (273 Ma [Finger <strong>and</strong>Broska, 1999]) <strong>and</strong> zircon U-Pb (275–245 Ma [Poller et al.,2002]) dating as Permian to Early Triassic. The maincleavage <strong>of</strong> the GCF is in the axial zone affected by kinkb<strong>and</strong>s with kink planes perpendicular or oblique to stronglydeveloped vertical anisotropy (Figure 5a). These structuresare interpreted as a result <strong>of</strong> vertical shortening <strong>of</strong> steepcleavage associated with the evolution <strong>of</strong> GCF.[20] The intensity <strong>and</strong> degree <strong>of</strong> metamorphism <strong>of</strong> thenorth dipping, spaced cleavage <strong>of</strong> southern flank <strong>of</strong> GCFrapidly decrease to the south. The cleavage intensity correspondsto stages 1 <strong>and</strong> 2 <strong>of</strong> Bell <strong>and</strong> Rubenach’s [1983]classification. Here, the Cretaceous deformation is so weakthat the structures associated with Jurassic obduction <strong>of</strong>Meliata accretionary wedge are well preserved. In thenorthern flank <strong>of</strong> GCF, the cleavage is dipping to the south.The dip angle gradually decreases to the north in conjunctionwith decreasing intensity <strong>of</strong> cleavage developmentcorresponding to the stages 1 <strong>and</strong> 2 <strong>of</strong> Bell <strong>and</strong> Rubenach[1983] (Figure 5b). In northern parts <strong>of</strong> the Gelnica Group,the north vergent kink b<strong>and</strong>s developed mainly withinincompetent shales while competent lithologies like quartzites<strong>and</strong> volcanics were gently folded by open folds withwavelengths <strong>of</strong> a few hundreds <strong>of</strong> meters (Figure 3a). Thedecrease in cleavage intensity is even more pronounced inamphibolite <strong>and</strong> greenschist facies metabasites <strong>of</strong> the RakovecGroup where a spaced to fracture cleavage is developedindicating very low finite strains [Passchier <strong>and</strong> Trouw,1996] (Figures 3a <strong>and</strong> 5c).[21] The Late Carboniferous <strong>and</strong> Permian cover <strong>of</strong> theGemer Unit, directly overlying the Rakovec Group, is alsoaffected by GCF. In this domain, the deformation is stronglyheterogeneous leading to the development <strong>of</strong> small-scale(up to several hundreds <strong>of</strong> meters) positive cleavage fanstructures. Carboniferous conglomerates contain pebbles <strong>of</strong>foliated metabasites <strong>of</strong> the Rakovec Group. However, thesecondary cleavage known from the Rakovec Group hasnever been discovered in the pebbles. This means that thesame post-Permian cleavage affects the Gemer Unit <strong>and</strong>Late Paleozoic sediments.[22] The features <strong>of</strong> GCF described above are developedprominently along the central part <strong>of</strong> the Gemer Unit(Figure 4a). Lateral extension <strong>of</strong> GCF toward western <strong>and</strong>eastern Vepor promontories (Figures 3a <strong>and</strong> 4b) is markedby change in cleavage trend, so that it becomes parallel totheir boundaries. In addition, the positive fan structuredisappears <strong>and</strong> the cleavage is dominantly vertical <strong>and</strong> veryintense. Triassic <strong>and</strong> Jurassic sediments <strong>of</strong> the Meliataaccretionary wedge in the front <strong>of</strong> the western Veporpromontory exhibit a strong cleavage development <strong>and</strong>reworking <strong>of</strong> Jurassic fabrics in continuation <strong>of</strong> GCF.Structural succession as well as 40 Ar/ 39 Ar cooling agesranging from 106 to 82 Ma from Gemer Unit [Dallmeyeret al., 1996] are our major arguments for Cretaceousshortening <strong>of</strong> the Gemer Unit associated with the development<strong>of</strong> GCF.3.2. The First Cretaceous Deformation:Vepor Extensional Mylonitization[23] In contrast to compressional deformation <strong>of</strong> theGemer Unit, the earliest Cretaceous structures developedin the Vepor basement are associated with extensionaltectonics (Figure 4a). This is manifested by mylonitization<strong>of</strong> basement rocks <strong>and</strong> Permian-Triassic cover, characterizedby the development <strong>of</strong> anastomosed network <strong>of</strong>greenschist facies shear zones. The intensity <strong>of</strong> mylonitizationis highest in the cover sequences <strong>and</strong> gradually89


5 - 6 LEXA ET AL.: COLLISION IN WEST CARPATHIANS90


LEXA ET AL.: COLLISION IN WEST CARPATHIANS 5 - 7decreases toward the deeper part <strong>of</strong> the Vepor basement,where the Variscan high-grade fabrics are preserved [Simanet al., 1996]. The Cretaceous age <strong>of</strong> this deformation isconfirmed by several Ar/Ar cooling ages <strong>of</strong> micas rangingfrom 90 to 85 Ma [Dallmeyer et al., 1996; Janák et al.,2001b; Kováčiketal., 1996; Maluski et al., 1993].[24] The style <strong>of</strong> deformation changes toward the NEedge <strong>of</strong> the Vepor basement, where narrow normal shearzone marked by strongly noncoaxial deformation associatedwith top-to-the-NE normal shearing is developed. Thisdeformation mainly affects the Late Paleozoic <strong>and</strong> TriassicVepor cover. Overlying Gemer slates <strong>and</strong> relics <strong>of</strong> theMeliata accretionary wedge contain a cleavage interpretedas manifestation <strong>of</strong> the GCF, but they are not affected byextensional reworking (Figure 4a).3.3. Cretaceous Transpressional Deformation:Trans-Gemer Shear Zone (TGSZ)[25] The transpressional deformation is marked by development<strong>of</strong> a several kilometer wide zone <strong>of</strong> NE trendingsteep cleavage along the southern boundary <strong>of</strong> the westernVepor promontory (Figure 4b). Here, in the Gemer Unit, theCarboniferous <strong>and</strong> Permian cover <strong>of</strong> the Vepor basement<strong>and</strong> Mesozoic rocks <strong>of</strong> the Meliata accretionary wedge, allpreviously developed structures, are intensively reworkedunder lower greenschist facies conditions (Figure 5d). In thestrongly attenuated Gemer Unit, relics <strong>of</strong> E-W trendingGCF fabric <strong>and</strong> the new NE trending cleavage form amap-scale sigmoidal domain surrounded by highly shearedLate Paleozoic rocks (Figure 4b). Locally, the early developedfoliation is refolded by synschistose noncylindricalfolds with steeply to subhorizontally plunging hinges,which become subparallel to horizontal stretching lineationswith increasing finite strain. These features are consistentwith progressive folding in transpressional shear zones[Fossen <strong>and</strong> Tik<strong>of</strong>f, 1998; Treagus <strong>and</strong> Treagus, 1992].[26] The boundary <strong>of</strong> the Vepor basement <strong>and</strong> LatePaleozoic cover is intruded by sheets <strong>of</strong> peraluminousgranites, which show magmatic fabric parallel to the transpressionalshear zone. In some places, a heterogeneous,steep NE trending mylonitic S-C fabric indicates sinistralshearing. Granite apophyses <strong>and</strong> dikes are folded <strong>and</strong>sheared or they crosscut the foliation <strong>of</strong> host rocks withoutinternal deformation. All these features are consistent withsyntectonic character <strong>of</strong> granite emplacement. Shallow level<strong>of</strong> magma emplacement is documented by contact metamorphism<strong>of</strong> host Late Paleozoic schists (500°C <strong>and</strong> 2 kbar[Vozárová, 1990]).[27] Toward the NE, this 5 km wide zone <strong>of</strong> steep cleavagecontinues into central part <strong>of</strong> the Gemer Unit (Figures 3a <strong>and</strong>4b). This NE trending zone <strong>of</strong> shear deformation, the Trans-Gemer Shear Zone (TGSZ), overprints all previouslydeveloped metamorphic fabrics (Figure 5e) exhibiting a20 to 25 km sinistral <strong>of</strong>fset <strong>of</strong> lithological stripes <strong>and</strong>axial zone <strong>of</strong> GCF. The displacement <strong>and</strong> intensity <strong>of</strong>deformation gradually vanishes toward the NE edge <strong>of</strong> theGemer Unit.[28] Two several kilometers wide, NE trending zones <strong>of</strong>greenschist facies transpressional deformation affectedinternal parts <strong>of</strong> the Western Vepor promontory (Figure 6b).Inside these shear zones, the Variscan as well as Mesozoicmylonitic extensional fabric are strongly refolded <strong>and</strong>transposed. The character <strong>of</strong> folding as well as numeroussense-<strong>of</strong>-shear criteria underlines the sinistral sense <strong>of</strong>movement. These zones are preferentially developed inweak lithologies such as garnetiferous micaschists orhighly anisotropic extensional mylonites. Outside <strong>of</strong> theseshear zones, a weak E-W trending crenulation cleavageaffects the flat Cretaceous extensional mylonitic foliation.The <strong>structural</strong> mapping shows anticlockwise rotation <strong>of</strong>this crenulation toward parallelism with shear zones,which confirms the sinistral displacement along shearzones.3.4. Eastern Gemer Thrust (EGT) <strong>and</strong>Associated Deformations[29] The southern part <strong>of</strong> the Gemer Unit is displacedalong sinistral TGSZ toward NE, <strong>and</strong> consequently is thrustover the eastern Vepor promontory along large-scale compressiveshear zone, the Eastern Gemer Thrust (EGT)(Figure 3b). This zone is marked by imbrication <strong>of</strong> basement<strong>and</strong> cover (both Paleozoic <strong>and</strong> Mesozoic), intenselower greenschist facies grade mylonitization <strong>of</strong> all lithologiesacross a few kilometers wide belt. An importantfeature is the incorporation <strong>of</strong> the Gemer Permian-Triassiccover as well as the Triassic-Jurassic Vepor cover intoimbricated thrust system in the form <strong>of</strong> large-scale isoclinalsynclines (Figure 3b). The forel<strong>and</strong> duplexes <strong>and</strong> myloniticFigure 3. (opposite) Idealized cross sections <strong>of</strong> central <strong>and</strong> eastern parts <strong>of</strong> the Gemer Unit. (a) Section characteristic <strong>of</strong>30 km wide <strong>and</strong> 50 km long almost E-W trending central part <strong>of</strong> the Gemer Unit. In the north, Late Carboniferous, Permian,<strong>and</strong> Triassic (Silica nappe) sequences unconformably overlie the metabasites <strong>of</strong> the Rakovec Group <strong>of</strong> the Gemer Unit.Gelnica Group builds the major part <strong>of</strong> the Gemer Unit farther to the south, where it is unconformably overlain by Permianmetas<strong>and</strong>stones. The southernmost part <strong>of</strong> the section is formed by most completely preserved sequences <strong>of</strong> Meliataaccretionary wedge <strong>and</strong> Silica nappe at the top. (b) Eastern part showing complete section through the Meliata accretionarywedge <strong>and</strong> the Silica nappe in the south <strong>and</strong> the Gemer Unit <strong>and</strong> the eastern Vepor promontory with its Mesozoic cover tothe NE. All tectonic units are trending NW-SE. The Vepor basement rocks are imbricated with their Mesozoic cover. In thehanging wall, there occurs Late Paleozoic cover <strong>of</strong> the Gemer Unit represented by Early Carboniferous slates <strong>and</strong> Permianmetaclastics <strong>and</strong> Triassic limestones. This sequence is tectonically overlain by the Rakovec <strong>and</strong> the Gelnica Groups whichare locally covered by klippens <strong>of</strong> Silica nappe. To the southwest, the Gelnica Group covered by Permian clastics <strong>and</strong> byMeliata accretionary wedge is located. The legend is same as for Figure 1.91


5 - 8 LEXA ET AL.: COLLISION IN WEST CARPATHIANSFigure 4. Idealized cross sections across the southern <strong>and</strong> eastern margin <strong>of</strong> the Vepor promontory.(a) Southwestern section <strong>of</strong> the studied area characteristic <strong>of</strong> 40 km wide <strong>and</strong> 70 km long outcrop <strong>of</strong> thewestern Vepor promontory. Southeastern margin <strong>of</strong> this crystalline complex is rimmed by low-grade LateCarboniferous metasedimentary formation (Slatviná Formation) intruded by peraluminous granites <strong>of</strong>uncertain age. Farther to the southeast, a belt <strong>of</strong> Permian greenschist facies metasediments occursfollowed by belt <strong>of</strong> Early Carboniferous greenschist facies slates (Ochtiná Formation) <strong>of</strong> variable width.The Gemer Unit consists exclusively <strong>of</strong> extremely reduced Gelnica Group. The southernmost part <strong>of</strong> thesection consists mostly <strong>of</strong> carbonates <strong>of</strong> Meliata accretionary wedge <strong>and</strong> Silica nappes. (b) NW-SEtrending contact <strong>of</strong> the western Vepor promontory with the Gemer Unit. The NE edge <strong>of</strong> the Veporbasement is rimmed by the Late Carboniferous formation (Slatviná Formation) <strong>and</strong> by Permian <strong>and</strong>Triassic greenschist facies grade metas<strong>and</strong>stones <strong>of</strong> the Vepor cover sequences. Along this border theLate Carboniferous formation is pinching out to the north so that the Permian rocks <strong>of</strong> the coversequences directly overlie the Vepor basement. Farther to the east, the Permian metasediments areoverlain by slates <strong>of</strong> the Gelnica Group. Large klippen <strong>of</strong> Permian rocks <strong>and</strong> metacarbonates <strong>of</strong> the Bôrkanappe overlie the Gelnica Formation. The legend is same as for Figure 1.Figure 5. (opposite) Field photographs <strong>of</strong> main <strong>structural</strong> features. (a) Steep cleavage (160/70) <strong>of</strong> the GCF affected by latesubhorizontal kink b<strong>and</strong>s (350/25) in axial part <strong>of</strong> GCF (Štós Saddle). (b) Weakly developed south dipping axial planecleavage (204/45) affecting early fabric (90/60) in the northern part <strong>of</strong> the Gelnica Group (road cut north <strong>of</strong> GemerskáPoloma). (c) Weak spaced cleavage (150/30) reworking Variscan metamorphic fabrics (335/60) <strong>of</strong> the Rakovec Group(Nálepkovo). (d) Steep cleavage (150/80) <strong>and</strong> vertical noncylindrical folds <strong>of</strong> TGSZ affecting early composite S 0–1 fabric<strong>of</strong> Triassic limestones (Blh valley near Ploské village). (e) Late steep cleavage <strong>of</strong> TGSZ (338/85) reworking the earlycleavage <strong>of</strong> the Gemer Cleavage Fan (178/70) in the Gelnica Group (Smolnícka Huta).92


LEXA ET AL.: COLLISION IN WEST CARPATHIANS 5 - 993


5 - 10 LEXA ET AL.: COLLISION IN WEST CARPATHIANSFigure 6. Schematic block diagrams showing tectonic evolution <strong>of</strong> the studied area. (a) Development <strong>of</strong>the Gemer Cleavage Fan due to indentation <strong>of</strong> sub-Gemer basement. (b) Development <strong>of</strong> the Trans-Gemer Shear Zone <strong>and</strong> Eastern Gemer Thrust Zone resulting from interaction between the Veporpromontories with sub-Gemer basement.foliation <strong>of</strong> the EGT system are dipping to the SW, bear anintense stretching lineation plunging to the SW <strong>and</strong> show atop-to-the-NE sense <strong>of</strong> shearing.4. Tectonic Model <strong>of</strong> Cretaceous Collision[30] The <strong>structural</strong> pattern <strong>and</strong> succession <strong>of</strong> deformationstructures described above allow modeling <strong>of</strong> the tectonicevolution <strong>of</strong> SW Carpathians <strong>and</strong> development <strong>of</strong> polyphasecleavage patterns. Our interpretation is based on mutualstrength relationships between individual units at the onset<strong>of</strong> Cretaceous convergence. Cretaceous tectonics <strong>of</strong> theWest Carpathians is characterized by north vergent collision<strong>of</strong> a southern continent with the northerly lying WestCarpathian domain (European plate). Fragments <strong>of</strong> thesouthern continental domain, now located in northern Hungary(Bükk mountains), are considered to be Neo-Proterozoicin age [Pantó et al., 1967]. Here, the absence <strong>of</strong>Variscan overprint is manifested by continuous sedimentationfrom the Early to Late Paleozoic. We suggest that thissouthern continent behaved as a rigid indenter controllingthe deformation <strong>of</strong> all northerly forel<strong>and</strong> crustal units, <strong>and</strong> inour coordinate system was actively moving toward thenorth. The mechanical contrast between the Vepor basementpromontories <strong>and</strong> the Gemer slates results from their contrastinglithologies <strong>and</strong> pre-Cretaceous evolution. Theinterval <strong>of</strong> thermal relaxation <strong>of</strong> the Vepor quartz<strong>of</strong>eldspathiccrust between the last (Late Carboniferous) importantthermal perturbation <strong>and</strong> Cretaceous collisioncorresponds to about 180 Ma. This indicates, that thegeotherm <strong>of</strong> the Vepor crust was equilibrated at the onset<strong>of</strong> Cretaceous orogeny [Cloetingh <strong>and</strong> Burov, 1996; Morgan<strong>and</strong> Ramberg, 1987]. Therefore we suggest that the VariscanVepor basement, composed <strong>of</strong> gneisses <strong>and</strong> granites, representeda mechanically strong promontory <strong>of</strong> irregular shape.In contrast, the Gemer Unit is represented mainly by lowgradeslates composed <strong>of</strong> fine-grained hydrous mineralswith rheology controlled by diffusion type <strong>of</strong> deformationmechanisms as pressure solution <strong>and</strong> diffusive mass transfer[Knipe, 1979, 1989]. For the same geotherm, as comparedwith laterally adjacent quartz<strong>of</strong>eldspathic rocks, the strength<strong>of</strong> Gemer slates was incomparably lower. Taking intoaccount these rheological assumptions, the Gemer Unitduring the Cretaceous event is considered to be the weakestdomain accommodating most <strong>of</strong> the viscous deformation.[31] In agreement with Woodcock et al. [1988] <strong>and</strong>Sintubin [1999], the cleavage patterns in deformable weakrocks reflect the geometry <strong>and</strong> direction <strong>of</strong> movement <strong>of</strong>rigid blocks. In order to model the development <strong>of</strong> superposedcleavage pattern described above, it is important todefine boundary conditions.4.1. Definition <strong>of</strong> Kinematic Frame[32] The asymmetry <strong>of</strong> the GCF can be interpreted as aresult <strong>of</strong> movement <strong>of</strong> rigid indenting block to the north <strong>and</strong>back thrusting <strong>of</strong> metasediments over its northern margin.The rigid basement does not crop out, but it can be traced indeep seismic lines 2T <strong>and</strong> G1 [Tomek, 1993; Vozár <strong>and</strong>Šantavý, 1999; Vozár et al., 1996]. The seismic pr<strong>of</strong>ilingshows that the Gemer Unit is about 5 km thick sheet-likebody resting on a basement <strong>of</strong> unknown age. This majorlithological boundary is represented by series <strong>of</strong> stronghorizontal reflectors. The most spectacular structure inseismic line G1 [Vozár <strong>and</strong> Šantavý, 1999; Vozár et al.,1996] is a highly reflective south dipping zone along whichthe horizontal base <strong>of</strong> the Gemer Unit is displaced to thenorth. This zone is interpreted as the major Sub-Gemerthrust fault responsible for northward thrusting <strong>of</strong> rigidbasement over weak sediments resulting in the development<strong>of</strong> GCF (Figure 6).[33] Important question in our model is the displacement<strong>of</strong> Vepor basement rocks with respect to the Gemer Unit. Itis well known that the whole central <strong>and</strong> southern Carpathi<strong>and</strong>omain was actively moving to the north (in recentcoordinates) as documented by Cretaceous progressiveclosure <strong>of</strong> Mesozoic Fatric basinal domain north <strong>of</strong> theVepor basement [Plašienka, 1997]. In this kinematic frameall units are shifted to the north but only differential move-94


LEXA ET AL.: COLLISION IN WEST CARPATHIANS 5 - 11Figure 7. (a) Initial geometry <strong>and</strong> boundary conditions <strong>of</strong> the <strong>numerical</strong> model. The arrow indicates thevelocity <strong>and</strong> trajectory <strong>of</strong> the indenter northern margin, with v = 0; zero Dirichlet boundary condition forvelocity; outflow, zero Neumann boundary condition for velocity. (b) Finite strain pattern developed inweak zone after 1 Myr <strong>of</strong> shortening. Distribution <strong>of</strong> strain intensity expressed in D value <strong>and</strong> orientationpattern <strong>of</strong> XY plane <strong>of</strong> finite strain ellipsoid. The foliations trajectories are shown by lines <strong>and</strong> theorientation <strong>of</strong> lineation is expressed by the color <strong>of</strong> foliation trace. White line corresponds to horizontallineation <strong>and</strong> black line to vertical one. (c) Finite strain pattern developed in weak zone after 3 Myr <strong>of</strong>shortening. Distribution <strong>of</strong> strain intensity expressed in D value <strong>and</strong> orientation pattern <strong>of</strong> XY plane <strong>of</strong>finite strain ellipsoid. (d) Distribution <strong>of</strong> finite strain symmetry expressed in K value.ments within the Vepor-Gemer system are responsible for itsinternal deformation. Important observation is that thedeformation intensity in central part <strong>of</strong> the Gemer Unitvanishes to the north. In this area, the Vepor basement ispresent (covered by Tertiary sediments) but no increase inCretaceous deformation intensity has been observed. Thismeans that this Vepor segment does not creating importantdeformation resulting from possible movement to the south.Therefore we suggest that the Vepor promontories did notmove actively to the south, <strong>and</strong> extreme deformation alongwestern <strong>and</strong> eastern Vepor promontories was imposed bygenerally northward flow <strong>of</strong> weak material. In conclusion,the only differentially moving rigid block is the northwardthrusted part <strong>of</strong> sub-Gemer basement. All other basementunits can be further considered kinematically fixed.[34] Our field studies showed that apart from GCF, anintense deformation was concentrated along southeasternedge <strong>of</strong> the western Vepor promontory producing TGSZ <strong>and</strong>also along southwestern edge <strong>of</strong> the eastern Vepor promontoryresponsible for the origin <strong>of</strong> EGT (Figure 6b). Thedevelopment <strong>of</strong> TGSZ probably results from a major changein mutual translation direction <strong>of</strong> southern sub-Gemer block<strong>and</strong> western Vepor promontory due to their oblique collisionat deeper crustal levels. Localized transpressional deformationin upper crustal levels is a typical expression <strong>of</strong> obliqueconvergence in many active regions, e.g., San Andreas faultzone [Teyssier <strong>and</strong> Tik<strong>of</strong>f, 1997], Sumatra [Tik<strong>of</strong>f <strong>and</strong>Teyssier, 1994], or Alpine fault in New Zeal<strong>and</strong> [Teyssieret al., 1995].4.2. Numerical Modeling <strong>of</strong> ProgressiveDeformation <strong>of</strong> the Gemer Unit[35] The presented <strong>numerical</strong> approach enables to modelthe deformation in a weak zone surrounded by rigid blocksor free boundaries. The approach is based on the thinviscous sheet approximation being similar to that one usedby Engl<strong>and</strong> et al. [1985] for modeling the deformation <strong>of</strong>the whole lithosphere. We assume a horizontal weak tabulardomain to have been subjected to flow with no tractions attop <strong>and</strong> bottom surface. We consider vertical gradients <strong>of</strong>the horizontal velocity to be negligible, which allows us tointegrate the equations <strong>of</strong> motion over the vertical dimension<strong>and</strong> to work with vertical averages <strong>of</strong> stress <strong>and</strong> strainrates. When linear constitutive relation between stress <strong>and</strong>strain rates is considered, the procedure leads to a system <strong>of</strong>elliptic partial differential equations for two horizontalvelocity components (see Appendix A). The system issolved by the finite element method, with the Dirichlet<strong>and</strong> Neumann boundary conditions applied to segments <strong>of</strong>the domain boundaries corresponding to the describedgeological settings (rigid indenter, free inflow or outflow95


5 - 12 LEXA ET AL.: COLLISION IN WEST CARPATHIANSFigure 8. Finite strain pattern developed in weak zone after 7 Myr <strong>of</strong> shortening. (a) Distribution <strong>of</strong>strain intensity expressed in D value <strong>and</strong> orientation pattern <strong>of</strong> XY planes <strong>and</strong> X axes <strong>of</strong> finite strainellipsoid. (b) Distribution <strong>of</strong> angles between instantaneous <strong>and</strong> finite XY planes expressing the degree <strong>of</strong>possible simple shear reactivation <strong>of</strong> existing finite anisotropy. (c) Distribution <strong>of</strong> finite strain symmetryexpressed in K value. (d) Distribution <strong>of</strong> finite topography in meters after 7 Myr <strong>of</strong> shortening in front <strong>of</strong>the indenter.<strong>of</strong> material). The vertical strain rate <strong>and</strong> velocity are relatedto the horizontal velocity field by the incompressibilityequation. Ježek et al. [2002] have described the thin sheetmodel as sensitive to the angle <strong>of</strong> collision <strong>and</strong> maybeproducing a zone dominated by lateral simple shear close tothe indenter <strong>and</strong> a zone <strong>of</strong> dominant pure shear farther awayfrom the indenting boundary. In addition, we show thatthese general features can strongly interfere with finitedimension <strong>of</strong> the modeled area <strong>and</strong> imposed boundaryconditions.4.2.1. Definition <strong>of</strong> Domain Geometry<strong>and</strong> Boundary Conditions[36] As the Vepor promontories were considered to havebeen kinematically fixed throughout the whole deformationalhistory, the zero velocity (Dirichlet) boundary conditionwas applied to their boundaries (Figure 7a), so that theirgeometry corresponds to the Vepor basement recent shape.The northern boundary <strong>of</strong> our model connecting eastern <strong>and</strong>western promontories is characterized by a zero velocitygradient (Neumann) boundary condition allowing free outflow<strong>of</strong> material to the north. This is in agreement with theabsence <strong>of</strong> a deformation gradient in this area. Similarboundary conditions are applied to the rest <strong>of</strong> examineddomain boundaries since the extension <strong>of</strong> all geologicalunits in these particular areas is unknown, being covered byTertiary sediments. The southern edge <strong>of</strong> the model representsthe actively moving rigid body <strong>of</strong> oval shape followingprescribed trajectory with constant velocity <strong>of</strong> 1 cm/yr(Figure 7a). This velocity <strong>of</strong> plate movement is deliberatelychosen to demonstrate the principal tendencies in finitestrain patterns.4.2.2. Results <strong>of</strong> Numerical Modeling[37] The main results <strong>of</strong> our modeling are presented in aseries <strong>of</strong> three time steps equal to 1, 3 <strong>and</strong> 7 Myr. Wepresent maps <strong>of</strong> strain intensities, orientations <strong>of</strong> XY planes<strong>of</strong> finite strain <strong>and</strong> orientations <strong>of</strong> X strain axes (Figures 7b,7c, <strong>and</strong> 8a). Already after 1 Myr we can observe thedevelopment <strong>of</strong> arcuate pattern <strong>of</strong> XY trajectories aroundthe rigid indenter (Figure 7b). Another feature is thedecrease in strain intensity from the south (D =0.5)tothenorth, where the strain intensity is negligible. The strainintensity increases in the western part <strong>of</strong> the model close tothe western promontory. The strain symmetry is <strong>of</strong> planestrain type in the deformed area irrespective to the strainintensity. The X axis <strong>of</strong> finite strain is vertical in the entiredomain indicating predominant pure shear deformationregime.[38] After 3 Myr we can observe several domains withcontrasting strain parameters (Figures 7c <strong>and</strong> 7d). The centraldomain in front <strong>of</strong> the indenter shows exponential decrease instrain intensity from indenter margin (D = 1) to the north(D = 0.1). The strain gradient is poorly defined toward easternpromontory <strong>and</strong> the XY planes tend to be parallel with thepromontory margin. In the west, the strain intensity is highestacross the whole width <strong>of</strong> the shortened domain <strong>and</strong> thetrajectories <strong>of</strong> XY planes are fully parallel with the westernVepor promontory margin. The strain symmetry shows dominantflattening, K close to zero, along wide zones parallel to96


LEXA ET AL.: COLLISION IN WEST CARPATHIANS 5 - 13the Vepor promontories, whereas in the central domain theincrease <strong>of</strong> K parameter up to 0.5 occurs close to the indentermargin (Figure 7d). Pure shear is dominating along the wholedeformed area except small part southeast <strong>of</strong> the indenter,where subhorizontal stretching starts to develop. In this area,we can examine angle f between XY planes <strong>of</strong> instantaneous<strong>and</strong> finite strain ellipsoid. Here, the angle f exceeds 25°, butwith progressive northward movement <strong>of</strong> the indenter, thearea <strong>of</strong> high angle f enlarges. According to Tik<strong>of</strong>f <strong>and</strong> Teyssire[1994] <strong>and</strong> Ježek et al. [2002], high angle f is critical for thedevelopment <strong>of</strong> discrete partitioning <strong>and</strong> subsequent slipalong highly inclined surfaces. In this concept, the instantaneousstretching tensor has similar significance as does astress tensor for the development <strong>of</strong> faulting <strong>and</strong> reactivation<strong>of</strong> preexisting anisotropies [Tommasi <strong>and</strong> Vauchez, 2001].Therefore the sufficiently high angle between finite stainanisotropy <strong>and</strong> instantaneous maximum stretching axis maygenerate high resolved shear stress on preexisting surfaces<strong>and</strong> reactivate these planes as strike-slip faults. In our model,the discrete partitioning starts to operate after 3 Myr leading tocontinuous change in the indenter movement direction. Weshow the pattern developed after 7 Myr, where the observedtendencies are fully developed (Figure 8b). The central part <strong>of</strong>weak domain shows exponential strain intensity decreasefrom D = 2.5 to approximately D = 0.5 in the most remotearea in the north (Figure 8a). Our highest calculated strainintensities correspond to maximum shortening <strong>of</strong> 70%,which is in agreement with values estimated for stronglydeveloped slaty cleavage [Siddans, 1972]. Therefore wesuggest that the zones <strong>of</strong> highest modeled strain intensitiescorrespond to the domains in the field with most intenselydeveloped cleavage mapped as stages 4 <strong>and</strong> 5 <strong>of</strong> Bell <strong>and</strong>Rubenach [1983]. Similarly, modeled low strain intensities <strong>of</strong>D = 0.5 (25% shortening) may be compared to heterogeneouslydeveloped cleavage stages 1 <strong>and</strong> 2 <strong>of</strong> Bell <strong>and</strong>Rubenach [1983] in the field. Modeled strain symmetriesreach values <strong>of</strong> K = 0.5, which roughly correspond tomeasured values <strong>of</strong> Németh et al. [1997]. The area betweenthe indenter <strong>and</strong> western promontory forms now a narrowchannel in which the strain intensity is significantly higher<strong>and</strong> strain symmetry more oblate. Moreover, the X axis <strong>of</strong>finite strain becomes horizontal <strong>and</strong> starts to be controlled bysimple shear deformation. At that time, following the prescribedtrajectory (Figure 7a), the indenter bulk translationvector becomes parallel to the southern margin <strong>of</strong> the westernVepor promontory (Figure 8a). This change in indentermovement strongly affects the style <strong>of</strong> deformation in theeastern part <strong>of</strong> deformable domain. In this area, an intensestrain develops up to D = 2, the foliation trajectories beingparallel to the eastern promontory margin. The strainsymmetry remains oblate with K parameter close to zero(Figure 8c).5. Discussion5.1. Validation <strong>of</strong> Numerical ModelWith Respect to Field Data[39] The <strong>numerical</strong> model <strong>of</strong> thin viscous sheet deformationgenerated by an indenter <strong>of</strong> oval shape simulates thedevelopment <strong>of</strong> deformation pattern characteristic <strong>of</strong> GCF.We note that strain intensities decrease exponentially fromthe margin <strong>of</strong> indenting block, which is in agreement withcleavage pattern observed in the field.[40] The development <strong>of</strong> discrete partitioning betweenwestern Vepor promontory <strong>and</strong> indenting block agrees wellwith the observed secondary cleavage associated withTGSZ. We suggest that the TGSZ accommodates thechange in bulk indenter translation from northward tonortheastward during the deformation. The transcurrentmovement along the western promontory is responsiblefor propagation <strong>of</strong> TGSZ into interior <strong>of</strong> the Gemer Unit.This leads to separation <strong>of</strong> GCF into northern domain withpreserved northward movement related structures, whilethe southern domain becomes frontally convergent withthe eastern promontory. This process is responsible for thedevelopment <strong>of</strong> EGT.[41] However, the presented model has serious limitations.We are not able to simulate the deformation <strong>of</strong> thoseparts <strong>of</strong> the viscous sheet, which were thrust over rigidpromontories. This particularly concerns extensional stripping<strong>of</strong> the Gemer Unit from northeastern part <strong>of</strong> thewestern Vepor promontory. The noncoaxial extensionaldeformation in this area is most likely related to the activity<strong>of</strong> TGSZ <strong>and</strong> corresponds to pulling the allochthonousGemer Unit associated with sinistral shearing along thismajor shear zone. The model is unable to demonstrate theeffects <strong>of</strong> strain localization associated with discrete partitioning.In fact, the TGSZ is passively translating southernpart <strong>of</strong> the Gemer Unit without significant internal deformation.Similarly, the development <strong>of</strong> EGT appears to be amore localized feature than is shown in our model, <strong>and</strong>leads also to passive thrusting <strong>of</strong> the Gemer Unit over theeastern Vepor promontory.[42] Despite <strong>of</strong> these limitations, the presented modelallows to predict the strain pattern in front <strong>of</strong> indenting platein an area with complex boundary conditions. Our model isintended to quantify the cleavage patterns developed due tothe movement <strong>of</strong> rigid blocks as suggested by Woodcock et al.[1988], Sintubin [1999], <strong>and</strong> others. The basis <strong>of</strong> our modelingis the assumption that the cleavage represents the XYplane<strong>of</strong> finite strain ellipsoid [Cloos, 1947; Sorby, 1853; Wood,1974]. Our model works with deformation <strong>of</strong> originallyisotropic medium <strong>and</strong> does not take into account problems<strong>of</strong> existing internal anisotropy [Cobbold et al., 1971]. However,the major advantage <strong>of</strong> our approach is the interconnection<strong>of</strong> complex kinematic frame with finite strain pattern,which was so far possible only for extremely simple boundarycondition models, e.g., simple shear, transpression, etc. Inaddition, the model explains the polyphase cleavagepatterns in terms <strong>of</strong> the complex shape <strong>of</strong> promontories <strong>and</strong>changes in movements <strong>of</strong> indenting blocks. Moreover, usingthe regional mapping <strong>of</strong> cleavage patterns, we are now able todistinguish actively moving blocks from stationary rigidpromontories.[43] We are aware that infinite numbers <strong>of</strong> boundaryconditions exist, which may generate different strain distribution<strong>and</strong> superposition <strong>of</strong> structures. Therefore we deliberatelyselected the set <strong>of</strong> boundary conditions, which satisfy97


5 - 14 LEXA ET AL.: COLLISION IN WEST CARPATHIANSthe <strong>structural</strong> evolution in the weak domain represented by theGemer Unit. Such a type <strong>of</strong> modeling could be used tovalidate chosen boundary conditions, i.e., the role <strong>of</strong> rigidpromontories for complex <strong>structural</strong> evolutions in terrainswith polyphase deformation.5.2. Timescales <strong>of</strong> the Proposed Model[44] The timescale <strong>of</strong> the model is controlled by velocity <strong>of</strong>the indenting block. We have chosen the arbitrary velocity <strong>of</strong>1 cm/yr for the sake <strong>of</strong> simplicity. However, in the case<strong>of</strong> Vepor <strong>and</strong> Gemer Units, we can define the velocity <strong>of</strong>movement <strong>of</strong> our kinematically fixed system. On the basis <strong>of</strong>the knowledge <strong>of</strong> approximate initial width <strong>and</strong> stratigraphicrecord <strong>of</strong> the Mesozoic (Fatric) basin in front <strong>of</strong> theVepor basement [Plašienka, 1997], the rate <strong>of</strong> shortening isestimated to be about 1 cm/yr, <strong>and</strong> the duration <strong>of</strong> theshortening process is estimated at about 20 Myr. Plašienka[1997] also demonstrated that the original frontal closure <strong>of</strong>the Fatric Basin passed to transpressive movements after20 Myr. This means that the differential movement <strong>of</strong> rigidindenter, which moves together with the whole kinematicsystem, has to generate a defined finite strain at the sameperiod <strong>of</strong> time. Moreover, the initiation <strong>of</strong> TGSZ activity maycorrespond to a transition from frontal to transpressionalmovements recorded in the northern edge <strong>of</strong> the wholekinematic system. Once this rough timescale is established,then the absolute velocity <strong>of</strong> our indenter should be four timesslower than suggested in the model to generate the observedstrain pattern.5.3. Development <strong>of</strong> Topography, Exhumation, <strong>and</strong>Asymmetry <strong>of</strong> GCF[45] The model allows estimation <strong>of</strong> average verticalstrains <strong>and</strong>, because <strong>of</strong> a fixed lower boundary condition,also the vertical elevation. We can expect that the surfaceelevation represent local topography generated by shortening<strong>of</strong> the viscous sheet. The lateral distribution <strong>of</strong> topographyfollows the exponential distribution <strong>of</strong> finite strain inareas <strong>of</strong> pure shear-dominated deformation. Figure 8dshows the distribution <strong>of</strong> topography in front <strong>of</strong> an indentingblock after 7 Myr <strong>of</strong> shortening. It is to be noted that thedomain <strong>of</strong> highest topography follows the axial zone <strong>of</strong> theGCF, where the degree <strong>of</strong> metamorphism associated withthe development <strong>of</strong> cleavage is most important.[46] Although our model predicts vertical cleavage in theentire domain, we observe that the cleavage forms a positivefan-like structure. We interpret this pattern as a result <strong>of</strong>different amount <strong>of</strong> vertical shortening due to differentgravitational potential across the GCF. This mechanism ismanifested by the development <strong>of</strong> late kink b<strong>and</strong>s with kinkplanes perpendicular or oblique to strongly developed verticalcleavage.Appendix A[47] The equations governing the deformation <strong>of</strong> a thinviscous sheet were published by Engl<strong>and</strong> <strong>and</strong> McKenzie[1982]. We provide here the derivation <strong>of</strong> equations for thesimplest case <strong>of</strong> Newtonian rheology as they have beenused for our modeling. The model assumes a relatively thinviscous plate with no tractions at top <strong>and</strong> bottom surface<strong>and</strong> negligible vertical gradients <strong>of</strong> horizontal velocitycomponents. Creep equation reads@t ij@x j¼ @p@x iðA1Þwhere T is the deviatoric stress tensor <strong>and</strong> p is the pressure,<strong>and</strong> repeated index means summation. Assuming a linearconstitutive relationt ij ¼ 2h_e ijequation (1) can be written as2h @_e ij@x j¼ @p@x i[48] The strain rate tensor <strong>of</strong> the form01_e 11 _e 12 0_E ¼_e 21 _e 22 0BC@A0 0 _e 33ðA2ÞðA3ÞðA4Þis assumed. Then the equation containing the vertical strainrate _e 33 reduces to2h @_e 33@x 3¼ @p@x 3ðA5Þ[49] Integrating the equation over the vertical dimensionx 3 , we obtain2h_e 33 ¼ p þ fðx 1 ; x 2 Þ ðA6Þwhere the upper strike means vertical average. Because <strong>of</strong>the model assumptions we can put f (x 1 ,x 2 ) = 0 everywhere.We substitute from equation (A6) in the first equation (A3)integrated over the vertical dimension <strong>and</strong> obtain2h @ _e 1j@x j¼ @p@x 1¼ 2h @ _e 33@x 1<strong>and</strong> similarly for the second equation. The resulting equationscan be written as@_e ij@x j@_e 33@x i¼ 0 i ¼ 1; 2; j ¼ 1; 2 ðA7Þwhere we omit the signs <strong>of</strong> vertical averaging. Massconservation for incompressible flow requires that_e 33 ¼ ð_e 11 þ _e 22 Þ ðA8Þ98


LEXA ET AL.: COLLISION IN WEST CARPATHIANS 5 - 15Thus2 @_e 11þ @_e 12þ @_e 22¼ 0@x 1 @x 2 @x 1@_e 11þ @_e 21þ 2 @_e 22¼ 0@x 2 @x 1 @x 2ðA9Þ[52] Although we compute the horizontal velocity field,the vertical strain rate can be estimated by equation (8) thatallows us to assess at every step the tensors <strong>of</strong> instantaneous<strong>and</strong> finite deformation <strong>and</strong> related parameters <strong>of</strong> deformation,such as intensity[50] We replace the strain rate tensor by horizontalcomponents <strong>of</strong> velocity u 1 , u 2_e ij ¼ 1 @u iþ @u jðA10Þ2 @x j @x i<strong>and</strong> finally obtain a system <strong>of</strong> elliptic partial differentialequationssymmetryqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiD ¼ R xy 1 2þ Ryz 1 2ðA12ÞK ¼ R xy 1 R yz 1 ðA13Þ4 @2 u 1@x 2 1@ 2 u 2@x 2 1þ @2 u 1@x 2 þ 3 @2 u 2¼ 02@x 1 @x 2þ 4 @2 u 2@x 2 þ 3 @2 u 1¼ 02@x 1 @x 2ðA11Þ<strong>and</strong> orientations <strong>of</strong> lineations <strong>and</strong> foliations. Because <strong>of</strong> theassumed form <strong>of</strong> the strain rate tensor, equation (A4), themodel can produce only horizontal or vertical lineations <strong>and</strong>foliations <strong>and</strong> can be regarded as generalized (locallyvariable) transpression.[51] The system can be solved by the finite elementmethod. Dirichlet <strong>and</strong> Neumann boundary conditions maybe applied to segments <strong>of</strong> the domain boundaries so thatthey correspond to geological settings (rigid indenter, freeinflow or outflow <strong>of</strong> material).[53] Acknowledgments. We are grateful to the Geological Survey <strong>of</strong>the Slovak Republic for significant financial support during the initialstages <strong>of</strong> our research. This work has been also supported by the CharlesUniversity Agency grant 216/1999/B-GEO. The salaries <strong>of</strong> K. Schulmann<strong>and</strong> O. Lexa were covered by Ministry <strong>of</strong> Education grant 24313005.ReferencesAllem<strong>and</strong>, P., <strong>and</strong> J. M. Lardeaux, Strain partitioning<strong>and</strong> metamorphism in a deformable orogenicwedge: Application to the Alpine belt, Tectonophysics,280, 157 – 169, 1997.Andrusov, D., Subtatranské príkrovy v Západných Karpatoch,Carpatica, 1, 3 – 50, 1936.Andrusov, D., Geológia Československých Karpát Zv.1, Vydavatel’stvo SAV, Bratislava, Slovakia, 1958.Bell, T. H., <strong>and</strong> M. J. Rubenach, Sequential porphyroblastgrowth <strong>and</strong> crenulation cleavage developmentduring progressive deformation, Tectonophysics,92, 171 – 194, 1983.Biely, A., <strong>and</strong> J. Bystrický, Mesozoic <strong>of</strong> the inner westCarpathians <strong>and</strong> the klippen belt, in Guide toExcursion, 44 pp., Int. Geol. Congr., Prague, 1967.Biely, A., J. Bystrický, <strong>and</strong> O. Fusán, De l’appartenancedes nappes des Karpates occidentales internesTranslated Title: Relations <strong>of</strong> the nappes <strong>of</strong> theinner western Carpathians, paper presented at Int.Geol. Congr., Prague, 1968.Bystrický, J., Uebersicht der Stratigraphie und Entwicklungder Trias in den Westkarpaten (Triassicstratigraphy <strong>of</strong> the eastern Carpathians), Geol. Sb.Slov. Akad. Vied, 18, 257 – 266, 1967.Cambel, B., J. Král’, <strong>and</strong> J. Burchart, Isotopic Geochronology<strong>of</strong> the Western Carpathians CrystallineComplex With Catalogue <strong>of</strong> Data, pp. 183, Veda,Bratislava, Slovakia, 1990.Cloetingh, S., <strong>and</strong> E. B. Burov, Thermomechanicalstructure <strong>of</strong> European continental lithosphere: Constraintsfrom rheological pr<strong>of</strong>iles <strong>and</strong> EET estimates,Geophys. J. Int., 124, 695 – 723, 1996.Cloos, E., Oolite deformation in the South Mountainfold, Maryl<strong>and</strong>, Geol. Soc. Am. Bull., 58, 843 – 918,1947.Cobbold, P. R., J. W. Cosgrove, <strong>and</strong> J. M. Summers,Development <strong>of</strong> internal structures in deformedanisotropic rocks, Tectonophysics, 12, 23 – 53, 1971.Dallmeyer, R. D., F. Neubauer, R. H<strong>and</strong>ler, H. Fritz,W. Mueller, D. Pana, <strong>and</strong> M. Putiš, Tectonothermalevolution <strong>of</strong> the internal Alps <strong>and</strong> Carpathians: Evidencefrom 40 Ar/ 39 Ar mineral <strong>and</strong> whole-rock data,Eclogae Geol. Helv., 89, 203 – 227, 1996.Engl<strong>and</strong>, P., <strong>and</strong> D. McKenzie, A thin viscous sheetmodel for continental deformation, Geophys. J. R.Astron. Soc., 70, 295 – 321, 1982. (Correction to ‘‘Athin viscous sheet model for continental deformation’’,Geophys. J. R. Astron. Soc., 73, 523 – 532,1983).Engl<strong>and</strong>, P., G. Houseman, <strong>and</strong> L. Sonder, Lengthscales for continental deformation in convergent,divergent, <strong>and</strong> strike-slip environments: Analytical<strong>and</strong> approximate solutions for a thin viscous sheetmodel, J. Geophys. Res., 90, 3551 – 3557, 1985.Faryad, S. W., Gneiss-amphibolite complex <strong>of</strong> Gemericum(in Slovak), Miner. Slovaca, 22, 303 – 318,1990.Faryad, S. W., Thermal overprint in the early Paleozoicsequences <strong>of</strong> Gemericum (Western Carpathians), inSpecial Volume to the Problems <strong>of</strong> the PaleozoicGeodynamic Domains: Western Carpathians, EasternAlps, Dinarides, edited by J. Vozár, pp. 49 – 56,Dionýz Štúr Inst. Geol., Bratislava, Slovakia, 1992.Faryad, S. W., Mineralogy <strong>of</strong> Mn-rich rocks fromgreenschist facies sequences <strong>of</strong> the Gemericum,West Carpathians, Slovakia, Neues Jarhb. Mineral.Monatsh., 10, 464 – 480, 1994.Faryad, S. W., Phase petrology <strong>and</strong> P-T conditions <strong>of</strong>mafic blueschists from the Meliata unit, WesternCarpathians, Slovakia, J. Metamorph. Geol., 13,701 – 714, 1995.Faryad, S. W., <strong>and</strong> H. J. Bernhardt, Taramite-bearingmetabasites from Rakovec (Gemeric Unit, the WesternCarpathians), Geol. Carpath., 47, 349 – 357,1996.Faryad, S. W., <strong>and</strong> F. Henjes-Kunst, Petrological <strong>and</strong>K-Ar <strong>and</strong> Ar-40-Ar-39 age constraints for thetectonothermal evolution <strong>of</strong> the high-pressureMeliata unit, Western Carpathians (Slovakia),Tectonophysics, 280, 141 – 156, 1997.Finger, F., <strong>and</strong> I. Broska, The Gemeric S-type granitesin southeastern Slovakia: Late Palaeozoic or Alpineintrusions? Evidence from electron-microprobe dating<strong>of</strong> monazite, Schweiz. Mineral. Petrogr. Mitt.,79, 439 – 443, 1999.Fossen, H., <strong>and</strong> B. Tik<strong>of</strong>f, Extended models <strong>of</strong> transpression<strong>and</strong> transtension, <strong>and</strong> application to tectonicsetting, in Continental Transpressional <strong>and</strong> TranstensionalTectonics, edited by R. E. Holdsworth,R. A. Strachan <strong>and</strong> J. F. Dewey, Geol. Soc. Spec.Publ., 135, 15 – 33, 1998.Genser, J., J. D. Van Wees, S. Cloetingh, <strong>and</strong> F. Neubauer,Eastern Alpine tectono-metamorphic evolution:Constraints from two-dimensional P-T-t modeling,Tectonics, 15, 584 – 604, 1996.Hók, J., P. Kováč, <strong>and</strong> M. Rakús, Structural investigations<strong>of</strong> the Inner Carpathians: Results <strong>and</strong> interpretation,Miner. Slovaca, 27, 231 – 235, 1995.Hovorka, D., <strong>and</strong> S. Méres, Relicts <strong>of</strong> high-temperaturemetamorphic rocks in the Tatro-Veporicum crystalline<strong>of</strong> the Western Carpathians (in Slovak), Miner.Slovaca, 21, 193 – 201, 1989.Jacko,S.,S.P.Korikovskij,<strong>and</strong>V.A.Boronichin,Equilibrium assemblages <strong>of</strong> gneisses <strong>and</strong> amphibolites<strong>of</strong> Bujanová complex (Čierna Hora), easternSlovakia, Miner. Slovaca, 22, 231 – 239, 1990.Janák, M., P. J. O’Brien, V. Hurai, <strong>and</strong> C. Reutel,Metamorphic evolution <strong>and</strong> fluid composition <strong>of</strong>garnet-clinopyroxene amphibolites from the TatraMountains, Western Carpathians, Lithos, 39, 57–79, 1996.Janák,M.,M.Cosca,F.Finger,D.Plašienka,B.Koroknai,B. Lupták, <strong>and</strong> P. Horváth, Alpine (Cretaceous)metamorphism in the Western Carpathians: P-T-tpaths <strong>and</strong> exhumation <strong>of</strong> the Veporic core complex,Geol. Paläontol. Mitt. Innsbruck, 25, 115–118,2001a.Janák, M., D. Plašienka, M. Frey, M. Cosca, S. T.Schmidt, B. Lupták, <strong>and</strong> S. Méres, Cretaceous evolution<strong>of</strong> a metamorphic core complex, the Veporic99


5 - 16 LEXA ET AL.: COLLISION IN WEST CARPATHIANSunit, Western Carpathians (Slovakia): P-T conditions<strong>and</strong> in situ Ar-40/Ar-39 UV laser probe dating<strong>of</strong> metapelites, J. Metamorph. Geol., 19, 197 – 216,2001b.Ježek, J., K. Schulmann, <strong>and</strong> A. B. Thompson, Strainpartitioning, strike slip faulting <strong>and</strong> the development<strong>of</strong> strain parameters in front <strong>of</strong> an obpliquelyconvergent indenter, in Continental Collision <strong>and</strong>Tectonosedimentary Evolution <strong>of</strong> Forel<strong>and</strong>s, EGSStephan Mueller Spec. Publ. Ser., vol.1,editedby G. Bertotti, K. Schulmann, <strong>and</strong> S. Cloetingh,pp. 145 – 165, Eur. Geophys. Soc., London, 2002.Kantor, J., Ar 40 /K 40 method <strong>of</strong> absolute age determination<strong>of</strong> rocks <strong>and</strong> their application to theBetliar granite, Geol. Práce Zprávy., 11, 188–200, 1957.Knipe, R. J., Chemical changes during slaty cleavagedevelopment, Bull. Mineral., 102, 206 – 209, 1979.Knipe, R. J., Deformation mechanisms: Recognitionfrom natural tectonites, J. Struct. Geol., 11, 127 –146, 1989.Korikovskij, S. P., E. Krist, <strong>and</strong> V. A. Boronikhin,Staurolite-chloritoid schists from Klenovec region:Prograde metamorphism <strong>of</strong> high-alumina rocks <strong>of</strong>the Kohút zone: Veporides, Geol. Zbor. Geol. Carpath.,39, 187 – 200, 1989.Korikovskij, S. P., S. Jacko, <strong>and</strong> V. A. Boronichin,Facial conditions <strong>of</strong> Variscan prograde metamorphismin the Lodina complex <strong>of</strong> Èierna hora crystalline,eastern Slovakia, Miner. Slovaca, 22, 225–230, 1990.Kováčik, M., J. Král’, <strong>and</strong> H. Maluski, Metamorphicrocks in the southern Veporicum: Their Alpinemetamorphism <strong>and</strong> thermochronologic evolution,Miner. Slovaca, 28, 185 – 202, 1996.Maluski, H., P. Rajlich, <strong>and</strong> P. Matte, 40 Ar- 39 Ar dating<strong>of</strong> the Inner Carpathians Variscan basement <strong>and</strong>Alpine mylonitic overprinting, Tectonophysics,223, 313 – 337, 1993.Máška, M., Poznámky k předtercierní metalogenesiZápadních Karpat, zvláště Spišsko-gemerskéhorudohoří, Geol. Práce, 46, 96 – 106, 1957.Matějka, A., <strong>and</strong> D. Andrusov, Overview <strong>of</strong> the Geology<strong>of</strong> West Carpathians in the Central Slovakia<strong>and</strong> Adjacent Regions, 163 pp., Státní Geol. Úst.Praha, Prague, 1931.Méres, S., <strong>and</strong> D. Hovorka, Alpine metamorphicrecrystallization <strong>of</strong> the pre-Carboniferous metapelites<strong>of</strong> the Kohut crystalline complex (the WesternCarpathians), Miner. Slovaca, 23, 435 – 442, 1991.Mock, R., <strong>and</strong> P. Reichwalder, The blueschist belt <strong>of</strong>the Slovak karst, Slovakia, Terra Abstr., 4, 46,1992.Morgan, P., <strong>and</strong> I. B. Ramberg, Physical changes in thelithosphere associated with thermal relaxation afterrifting, Tectonophysics, 143, 1 – 11, 1987.Németh, Z., L. Gazdačko, D. Návesňák, <strong>and</strong> J. Kobulský,Polyphase tectonic evolution <strong>of</strong> the Gemericum(the Western Carpathians) outlined by review <strong>of</strong><strong>structural</strong> <strong>and</strong> deformational data, in GeologicalEvolution <strong>of</strong> the Western Carpathians, edited byP. Grecula, D. Hovorka, <strong>and</strong> M. Putiš, pp. 215 –224, Geocomplex, Bratislava, Slovakia, 1997.Pantó, G.,A.Kovács, K. Balogh, <strong>and</strong> Z. Sámsoni,Rb-Sr check <strong>of</strong> Assyntian <strong>and</strong> Caledonian metamorphism<strong>and</strong> igneous activity in NE Hungary,Acta Geol. Acad. Sci. Hung., 11, 279 – 282, 1967.Passchier, C. W., <strong>and</strong> R. A. J. Trouw, Microtectonics,289 pp., Springer-Verlag, New York, 1996.Pl<strong>and</strong>erová, E., <strong>and</strong> A. Vozárová, Upper Carboniferousin southern Veporides, Geol. Práce Zprávy., 70,129 – 141, 1978.Plašienka, D., Mosozoic tectonic evolution <strong>of</strong> the epi-Variscan continental crust <strong>of</strong> the Western Carpathians:A tentative model, Miner. Slovaca, 23,447 – 457, 1991.Plašienka, D., Cleavages <strong>and</strong> folds in changing tectonicregimes: The Vel’ký Bok Mesozoic cover unit <strong>of</strong>the Veporicum (Nízke Tatry Mts., central WesternCarpathians), Slovak Geol. Mag., 2–95, 97 – 113,1995.Plašienka, D., Cretaceous tectonochronology <strong>of</strong> thecentral Western Carpathians, Slovakia, Geol. Carpath.,48, 99 – 111, 1997.Plašienka, D., P. Grecula, M. Putiš, M. Kováč, <strong>and</strong>D. Hovorka, Evolution <strong>and</strong> structure <strong>of</strong> the WesternCarpathians: An overview, in Geological Evolution<strong>of</strong> the Western Carpathians, edited by P. Grecula,D. Hovorka, <strong>and</strong> M. Putiš, pp. 1 – 24, Geocomplex,Bratislava, Slovakia, 1997.Poller, U., P. Uher, I. Broska, D. Plašienka, <strong>and</strong>M. Janák, First Permian-Early Triassic zircon agesfor tin-bearing granites from the Gemeric unit(Western Carpathians, Slovakia): Connection tothe post-collisional extension <strong>of</strong> the Variscanorogen <strong>and</strong> S-type granite magmatism, Terra Nova,14, 41 – 48, 2002.Schmid, S. M., O. A. Pfiffner, N. Froitzheim,G. Schoenborn, <strong>and</strong> E. Kissling, Geophysicalgeologicaltransect <strong>and</strong> tectonic evolution <strong>of</strong> theSwiss-Italian Alps, Tectonics, 15, 1036 – 1064,1996.Siddans, A. W. B., Slaty cleavage: A review <strong>of</strong> researchsince 1815, Earth Sci. Rev., 8, 205 – 232, 1972.Siman, P., V. Johan, P. Ledru, V. Bezák, <strong>and</strong> J. Madarás,Deformation <strong>and</strong> p-T conditions estimated in‘‘layered migmatites’’ from southern part <strong>of</strong> Veporicumcrystalline basement (Western Carpathians,Slovakia), Slovak Geol. Mag., 3 – 4/96, 209 – 213,1996.Sintubin, M., Arcuate fold <strong>and</strong> cleavage patterns in thesoutheastern part <strong>of</strong> the Anglo-Brabant Fold Belt(Belgium): Tectonic implications, Tectonophysics,309, 81 – 97, 1999.Snopko, L., <strong>and</strong> P. Reichwalder, Tektonická mapaSpišsko-Gemerského rudohoria, Geol. Ústav DionýzaŠtúra, a Bratislava, Slovakia, 1970.Sorby, H. C., On the origin <strong>of</strong> slaty cleavage, NewPhilos. J., 55, 137 – 148, 1853.Soták, J., A. Vozárová, <strong>and</strong> J. Ivanička, New micr<strong>of</strong>ossilsfrom the early Paleozoic formations <strong>of</strong> theGemericum (Foraminiferida), Geol. Carpath., 50,72 – 74, 1999.Šucha, V., <strong>and</strong> D. D. Eberl, Burial metamorphism <strong>of</strong> thePermian sediments from the Western Carpathians,Miner. Slovaca, 24/5, 399 – 405, 1992.Teyssier, C., <strong>and</strong> B. Tik<strong>of</strong>f, Study <strong>of</strong> strike-slip partitioningin California resolves San Andreas discrepancy<strong>and</strong> constrains lithospheric structure, Geol.Soc. Am. Abstr. Programs, 29, 347, 1997.Teyssier, C., K. Kleinspehn, <strong>and</strong> J. Pershing, Analysis<strong>of</strong> fault populations in western Spitsbergen: Implicationsfor deformation partitioning along transformmargins, Geol. Soc. Am. Bull., 107, 68 – 82, 1995.Tik<strong>of</strong>f, B., <strong>and</strong> C. Teyssier, Strain <strong>modelling</strong> <strong>of</strong> displacement-fieldpartitioning in transpressional orogens,J. Struct. Geol., 16, 1575 – 1588, 1994.Tomek, C., Deep crustal structure beneath the central<strong>and</strong> inner West Carpathians, Tectonophysics, 226,417 – 431, 1993.Tommasi, A., <strong>and</strong> A. Vauchez, Continental rifting parallelto ancient collisional belts: An effect <strong>of</strong> themechanical anisotropy <strong>of</strong> the lithospheric mantle,Earth Planet. Sci. Lett., 185, 199 – 210, 2001.Treagus, S. H., <strong>and</strong> J. E. Treagus, Transected folds <strong>and</strong>transpression; how are they associated?, J. Struct.Geol., 14, 361 – 367, 1992.Trumpy, R., The timing <strong>of</strong> orogenic events in the centralAlps, in Gravity <strong>and</strong> Tectonics, pp. 229 – 251,John Wiley, Hobobken, N. J., 1973.Vozár, J., Paleogeografický vývoj Západných Karpát(Paleogeographical Evolution <strong>of</strong> the West Carpathians),346 pp., Geol. Ústav Dionýza Štúra, Bratislava,Slovakia, 1978.Vozár, J., <strong>and</strong> J. Šantavý, Atlas <strong>of</strong> Deep Reflection SeismicPr<strong>of</strong>iles <strong>of</strong> the Western Carpathians <strong>and</strong> TheirInterpretation, GSSR, Bratislava, Slovakia, 1999.Vozár, J., Č. Tomek, A. Vozárová, J. Mello, <strong>and</strong>J. Ivanička, Seismic section G-1, Geol. Práce,101, 32 – 34, 1996.Vozárová, A., Development <strong>of</strong> metamorphism in theGemeric/Veporic contact zone (Western Carpathians),Geol. Zbor. Geol. Carpath., 41, 475–502, 1990.Vozárová, A., Variscan metamorphism <strong>and</strong> crustal evolutionin the Gemericum (in Slovak), Záp. Karp.Sér. Miner. Petrol. Geochém. Metalogen., 16, 55–117, 1993.Vozárová, A., J. Soták, <strong>and</strong> J. Ivanička, Cambro-Ordovicianfossils (conodontes, foraminifers, chitinousshields) from the methamorphic series <strong>of</strong> theGemericum (western Carpathians), in Tenth Meeting<strong>of</strong> European Union <strong>of</strong> Geosciences, vol. 4, Abstracts,p. 266, Cambridge Univ. Press, New York, 1999.Wood, D. S., Current views <strong>of</strong> the development <strong>of</strong> slatycleavage, Annu. Rev. Earth Planet. Sci., 2, 369 –401, 1974.Woodcock, N. H., M. A. Awan, T. E. Johnson, A. H.Mackie, <strong>and</strong> R. D. A. Smith, Acadian tectonics <strong>of</strong>Wales during Avalonia Laurentia convergence, Tectonics,7, 483 – 495, 1988.J. Ježek, Institute <strong>of</strong> Applied Mathematics <strong>and</strong>Computer Science, Albertov 6, 128 43 Prague 2, CzechRepublic.O. Lexa <strong>and</strong> K. Schulmann, Institute <strong>of</strong> Petrology<strong>and</strong> Structural Geology, Albertov 6, 128 43 Prague 2,Czech Republic. (lexa@natur.cuni.cz)100


Journal <strong>of</strong> Structural Geology 26 (2004) 155–161www.elsevier.com/locate/jsgApparent shear-b<strong>and</strong> geometry resulting fromoblique fold sectionsOndrej Lexa a, *, John Cosgrove b , Karel Schulmann aa Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University, Prague, Czech Republicb Department <strong>of</strong> Earth Sciences <strong>and</strong> Engineering, Royal School <strong>of</strong> Mines, Imperial College, London SW7 2BP, UKReceived 1 May 2002; received in revised form 1 February 2003; accepted 4 April 2003AbstractSmall-scale shear zones inclined at intermediate angles to an earlier anisotropy are <strong>of</strong>ten observed in deformed rocks. They aretraditionally described as shear-b<strong>and</strong>s, C-b<strong>and</strong>s, extensional crenulation cleavage or normal kink-b<strong>and</strong>s formed as a result <strong>of</strong> extension alongthe anisotropy. Their asymmetries are widely used to describe the large-scale kinematics <strong>of</strong> deformation <strong>and</strong> the deformational history <strong>of</strong> agiven area. We demonstrate that when various three-dimensional fold structures are observed on two-dimensional outcrop surfaces or in thinsection, they can appear geometrically identical. We have developed a simple technique that allows the geometrical evaluation <strong>of</strong> any sectionacross a cylindrical fold <strong>of</strong> arbitrary geometry. The ranges <strong>of</strong> planar sections on which a fold exhibits shear-b<strong>and</strong> like geometry are presentedon a stereographic projection in order to simplify the determination <strong>of</strong> critical orientations. We demonstrate that for any fold geometry, thereare two distinct groups <strong>of</strong> sections showing shear-b<strong>and</strong> like geometry with opposite ‘senses <strong>of</strong> shear’ criteria systematically arranged aroundthe axial plane <strong>and</strong> which are inclined at a high angle to the major anisotropy. We provide a field example from Western Carpathians, wherekinematic analysis, mainly based on apparent extensional shear-b<strong>and</strong>s, led to overemphasis <strong>of</strong> the role <strong>of</strong> post-orogenic extension on the final<strong>structural</strong> pattern <strong>of</strong> the belt.q 2003 Elsevier Ltd. All rights reserved.Keywords: Shear-b<strong>and</strong> geometry; Oblique fold sections; Small-scale shear zones1. IntroductionSmall-scale shear zones inclined at intermediate angles to aprevious anisotropy are commonly observed in deformedrocks. They are traditionally presented as shear-b<strong>and</strong>s (White,1979), C 0 -b<strong>and</strong>s (Ponce <strong>and</strong> Choukroune, 1980), extensionalcrenulation cleavage (Platt, 1979, 1984; Platt <strong>and</strong> Vissers,1980), asymmetric boudinage, asymmetric folds or normalkink-b<strong>and</strong>s (Dewey, 1965; Cobbold et al., 1971; Cosgrove,1976) formed as a result <strong>of</strong> extension along the olderanisotropy or shortening normal to the anisotropy. Their‘sense <strong>of</strong> shear’ <strong>and</strong> geometrical relations are widely used todescribe the large-scale kinematics <strong>of</strong> deformation (Berthéet al., 1979; Simpson <strong>and</strong> Schmid, 1983; Lister <strong>and</strong> Snoke,1984) or the tectonic settings <strong>of</strong> the deformational history(Platt <strong>and</strong> Vissers, 1980; Behrmann, 1987).Shear b<strong>and</strong>s may resemble the compressional crenulation* Corresponding author. Tel.: þ420-22195-1531; fax: þ420-22195-1524.E-mail address: lexa@natur.cuni.cz (O. Lexa).0191-8141/03/$ - see front matter q 2003 Elsevier Ltd. All rights reserved.doi:10.1016/S0191-8141(03)00072-5cleavage but develop by extension <strong>of</strong> the older foliationrather than by shortening (Passchier <strong>and</strong> Trouw, 1996). Thisled some authors to use the terms compressional (CCC) <strong>and</strong>extensional (SBC) crenulation cleavages (Platt <strong>and</strong> Vissers,1980). Passchier <strong>and</strong> Trouw (1996) presented a summary <strong>of</strong>differences between these two contrasting structures. Theirmain argument for distinction between both kinds <strong>of</strong>structures is the angle <strong>of</strong> CCC with the older foliation,which generally ranges between 45 <strong>and</strong> 908, while for SBCthe angle to earlier foliation is less than 458. However, theangular distinction between CCC <strong>and</strong> SBC is not alwaysvalid. The compressional crenulation cleavage changes thegeometry in the pr<strong>of</strong>ile section towards the hinge direction<strong>of</strong> the folded domain, so that the internal rotation becomesless than 458 <strong>and</strong> may be easily misinterpreted as an SBC(Price <strong>and</strong> Cosgrove, 1994, p. 263, Fig. 10.50).From a kinematic point <strong>of</strong> view, CCC develops at a highangle to bulk shortening while SBC represents a single shearplane at small angle to the foliation (Passchier <strong>and</strong> Trouw,1996). In order to interpret the kinematic significance <strong>of</strong>both kinds <strong>of</strong> structures, they have to be observed in plane,101


156O. Lexa et al. / Journal <strong>of</strong> Structural Geology 26 (2004) 155–161Fig. 1. Shear b<strong>and</strong> geometry <strong>and</strong> definition <strong>of</strong> angles a, b <strong>and</strong> g according toPlatt <strong>and</strong> Vissers (1980).perpendicular to the intersection <strong>of</strong> CCC <strong>and</strong> SBC with theolder foliation.With the advent <strong>of</strong> modern kinematic analysis in <strong>structural</strong>geology in the early 1980’s, the XZ plane <strong>of</strong> finite strainellipsoid became extremely important. This plane is traditionallydefined as a plane parallel to the stretching lineation <strong>and</strong>perpendicular to the foliation. However, the stretchinglineation in phyllites or phyllonites can be difficult to define,<strong>and</strong> it can be easily confused with corrugations or intersectionlineation. Moreover, the presence <strong>of</strong> well-defined shear b<strong>and</strong>son a rock surface in the field is in many cases considered asatisfactory indicator to consider this surface as an XZ plane.Because, shear b<strong>and</strong>s are such noticeable structures <strong>and</strong> theirkinematic significance is straightforward, these structureshave been widely used as first order kinematic indicators inmany orogenic belts (Behrmann, 1987).This paper aims to demonstrate that: (1) distinguishingbetween CCC <strong>and</strong> SBC is not always an easy task, (2) unlessthis is fully appreciated there is a great danger <strong>of</strong>misinterpretation when the shear b<strong>and</strong>s are used as kinematicindicators, <strong>and</strong> (3) the CCC <strong>and</strong> SBC can appear identicalwhen seen on flat outcrop surfaces or in thin sections.2. Geometrical characteristics <strong>of</strong> oblique sections <strong>of</strong> foldsFolding or flow partitioning are commonly presented intwo dimensions while geological structures are threedimensional(3D) features. In order to justify the use <strong>of</strong>the two-dimensional (2D) <strong>analyses</strong> to investigate a threedimensionalproblem, certain assumptions are made.Because the displacement fields predicted by the 2D-theoryare limited to a plane (usually one <strong>of</strong> the principal planes <strong>of</strong>the strain ellipsoid), it is assumed that displacement isidentical in any parallel plane <strong>and</strong> therefore that theresulting fold structures are cylindrical. As a result, theaxes <strong>of</strong> the folds are perpendicular to the plane <strong>of</strong>the displacement field. Based on these assumptions wewill consider 2D sections normal to the fold axes to be ‘true’sections <strong>and</strong> all other sections ‘apparent’.Geometrically, shear-b<strong>and</strong>s can be characterized by threeangles a, b <strong>and</strong> g (Fig. 1), corresponding to the relativeangles <strong>of</strong> the limbs, hinge trace <strong>and</strong> enveloping surface(Platt <strong>and</strong> Vissers, 1980). In order for the fold or flexurepr<strong>of</strong>iles to have shear-b<strong>and</strong> geometry, a should have a valuebetween 10 <strong>and</strong> 508, while b <strong>and</strong> g should range between 10<strong>and</strong> 258, so that the interlimb angle exceeds 1308. Platt <strong>and</strong>Vissers (1980) show that these angles are mean values fromobservations in the field, for example from the Beticmovement zone (Behrmann, 1987). We will show that for agiven fold pr<strong>of</strong>ile in the ‘true’ section (i.e. a section normalto the fold axis; Fig. 2) there is a specific range <strong>of</strong> ‘apparent’sections (i.e. a section oblique to the fold axis; Fig. 2) inwhich the fold pr<strong>of</strong>iles satisfy the geometrical criteria <strong>of</strong>shear-b<strong>and</strong>s.Apparent sections through three-dimensional cylindrical,coaxial folds can be treated as fold axis parallel projections<strong>of</strong> ‘true’ fold section onto the section plane. When ax þby þ cz ¼ 0 is the equation <strong>of</strong> a section plane <strong>and</strong> ( p, q, r)isthe vector parallel to the fold axis in an arbitrarily chosenreference frame, then the coordinates ½x yzŠ <strong>of</strong> points in a‘true’ section <strong>and</strong> the coordinates ½x 0 y 0 z 0 Š <strong>of</strong> points on an‘apparent’ section are related to each other by matrixequation:23bq þ cr 2ap 2arD D D2bp ap þ cr 2br½x yzŠ6 D D D¼½x 0 y 0 z 0 Š7642cpD2cqDap þ bqDwhere D ¼ðap þ bq þ crÞ.The 2D coordinates on the chosen section are obtained ascoefficients <strong>of</strong> linear combinations <strong>of</strong> any two perpendicularunit vectors coplanar with the section plane. We developeda simple MATLABw script, which visualises any sectionpr<strong>of</strong>ile across a given fold geometry <strong>and</strong> determine the threeangles a, b <strong>and</strong> g (Fig. 1) for the oblique section pr<strong>of</strong>ile.Using this script we can examine any section across a fixedfold geometry <strong>and</strong> determine which <strong>of</strong> them satisfies thegeometric requirements for shear-b<strong>and</strong>s. The results arepresented in a stereographic projection (Figs. 3 <strong>and</strong> 4)sharing the area containing poles to sections that satisfy theshear-b<strong>and</strong> criteria. The area <strong>of</strong> sections exhibiting shearb<strong>and</strong>like geometry is shaded on the basis <strong>of</strong> the interlimbangle a (Fig. 1).Sections through symmetrical folds generally have foldpr<strong>of</strong>iles with an asymmetric geometry (Fig. 3) <strong>and</strong> theplanes on which the apparent fold geometry satisfies theshear-b<strong>and</strong> criteria fall into two distinct groups separated bythe fold axial plane. These two groups contain poles tosections on which shear-b<strong>and</strong> geometry exhibit opposingasymmetry, i.e. opposing ‘shear-sense’ criteria. The area <strong>of</strong>these domains on a stereographic projection, i.e. the range <strong>of</strong>orientations <strong>of</strong> section planes in space that display shearb<strong>and</strong>geometries is related to the interlimb angle <strong>of</strong> thesymmetrical folds, so as the interlimb angle increasestowards 1808, the range <strong>of</strong> suitable sections displayingshear-b<strong>and</strong> geometry increases.Sections across asymmetric folds that display shear-b<strong>and</strong>75102


O. Lexa et al. / Journal <strong>of</strong> Structural Geology 26 (2004) 155–161 157Fig. 2. (a) Block diagrams <strong>of</strong> folded anisotropy <strong>and</strong> ‘apparent’ fold sections showing shear-b<strong>and</strong> geometries. (b) Orientation <strong>of</strong> apparent fold sectionsexhibiting opposite sense <strong>of</strong> shear criteria. (c), (d) Principle <strong>of</strong> presentation <strong>of</strong> ‘apparent’ fold sections in stereographic projection.geometries are shown in Fig. 4. It can be seen that the range <strong>of</strong>sections suggesting a sinistral <strong>and</strong> dextral ‘sense <strong>of</strong> shear’ isunequal in this example <strong>and</strong> a preferred enlargement <strong>of</strong> onegroup occurs. This asymmetry is controlled by the anglebetween the fold axial plane <strong>and</strong> the main anisotropy (Fig. 4).3. Geological exampleIn the Vepor basement <strong>of</strong> the Central WesternCarpathians, a flat-lying mylonitic foliation containing awell-developed stretching lineation is affected by late103


158O. Lexa et al. / Journal <strong>of</strong> Structural Geology 26 (2004) 155–161Fig. 3. Stereographic projection showing influence <strong>of</strong> interlimb angle. The range <strong>of</strong> section planes showing shear-b<strong>and</strong> geometry increases with interlimb angle<strong>of</strong> folds affecting main anisotropy.Fig. 4. Stereographic projections showing influence <strong>of</strong> angle between fold axial plane <strong>and</strong> main anisotropy. The probability <strong>of</strong> occurrence <strong>of</strong> opposite shearcriteria decreases with decreasing angle <strong>of</strong> axial plane <strong>and</strong> plane <strong>of</strong> main anisotropy.104


O. Lexa et al. / Journal <strong>of</strong> Structural Geology 26 (2004) 155–161 159Fig. 5. Structural elements from studied area. (a) Poles to main Alpine metamorphic anisotropy showing girdle distribution around the axis sub-parallel tohinges <strong>of</strong> late Alpine folds—116 measurements; (b) orientation <strong>of</strong> stretching lineation—68 measurements; (c) distribution <strong>of</strong> poles to main fracture systemsdeveloped in studied area (Hók et al., 2001) overlaid by shaded areas indicating ‘apparent’ fold sections in which shear-b<strong>and</strong> geometry can be observed; (d)rose diagram <strong>of</strong> main fracture directions in studied area (Hók et al., 2001). Note that a main maximum <strong>of</strong> fracture orientations coincides with orientations <strong>of</strong>apparent fold sections. All data are plotted on Schmidt net <strong>and</strong> projected from lower hemisphere. In (a) <strong>and</strong> (b), the contour levels are even multiples <strong>of</strong> st<strong>and</strong>arddeviation.folding, which generated a crenulation cleavage. The foldaxes are generally sub-parallel to the stretching lineation(Fig. 5). Most <strong>of</strong> the studies on the kinematic <strong>and</strong>tectonic evolution <strong>of</strong> this region are based on the use <strong>of</strong>shear-b<strong>and</strong>s <strong>and</strong> asymmetric porphyroclasts as kinematicindicators. These studies have resulted in the generallyaccepted model <strong>of</strong> post-orogenic, orogen-parallel extension(Plašienka et al., 1999; Lupták et al., 2000; Janáket al., 2001). In contrast to the interpretation given byprevious workers in which the major extensionaldeformation post-dates the episode <strong>of</strong> compression, astudy by these authors shows that extension was the firstAlpine deformation in the area <strong>and</strong> that this pre-dates thecompressional stage <strong>of</strong> tectonic evolution (Lexa et al.,2003). The extensional phase is highly asymmetrical <strong>and</strong>locally non-coaxial. The majority <strong>of</strong> the studied outcropsshow that we are dealing with oblique sections acrosssmall-scale folds, which are likely to be misinterpreted asshear-b<strong>and</strong>s (Fig. 6).In order to evaluate the probability <strong>of</strong> encounteringoblique sections with shear-b<strong>and</strong> like geometry in aparticular field area, we need to identify the dominantgeometries <strong>of</strong> the small-scale folds <strong>and</strong>, moreover, wehave to underst<strong>and</strong> the distribution <strong>of</strong> surfaces on whichthe structures are observed. It should be pointed out thatthe majority <strong>of</strong> observations are from natural rather thanman-made outcrops, where the orientation <strong>of</strong> the exposedsurfaces is controlled mainly by fractures (typicallyjoints). In addition, sections that are sub-parallel to thelineations are specifically selected as being appropriatefor the study <strong>of</strong> kinematic indicators. We plotted therange <strong>of</strong> sections with shear-b<strong>and</strong> like geometry, therange <strong>of</strong> naturally occurring fractures <strong>and</strong> the distribution<strong>of</strong> lineations on a stereographic projection (Fig. 5c),which shows that there is a high probability <strong>of</strong> systematicallyobserving oblique sections across the small-scalefolds showing opposite shear-b<strong>and</strong> geometries. Wepropose that such field observations led some authors105


160O. Lexa et al. / Journal <strong>of</strong> Structural Geology 26 (2004) 155–161Fig. 6. Field photos <strong>of</strong> late folding <strong>and</strong> crenulation cleavage affecting early Alpine fabric in the Vepor micaschists <strong>of</strong> the West Carpathians. True fold sections(a) micaschists in the Mútnik area; (b) micaschists in the Katarínska huta area. Apparent fold sections showing shear-b<strong>and</strong>s geometry (c), (d) Katarínska huta.Axial plane: 326/75; fold axis: 52/14; <strong>and</strong> general outcrop surface: 318/65.(e.g. Siman et al., 1996) to interpret the extensionaltectonics in Vepor basement to be symmetrical.4. Discussion <strong>and</strong> conclusionsShear-b<strong>and</strong>s <strong>and</strong> folds can generate identical geometrieswhen seen on a flat exposure surfaces in some orientations.These geometries can lead to misinterpretation <strong>of</strong> folds asshear b<strong>and</strong>s <strong>and</strong> to erroneous <strong>structural</strong> <strong>and</strong> kinematicinterpretation.The ambiguity arises since oblique sections throughsmall-scale folds in anisotropic materials have identicalgeometries to XZ finite strain sections through shear-b<strong>and</strong>bearing rocks. Misinterpretations are most likely in areas <strong>of</strong>multiple deformations when earlier fabrics are folded.Determination <strong>of</strong> the XZ section in phyllonites <strong>and</strong> phyllitesis <strong>of</strong>ten complicated by the presence <strong>of</strong> microscopiccorrugations on foliation surfaces, which may be easilyconfused with stretching lineation. These corrugations arecommonly associated with gentle folds <strong>of</strong> a larger scale.Particularly in this case, geologists would tend to look forkinematic criteria in sections in which the larger folds wouldgenerate pr<strong>of</strong>iles similar to shear b<strong>and</strong> geometry.In order to be confident that the geometry displayed on a2D exposure surface is related to true shear-b<strong>and</strong>s <strong>and</strong> cantherefore be used as a valid kinematic indicator, it isessential to know the 3D geometry <strong>of</strong> the structure <strong>and</strong> theorientation <strong>of</strong> the exposure surface with respect to fold axes<strong>and</strong> lineations. For this reason, systematic studies <strong>of</strong> fracture<strong>and</strong> joint systems in folded areas should accompanykinematic <strong>analyses</strong>.106


O. Lexa et al. / Journal <strong>of</strong> Structural Geology 26 (2004) 155–161 161AcknowledgementsWe are grateful to the Geological Survey <strong>of</strong> the SlovakRepublic for significant financial support during the initialstages <strong>of</strong> our research. This work has been also supported bythe Grant <strong>of</strong> Charles University Agency No. 216/1999/B-GEO. The salaries <strong>of</strong> K. Schulmann <strong>and</strong> O. Lexa werecovered from the grant <strong>of</strong> the Ministry <strong>of</strong> Education No.24313005.ReferencesBehrmann, J.H., 1987. A precautionary note on shear b<strong>and</strong>s as kinematicindicators. In: Cobbold, P.R., Gapais, D., Means, W.D., Treagus, S.H.(Eds.), Shear Criteria in Rocks. Journal <strong>of</strong> Structural Geology 9(5–6).Pergamon, Oxford–New York, International, pp. 659–666.Berthé, D., Choukroune, P., Jegouzo, P., 1979. Orthogneiss, mylonite <strong>and</strong>non coaxial deformation <strong>of</strong> granites; the example <strong>of</strong> South Armoricanshear zone. Journal <strong>of</strong> Structural Geology 1 (1), 31–42.Cobbold, P.R., Cosgrove, J.W., Summers, J.M., 1971. Development <strong>of</strong>internal structures in deformed anisotropic rocks. Tectonophysics 12,23–53.Cosgrove, J.W., 1976. The formation <strong>of</strong> crenulation cleavage. Journal <strong>of</strong> theGeological Society <strong>of</strong> London 132 (Part 2), 155–178.Dewey, J.F., 1965. Nature <strong>and</strong> origin <strong>of</strong> kink-b<strong>and</strong>s. Tectonophysics 1 (6),459–494.Hók, J., Lacika, J., Madarás, J., Kohút, M., Nagy, A., Ivanička, J., Siman,P., Král’, J., Töröková, I., 2001. Neotektonický a geomorfologickývývoj študijných lokalít – 1. čast’. Projekt: Vývoj hlbinného úložiskavyhoreného jadrového paliva a vysokoaktívnych Ra – odpadov vpodmienkach Slovenskej republiky pre obdobie 1998–2000. Úloha:Výber lokality. Číslo etapy: VYL-01-00. Štátny geologický ústavDionýza Štúra, Bratislava. Manuskript–archív ŠGÚDŠ, Bratislava,pp. 1–171.Janák, M., Plašienka, D., Frey, M., Cosca, M., Schmidt, S.T., Lupták, B.,Méres, S., 2001. Cretaceous evolution <strong>of</strong> a metamorphic core complex,the Veporic unit, Western Carpathians (Slovakia): P–T conditions <strong>and</strong>in situ Ar-40/Ar-39 UV laser probe dating <strong>of</strong> metapelites. Journal <strong>of</strong>Metamorphic Geology 19 (2), 197–216.Lexa, O., Schulmann, K., Ježek, J., 2003. Cretaceous collision <strong>and</strong>indentation in SW part <strong>of</strong> West Carpathians:view based on <strong>structural</strong>analysis <strong>and</strong> <strong>numerical</strong> modeling. Tectonics, accepted.Lister, G.S., Snoke, A.W., 1984. S–C mylonites. Journal <strong>of</strong> StructuralGeology 6 (6), 617–638.Lupták, B., Janák, M., Plašienka, D., Schimdt, S.T., Frey, M., 2000.Chloritoid-kyanite schists from the Veporic unit, Western Carpathians,Slovakia: implications for Alpine (Cretaceous) metamorphism. SchweizerischeMineralogische Und Petrographische Mitteilungen 80 (2),213–223.Passchier, C.W., Trouw, R.A.J., 1996. Microtectonics, Springer-Verlag,Berlin, Heidelberg.Plašienka, D.a., Janák, M., Lupták, B., Milovskú, R., Frey, M., 1999.Kinematics <strong>and</strong> metamorphism <strong>of</strong> a Cretaceous core complex: theVeporiuc Unit <strong>of</strong> the Western Carpathians. Physics <strong>and</strong> Chemistry <strong>of</strong>the Earth (A) 24 (8), 651–658.Platt, J.P., 1979. Extensional crenulation cleavage. In: Cobbold, P.R.,Ferguson, C.C. (Eds.), Description <strong>and</strong> Origin <strong>of</strong> Spatial Periodicity inTectonic Structures; Report on a Tectonic Studies Group Conference.Journal <strong>of</strong> Structural Geology 1. Pergamon, Oxford–New York,International, pp. 95–96.Platt, J.P., 1984. Secondary cleavages in ductile shear zones. Journal <strong>of</strong>Structural Geology 6 (4), 439–442.Platt, J.P., Vissers, R.L.M., 1980. Extensional structures in anisotropicrocks. Journal <strong>of</strong> Structural Geology 2 (4), 397–410.Ponce, d.L.M.I., Choukroune, P., 1980. Shear zones in the Iberian Arc. In:Carreras, J., Cobbold, P.R., Ramsay, J.G., White, S.H. (Eds.), ShearZones in Rocks. Journal <strong>of</strong> Structural Geology 2(1/2). Pergamon,Oxford–New York, International, pp. 63–68.Price, N.J., Cosgrove, J.W., 1994. Analysis <strong>of</strong> Geological Structures,Cambridge University Press, Cambridge.Siman, P., Johan, V., Ledru, P., Bezák, V., Madarás, J., 1996. Deformation<strong>and</strong> P–T conditions estimated in “layered migmatites” from southernpart <strong>of</strong> Veporicum crystalline basement (Western Carpathians,Slovakia). Slovak Geological Magazine 3–4/96, 209–213.Simpson, C., Schmid, S.M., 1983. An evaluation <strong>of</strong> criteria to deduce thesense <strong>of</strong> movement in sheared rocks. Geological Society <strong>of</strong> AmericaBulletin 94 (11), 1281–1288.White, S., 1979. Large strain deformation; report on a Tectonic StudiesGroup discussion meeting held at Imperial College, London on 14November 1979. Journal <strong>of</strong> Structural Geology 1 (4), 333–339.107


DTD 5ARTICLE IN PRESSJournal <strong>of</strong> Structural Geology xx (xxxx) 1–24www.elsevier.com/locate/jsgThe quantitative link between fold geometry, mineral fabric <strong>and</strong>mechanical anisotropy: as exemplified by the deformation <strong>of</strong> amphibolitesacross a regional metamorphic gradientLenka Baratoux a, *, Ondrej Lexa a,b , John W. Cosgrove c , Karel Schulmann ba Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University, Albertov 6, Praha 2, 12843, Czech Republicb Université Louis Pasteur, EOST, UMR 7517, 1 Rue Blessig, Strasbourg 67084, Francec Department <strong>of</strong> Earth Science & Engineering, Royal School <strong>of</strong> Mines, Imperial College <strong>of</strong> Science, Technology & Medicine, Prince Consort Road,London SW7 2BP, UKReceived 8 June 2004AbstractThis work shows lateral variations in fold geometry within an amphibolite unit <strong>of</strong> constant mineralogical composition under increasingmetamorphic grade. Analysis <strong>of</strong> the fold geometries indicate: (1) medium amplification associated with low post-buckle flattening in thegarnet zone, (2) high amplification coupled with medium post-buckle flattening in the staurolite zone, <strong>and</strong> (3) passive amplificationdominated by intense post-buckle flattening in the sillimanite zone. A systematic decrease in the mechanical anisotropy <strong>of</strong> the folded fabric isobserved with increase in metamorphic grade. <strong>Quantitative</strong> micro<strong>structural</strong> study shows contrasting deformation micro-mechanismsassociated with folding manifested by: (1) brittle dominated deformation <strong>of</strong> amphiboles that form a stress supporting network with a highcompetence contrast with respect to plagioclase in the garnet zone, (2) ductile dominated heterogeneous deformation <strong>of</strong> an interconnectedweak layer structure with low competence contrast in the staurolite zone, <strong>and</strong> (3) homogeneous deformation <strong>of</strong> a stress supporting frameworkwith low competence contrast in the sillimanite zone. The difference in the folding style between the garnet <strong>and</strong> staurolite zones is associatedwith the lateral variations in microstructure <strong>of</strong> the amphibolites inherited from a pre-folding metamorphic zonation <strong>and</strong> with differentdeformation mechanisms in the hinge zones. However, the change in fold style observed as one moves into the sillimanite zone is controlledby the recrystallization associated with an important syn-folding heat input from an adjacent granite intrusion.q 2005 Elsevier Ltd. All rights reserved.Keywords: Fold geometry; <strong>Quantitative</strong> link; Amphibolite; Metamorphic gradient; Deformation microstructures1. IntroductionThe fold shape in a folded bilaminate is determined bythe ratio <strong>of</strong> competent to incompetent layer thickness <strong>and</strong>the viscosity contrast (m 1 /m 2 ) between them (Ramberg,1961). Ramsay <strong>and</strong> Huber (1987, p. 405) discuss the factorsthat influence the fold geometry <strong>of</strong> such buckled bilaminates.The most important factors are the primaryrheological properties <strong>of</strong> the layers, which depend on themineralogical composition <strong>and</strong> grain size, external parameterssuch as temperature <strong>and</strong> pressure, <strong>and</strong> the* Corresponding author. Tel.: C420-257189509E-mail address: lka@natur.cuni.cz (L. Baratoux).0191-8141/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.doi:10.1016/j.jsg.2005.01.001development <strong>of</strong> a preferred orientation <strong>of</strong> minerals duringdeformation. The seminal work <strong>of</strong> Biot (1961) provided atheoretical treatment capable <strong>of</strong> analysing these systems <strong>and</strong>his work has been successfully applied by geologists toexplain the buckling behaviour <strong>of</strong> both discretely layeredsystems <strong>and</strong> mineral fabrics. In Biot’s work the mostimportant factor controlling the mode <strong>and</strong> style <strong>of</strong> folding isthe mechanical anisotropy <strong>of</strong> the material. It can be shownthat there is a direct link between the mechanical anisotropy<strong>and</strong> the competence contrast <strong>of</strong> a bilaminate (m 1 /m 2 ) <strong>and</strong> thisallows one to use both theories to study their foldingbehaviour (Price <strong>and</strong> Cosgrove, 1990). By analysing thegeometries <strong>of</strong> the studied folds it is possible to determinewhich <strong>of</strong> the two theoretical approaches is more appropriate.Although a large quantity <strong>of</strong> work has been carried out onthe folding mechanics <strong>of</strong> layered systems (e.g. thetheoretical work by Biot (1961, 1963a,b), Ramberg (1963,109


DTD 5ARTICLE IN PRESS2L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–241964) <strong>and</strong> Fletcher (1995), the experimental studies byGhosh (1968), Dubey <strong>and</strong> Cobbold (1977), Dubey (1980),Mancktelow <strong>and</strong> Abbassi (1992) <strong>and</strong> Hobbs (1971) <strong>and</strong> thefield studies by Oertel <strong>and</strong> Ernst (1978) <strong>and</strong> Fowler <strong>and</strong>Winsor (1996)), this work is not directly applicable to thefolding <strong>of</strong> mineral fabrics where the concept <strong>of</strong> layerthickness <strong>and</strong> competence contrast are more difficult todefine. Fortunately a theory <strong>of</strong> the deformation <strong>of</strong>mechanically anisotropic materials has been developed byBiot (1967), which enables the folding <strong>of</strong> mineral fabrics tobe considered.The application <strong>of</strong> modern computer graphics, inparticular geographic information systems (GIS) in quantitativepetrography, <strong>and</strong> computer techniques, which allowthe precise evaluation <strong>of</strong> grain shapes (Panozzo, 1983,1984), grain sizes <strong>and</strong> grain spatial distribution, haveprovided tools for the assessment <strong>of</strong> the relationshipbetween mineral fabrics <strong>and</strong> the folding that developswithin them. In this paper we present a method <strong>of</strong>quantitative micro<strong>structural</strong> analysis that allows the semiquantitativeassessment <strong>of</strong> the mechanical anisotropy <strong>of</strong>mineral fabrics. We demonstrate that quantitative computeraidedanalysis <strong>of</strong> natural fold shapes in conjunction with astudy <strong>of</strong> the internal microstructure allows a better underst<strong>and</strong>ing<strong>of</strong> the mechanical behaviour <strong>of</strong> naturally foldedmultilayers. Our observations show that it is possible todistinguish folds formed by active buckling from thosedominantly controlled by passive amplification. This can beclearly correlated with major variations in mineral microstructures,semi-quantitatively estimated mechanical anisotropy<strong>and</strong> the deformation mechanisms associated withfolding.The rock unit studied in this analysis consists <strong>of</strong>metabasites <strong>of</strong> relatively constant composition that showmetamorphic zonality ranging from the lower amphibolitefacies conditions in the east to the upper amphibolite faciesconditions in the west. The metamorphism wasaccompanied by the development <strong>of</strong> progressively evolvingmineral micr<strong>of</strong>abrics forming a gently dipping to subhorizontalmetamorphic foliation. This consistently orientedbut systematically varying rock anisotropy has beensubsequently affected by horizontal compression responsiblefor the development <strong>of</strong> upright post-metamorphicfolds. In addition, a granite intruded into the western part <strong>of</strong>the amphibolite sheet <strong>and</strong> thermally influenced the host rockwithin a zone several kilometres wide. Field observationsshow a remarkable change in the style <strong>of</strong> the post-peakmetamorphic folding approaching the intrusion (Schulmann<strong>and</strong> Gayer, 2000) <strong>and</strong> the question arises as to whether thechange in fold style reflects the variation in theinherited mechanical anisotropy associated withthe east–west metamorphic–micr<strong>of</strong>abric zonation,whether it is related to the lateral heat input from thewesterly granite, or both.The aims <strong>of</strong> this contribution are: (1) the quantitativecharacterization <strong>of</strong> the fold shapes <strong>and</strong> semi-quantitativedetermination <strong>of</strong> mechanical anisotropy based on quantitativemicro<strong>structural</strong> analysis, (2) the determination <strong>of</strong>possible folding mechanisms using fold shape analysis aswell as the micro-deformational mechanisms associatedwith folding, <strong>and</strong> (3) the evaluation <strong>of</strong> the relative role <strong>of</strong> theinherited paleometamorphic micro<strong>structural</strong> zonation(mechanical anisotropy) <strong>and</strong> lateral heating, on folding.2. Geological settingThe northeastern margin <strong>of</strong> the Bohemian massif is builtup <strong>of</strong> the Neo-Proterozoic basement called Brunia <strong>and</strong> itssedimentary cover (Dudek, 1980) in the east, <strong>and</strong> a belt <strong>of</strong>Brunia derived Variscan nappes (Silesian domain) with highgrade rocks <strong>of</strong> the orogenic root belonging to the Lugi<strong>and</strong>omain in the west (Fig. 1).The studied area is located in the Brunia derived parautochthoncalled the Desná dome, which consists <strong>of</strong>metamorphic Neo-Proterozoic basement core surroundedby a volcano-sedimentary Devonian envelope (Fig. 1b).Metabasites <strong>of</strong> the Jeseník amphibolite massif, the subject<strong>of</strong> this study, form part <strong>of</strong> this Devonian sequence. The hugeaccumulation <strong>of</strong> basic rocks was formed during the EarlyDevonian rifting, as indicated by their tectono-stratigraphicposition <strong>and</strong> back-arc tholeiitic composition (Patočka <strong>and</strong>Valenta, 1990). Metabasites are present on both the eastern<strong>and</strong> western flanks <strong>of</strong> the Desná gneissic core. The basites atthe eastern margin show more tuffitic character while themain amphibolite body situated in the west developedmostly from basalts <strong>and</strong> possibly from gabbros as indicatedby preserved textures in the less deformed domains. Themetabasites are locally stratigraphically or tectonicallyintercalated with metapelites <strong>and</strong> quartzites <strong>of</strong> Pragian age(Chlupáč, 1994), which show the same tectono-metamorphichistory.2.1. Structural geologyThe complex polyphase tectonic evolution <strong>of</strong> the regioncomprises the early pre-Variscan deformations termed D 1 .The first Variscan structures are associated with theBarrovian metamorphism (termed D 2 ), <strong>and</strong> the late Variscanstructures (D 3 ), developed during the transpressionalregime. The most important D 2 structure is represented bya planar metamorphic fabric expressed as localized shearzones in the basement orthogneisses <strong>and</strong> as a penetrative,originally sub-horizontal metamorphic foliation (S 2 ) in allthe rocks <strong>of</strong> the Devonian cover. This fabric is now dippingto the NW or SE at moderate to steep angles as a result <strong>of</strong>subsequent large-scale folding (Fig. 2a). Within the cover,syn-schistose folds (F 2 ) as well as mineral lineation (L 2 ) arelocally well developed.The D 3 deformation had various effects on differentlithologies, depending mainly on the intensity <strong>of</strong> theprevious S 2 anisotropy (Schulmann <strong>and</strong> Gayer, 2000).110


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 3Fig. 1. (a) Geological map <strong>of</strong> the eastern margin <strong>of</strong> the Bohemian Massif. Inset shows location <strong>of</strong> the studied area in the frame <strong>of</strong> European Variscides. (b).Geological map <strong>of</strong> the Silesian domain based on the geological map 1:200,000 by Pouba (1962) <strong>and</strong> Roth (1962). Important thrust faults <strong>and</strong> normal faults areindicated. A–A 0 indicates the position <strong>of</strong> the cross-section in Fig. 2.Basement orthogneisses display rare upright folding.Metabasites <strong>and</strong> metasediments <strong>of</strong> the Devonian volcanosedimentarycover show well-developed mostly asymmetricF 3 folds with axial planes dipping to the NW or SE <strong>and</strong>subhorizontal NE–SW-trending hinges (Fig. 2a). The foldsize ranges from the microscopic grain-scale up to metresscalefolds. In the eastern <strong>and</strong> central parts <strong>of</strong> the Desnádome the D 3 deformation was post-peak metamorphic <strong>and</strong>took place under greenschist conditions as marked bygrowth <strong>of</strong> chlorite, muscovite <strong>and</strong> albite along axialcleavage <strong>of</strong> F 3 folds in metapelites (Schulmann <strong>and</strong>Gayer, 2000). However, it was contemporaneous with theemplacement <strong>of</strong> the Žulová granite in the west (Cháb <strong>and</strong>Žáček, 1994; Schulmann <strong>and</strong> Gayer, 2000), which resultedin the overprinting <strong>of</strong> the S 2 foliation by a NE–SW-trendingsteep <strong>and</strong> penetrative foliation S 2–3 (Fig. 2a).111


DTD 5ARTICLE IN PRESS4L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24Fig. 2. Geological cross-section A–A 0 (shown in Fig. 1b) <strong>of</strong> the Desná dome showing major structures, lithology <strong>of</strong> individual units, <strong>and</strong> major tectonicboundaries (a). Metamorphic zones extended from metapelites are indicated as well as approximate actinolite-out <strong>and</strong> clinopyroxene-in isograds (for Ca-richmetabasites). Equal area lower hemisphere stereoplots are shown for D 2 , <strong>and</strong> D 3 planar <strong>and</strong> linear structures. Each stereoplot contains between 50 <strong>and</strong> 100 polescontoured as multiples <strong>of</strong> uniform distribution. The cross-section is constructed from photographs <strong>and</strong> field notes <strong>and</strong> shows the principal <strong>structural</strong> features.The vertical axis is not to scale. Note that the metamorphic isograds cross-cut lithological boundaries <strong>and</strong> their apparent steep attitude is associated with a largescale F 3 folding <strong>of</strong> their original flat dip as shown by Schulmann <strong>and</strong> Gayer (2000) <strong>and</strong> Štípská et al. (2000). (b) Lower left inset shows PT plot with indicatedages <strong>of</strong> M 3 <strong>and</strong> M 2 metamorphisms <strong>and</strong> associated fabrics; a—Schulmann <strong>and</strong> Gayer (2000); b—Jehlička (1995); c—Maluski et al. (1995). PA — plutonaureole.112


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 52.2. Metamorphic zonalityThe Variscan metamorphic evolution <strong>of</strong> the study areainvolves two main phases; an M 2 metamorphism <strong>of</strong>Barrovian character with the intensity increasing westwards<strong>and</strong> an M 3 metamorphism induced by the Žulová graniteintrusion in the western margin <strong>of</strong> the Desná dome. TheBarrovian M 2 metamorphic grade ranges from chlorite zonein the eastern margin <strong>of</strong> the Desná dome up to staurolite <strong>and</strong>possibly sillimanite zone in the west. The M 3 periplutonicHT/LP overprint also attains its maximum in the westernpart <strong>of</strong> the studied area, where it is documented by thepresence <strong>of</strong> K-feldspar–cordierite migmatites (Rozkošný<strong>and</strong> Souček, 1989; Cháb <strong>and</strong> Žáček, 1994), as well as by thegrowth <strong>of</strong> sillimanite <strong>and</strong> new garnet in the staurolitemicaschists in the pluton aureole (Fig. 2).Although the metabasites cannot be precisely dividedinto particular metamorphic zones once they have reachedamphibolite facies conditions, it is nevertheless necessaryfor the purpose <strong>of</strong> this work to establish a metamorphiczoning <strong>of</strong> the amphibolite massif based on the increasingmetamorphic conditions. The metamorphic zones determinedin intercalated metasedimentary rocks by Souček(1978) are therefore extended into the adjacent metabasites<strong>and</strong> used as a reference for the definition <strong>of</strong> the metamorphicgrade therein. The justification for this procedure is based onthe mutual field relations <strong>of</strong> the amphibolites <strong>and</strong> metasediments,which indicate that both lithologies have experiencedthe same tectonic <strong>and</strong> metamorphic history(Schulmann <strong>and</strong> Gayer, 2000). Therefore, PT conditionswere determined in metasediments applying variousthermobarometry methods (Baratoux, 2004). The degree<strong>of</strong> metamorphism in the east <strong>of</strong> the metabasite massifcorresponds to the garnet zone in the contiguous metasediments<strong>and</strong> the mineral assemblage in the metabasitesincludes hornblende, plagioclase, actinolite, chlorite, <strong>and</strong>epidote corresponding to PT conditions <strong>of</strong> 540G10 8C <strong>and</strong>5G1 kbar. Further to the west actinolite <strong>and</strong> epidotedisappear <strong>and</strong> HblCPlGQtzCIlmCTtn, which correspondsto the staurolite zone, appears with metamorphicconditions <strong>of</strong> M 2 estimated to be 570G30 8C <strong>and</strong> 5.5G1 kbar (Baratoux, 2004). The mineral assemblage in themetabasites corresponding to the sillimanite zone isrepresented by HblCPlGQtzCIlmCTtn <strong>and</strong> AmpCPlGQtzGCpxGCalCIlmCTtn in calcium-rich lithologies <strong>and</strong>metamorphic conditions <strong>of</strong> this zone were estimated to590G20 8C <strong>and</strong> 5.5G1 kbar. In the pluton aureole,sillimanite–cordierite–K-feldspar assemblage occur inmetapelites <strong>and</strong> the PT conditions <strong>of</strong> periplutonic M 3metamorphism reached 700G15 8C <strong>and</strong> 4.2G0.8 kbar (Fig.2b). The mineral assemblage in the amphibolites correspondsto that <strong>of</strong> the sillimanite zone.The age <strong>of</strong> the main fabric-forming M 2 metamorphicevent is difficult to establish in the studied area, but it issupposed to have occurred during the main collisionalevent, which is dated elsewhere at w340 Ma (Schulmann<strong>and</strong> Gayer, 2000; Štípská et al., 2004). The termination <strong>of</strong>the D 2 –M 2 event can be constrained by Rb–Sr dating (335G7.5 Ma) <strong>of</strong> the crystallization <strong>of</strong> the Žulová granite(Jehlička, 1995). However, 40 Ar/ 39 Ar dating <strong>of</strong> muscovite<strong>and</strong> biotite from mylonitic gneisses <strong>of</strong> the Desná dome <strong>and</strong><strong>of</strong> the Žulová granite (Maluski et al., 1995) suggest thatcooling through the white mica <strong>and</strong> biotite closuretemperatures (350 <strong>and</strong> 300 8C, respectively) occurredbetween 310 <strong>and</strong> 300 Ma. Therefore, this age maycorrespond to the greenschist facies F 3 folding activity inthe east <strong>and</strong> to the cooling <strong>of</strong> the Žulová pluton to the west.3. D 2 <strong>and</strong> D 3 amphibolite microstructures acrossmetamorphic zones3.1. Eastern part <strong>of</strong> the massif (the garnet zone)In the eastern part <strong>of</strong> the massif, i.e. in the garnet zone,some weakly deformed metagabbros with large grains <strong>of</strong>hornblende <strong>and</strong> plagioclase (1–2 mm) are still present. Themain metamorphic fabric S 2 <strong>of</strong> amphibolites is preserved indomains unaffected by F 3 folding or in F 3 fold limbs <strong>and</strong> it ischaracterized by a strong mineral shape preferred orientation(SPO) <strong>of</strong> amphibole 0.2–3 mm in size (Fig. 3a). Thesegrains <strong>of</strong>ten show a strong chemical zonality with actinoliticcores <strong>and</strong> tschermakitic rims (Fig. 4a). Fine-grained (30–60 mm) sub-equant plagioclase (An 25–35 ) forms elongatepolycrystalline aggregates surrounded by laths <strong>of</strong> amphibole.These new grains develop from large relict clasts(An 40–50 ) (Fig. 4d). The plagioclase in these domainsexhibits features typical for dynamic recrystallization (in thesense <strong>of</strong> Poirier <strong>and</strong> Guillopé (1979)) such as undulatoryextinction, the development <strong>of</strong> sub-grain boundaries, <strong>and</strong> auniform grain-size distribution. However, a contribution <strong>of</strong>metamorphic nucleation is evidenced by the differentchemical composition <strong>of</strong> the host <strong>and</strong> the new grains(Rosenberg <strong>and</strong> Stünitz, 2003).In the hinge zones <strong>of</strong> F 3 micr<strong>of</strong>olds, amphibole grainswith irregular boundaries are commonly bent <strong>and</strong> broken(Fig. 3b). Micr<strong>of</strong>ractures associated with domainal undulatoryextinction fragment large grains into smaller elongatesegments. There is no difference in plagioclase microstructurein the non-folded foliation <strong>and</strong> in the crenulateddomains. It is concluded, therefore, that the main D 3deformation mechanism for amphibole is fracturing (in thesense <strong>of</strong> Nyman et al. (1992) <strong>and</strong> Stünitz (1993)) <strong>and</strong>passive grain rotation within the plagioclase matrix.3.2. Central part <strong>of</strong> the massif (the staurolite zone)In the eastern part <strong>of</strong> the central zone <strong>of</strong> the massifactinolite is still locally present in the cores <strong>of</strong> dark greengrains <strong>of</strong> magnesio-hornblende. Amphibole compositionfollows the pargasitic line (Fig. 4b) within the S 2 foliation.The size <strong>of</strong> the dark green grains <strong>of</strong> magnesio-hornblende113


DTD 5ARTICLE IN PRESS6L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24114


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 7Fig. 4. The evolution <strong>of</strong> amphibole (a)–(c) <strong>and</strong> plagioclase (d) compositions in the garnet, staurolite <strong>and</strong> sillimanite zones. In (a)–(c) the compositions situatedbelow the pargasitic line (PRG) in the garnet <strong>and</strong> straurolite zones are interpreted as relics <strong>of</strong> previous metamorphic or magmatic stage while those following orsituated above the pargasitic line are thought to be <strong>of</strong> metamorphic origin (for further discussion see Baratoux (2004)). The numbers next to the symbolscorrespond to the analysed samples. Sample 1 in the sillimanite zone is situated in the pluton aureole. Compositions <strong>of</strong> end members edenite (ED), pargasite(PRG), tschermakite (TS), actinolite (AC), <strong>and</strong> tremolite (TR) are indicated.varies from fine-grained (0.01–0.1 mm), through mediumgrained(0.1–1 mm), up to coarse-grained (0.5–3 mm)porphyroblasts. The finer the amphibole grains, the strongertheir SPO <strong>and</strong> alignment are. The sub-equant, weaklyelongate plagioclase <strong>of</strong> An 25–35 composition <strong>and</strong> 20–50 mmin size is either r<strong>and</strong>omly distributed among the amphibolesor forms either elongate polycrystalline aggregates or layersalternating with the amphibole planar fabric (Fig. 3c).Straight boundaries <strong>of</strong> sub-equant plagioclase grainscommonly meet in triple junctions.In the hinge zones <strong>of</strong> F 3 folds, large amphiboles aregenerally bent <strong>and</strong> fractured (Fig. 3d). The main differencewith respect to the previous zone is an important grain sizereduction <strong>of</strong> amphibole down to 0.01–0.05 mm as a result <strong>of</strong>fracturing. New grains attain lower aspect ratios <strong>and</strong>become mixed with plagioclase <strong>of</strong> the same grain size.There is no difference in chemical composition <strong>of</strong> either theplagioclase or amphibole in the hinge zones compared withthe unfolded S 2 fabric.3.3. Western area <strong>of</strong> the massif (the sillimanite zone)In the western area <strong>of</strong> the massif, i.e. in the sillimanitezone, there is a major change in the micro<strong>structural</strong>Fig. 3. Photomicrographs <strong>of</strong> representative microstructures for each zone. (a) Large amphibole grains parallel to the S 2 foliation in the garnet zone aresurrounded by fine-grained recrystallized plagioclase. (b) In the hinge area, amphiboles included within a sub-equant plagioclase matrix grains are bent <strong>and</strong>broken as a result <strong>of</strong> micro-folding. (c) Amphibole <strong>and</strong> plagioclase in the limb area <strong>of</strong> the staurolite zone tend to form alternating aggregates elongated parallelto the foliation. (d) Microcrenulation leads to strong grain size reduction via brittle deformation <strong>of</strong> amphibole resulting in the mixing <strong>of</strong> amphibole <strong>and</strong>plagioclase in the hinge areas. (e) Plagioclase <strong>and</strong> amphibole in the sillimanite zone show straight equilibrated boundaries. Both minerals are arranged parallelto the S 2 foliation. (f) Amphibole <strong>and</strong> plagioclase grains <strong>of</strong> the hinge area deformed by microcrenulation achieve a high aspect ratio. They reorient sub-parallelto the axial plane without any internal deformation. (g) Foam-like texture with straight equilibrated grain boundaries meeting at triple junctions <strong>of</strong> 1208 as wellas a high grain size in both minerals is typical <strong>of</strong> the metabasites <strong>of</strong> the pluton aureole. Amphibole <strong>and</strong> plagioclase tend to form separate monomineral layers.(h) Amphibole <strong>and</strong> plagioclase texture in the hinge zone <strong>of</strong> micr<strong>of</strong>olds is similar to that in the limbs.115


DTD 5ARTICLE IN PRESS8L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24character <strong>of</strong> the metabasites compared with those in thegarnet <strong>and</strong> staurolite zones. Brownish green ferroanpargasitic grains <strong>of</strong> hornblende with aspect ratios <strong>of</strong> 2–4are arranged parallel to the S 2 fabric (Fig. 4c). They displaymutual equilibrated straight grain boundaries typical <strong>of</strong>high-grade amphibolite textures (Brodie <strong>and</strong> Rutter, 1985).Unlike the staurolite zone where the plagioclase (An 30–60 )tends to form layers, we observe isolated grains <strong>and</strong>aggregates <strong>of</strong> plagioclase surrounded by elongate <strong>and</strong> welloriented crystals <strong>of</strong> amphibole (Fig. 3e). Chemical zonality<strong>and</strong> a large span <strong>of</strong> plagioclase compositions evolvingtowards anorthite document prograde metamorphic growthassociated with dynamic recrystallization (Yund <strong>and</strong> Tullis,1991) (Fig. 4d).The hinge zones <strong>of</strong> micro-crenulations are very narrow<strong>and</strong> the amphibole grains tend to reorient parallel with thefold axial planes without any bending <strong>and</strong> fracturing (Fig.3f). The hornblende grains show straight boundaries, localdecussate structures, <strong>and</strong> a similar aspect ratio <strong>and</strong> size tothose within limb zones. Plagioclase reaches slightly higheraspect ratios <strong>and</strong> grain size compared with the limbs. Allthese features together with the absence <strong>of</strong> any compositionalzoning <strong>of</strong> the pargasitic hornblende (Fig. 4c)indicate grain growth related to prograde metamorphism(Vernon, 1976).3.4. Western area <strong>of</strong> the massif (the sillimanite zone <strong>of</strong> thepluton aureole)Unlike the zones described above, brown pargasitichornblende rich in Ti within the sillimanite zone <strong>of</strong> thepluton aureole, exhibits straight or slightly concaveequilibrated boundaries <strong>and</strong> fairly low aspect ratios (1.5–2). It is difficult if not impossible to distinguish in thinsection the S 2 <strong>and</strong> S 3 fabrics. In most <strong>of</strong> the studied samplesthe pargasitic hornblende is associated with plagioclase(An 30–60 ) <strong>of</strong> similar aspect ratios <strong>and</strong> size <strong>and</strong> arranged intoa regular ’static’ foam-like structure (Fig. 3g). The latterexhibits straight growth-related twins. Plagioclase-rich <strong>and</strong>amphibole-rich compositional b<strong>and</strong>s are locally developed.Both minerals attain the same size ranging between 0.05 <strong>and</strong>0.5 mm. 1208 triple point junctions are formed byamphibole–amphibole, plagioclase–plagioclase, <strong>and</strong> evenamphibole–plagioclase grain boundaries. All <strong>of</strong> the featuresdescribed above are consistent with re-equilibration underhigh temperature conditions corresponding to the M 3 HT/LPmetamorphic overprint. The hinges <strong>of</strong> F 3 micr<strong>of</strong>olds areextremely rare but if present they show very similarmicro<strong>structural</strong> relations <strong>of</strong> hornblende <strong>and</strong> plagioclase tothat <strong>of</strong> the main fabric (Fig. 3h). A characteristic feature isthe growth <strong>of</strong> large elongate hornblende crystals parallel tothe axial plane <strong>of</strong> the F 3 micr<strong>of</strong>olds. The texture <strong>of</strong> theserocks bears a remarkable resemblance to the upperamphibolite to granulite facies example <strong>of</strong> Brodie <strong>and</strong>Rutter (1985, p. 155).4. Fold shape analysis4.1. Methods <strong>of</strong> quantitative analysisFor the purpose <strong>of</strong> the fold analysis, 3–6 photographs <strong>of</strong>representative fold types from each metamorphic zone(garnet, staurolite <strong>and</strong> sillimanite with a low degree <strong>of</strong> HToverprint) were selected (Fig. 5). The fold analysis could notbe carried out on the rocks from the pluton aureole because<strong>of</strong> the scarcity <strong>of</strong> macroscopic folds. The fold shapes,redrawn from photographs, were transformed by parallelprojection onto a plane perpendicular to the fold axis. Twoquantitative methods have been applied to quantify the foldshape: the method <strong>of</strong> Lisle (1997) based on Ramsay’s(1967) classification <strong>and</strong> the harmonic fold shape analysis <strong>of</strong>Hudleston (1973). The principles <strong>of</strong> these methods are givenin Appendix A.The method <strong>of</strong> Lisle is based on the polar projection <strong>of</strong>the normalized thickness <strong>of</strong> a folded layer. Each fold can becharacterized by a single number (the index F), whichexpresses the amount <strong>of</strong> flattening within each fold limb.Class 1 folds are characterized by positive F values (0 to N)<strong>and</strong> these give a measure <strong>of</strong> the amount <strong>of</strong> homogeneousflattening perpendicular to the axial plane required togenerate this fold shape from a parallel fold (class 1C forFO1, class 1B for FZ1 <strong>and</strong> class 1A for 0!F!1). Class 3folds are folds with strong thinning <strong>of</strong> the limbs (attenuatedfolds), typically developed in incompetent layers.These folds are characterized by negative F values (0 toKN). Lisle (1997) divided the field <strong>of</strong> class 3 folds intothree subfields with fold shape 3B marking the boundarybetween the newly defined 3A <strong>and</strong> 3C classes. Class 3Bfolds are defined as a ‘pure’ class 3 fold geometry fromwhich the other types (3A <strong>and</strong> 3C folds) develop by thesuperposition <strong>of</strong> flattening strains. Class 3A folds aregenerated from 3B by flattening in the direction normal tothe axial plane <strong>and</strong> have F values ranging from K1toKN.Class 3C folds are the result <strong>of</strong> flattening parallel to the axialplane <strong>and</strong> are relatively rare in the nature.Hudleston’s (1973) fold classification, based on theFourier harmonic analysis <strong>of</strong> folds, exploits the fact that thegeometry <strong>of</strong> a quarter wavelength is sufficient to characterizethat <strong>of</strong> the whole fold. The shape is approximated interms <strong>of</strong> its harmonics <strong>and</strong> it is found that the first two oddharmonics (coefficients b 1 <strong>and</strong> b 3 ) are sufficient toadequately describe the fold’s pr<strong>of</strong>ile shape. b 1 expressesthe amplitude <strong>of</strong> the fold pr<strong>of</strong>ile, b 3 is a measure <strong>of</strong> the‘angularity’ or ‘sharpness’ <strong>of</strong> the fold’s hinge. For sinewaves, b 3 Z0, for box-like folds b 3 O0 <strong>and</strong> for chevron-likefolds b 3 !0. The ratios <strong>of</strong> b 3 /b 1 describe a continuous series<strong>of</strong> shapes between the chevron <strong>and</strong> box-end members. Themethod for calculating b 1 <strong>and</strong> b 3 values is given inAppendix A.In the present analysis <strong>of</strong> fold shape, these two methodshave been combined by plotting b 1 values (a measure <strong>of</strong> theactive amplification <strong>of</strong> the fold) against F values (a measure116


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 9Fig. 5. Field photographs <strong>of</strong> typical fold shapes: (a) garnet zone, (b) staurolite zone, (c) sillimanite zone <strong>and</strong> (d) sillimanite zone <strong>of</strong> the pluton aureole. Chevronfolds are typical for the garnet zone (a), flattened folds with high amplitude are characteristic <strong>of</strong> the staurolite zone (b), <strong>and</strong> similar folds with rather lowamplitudes are present in the sillimanite zone (c). Steep foliation with relic F 3 folds characterize the sillimanite zone in the pluton aureole (d).<strong>of</strong> the post buckle flattening). This graph shows therelationship between these two different expressions <strong>of</strong>shortening. In order to illustrate the geometric implications<strong>of</strong> this graph we have plotted the fold patterns <strong>of</strong> the modelmultilayer sequences given by Ramsay <strong>and</strong> Huber (1987,pp. 414–415) (Fig. 6). This shows clearly the effect <strong>of</strong> theviscosity ratio (m 1 /m 2 ), layer thickness ratio (nZd 1 /d 2 ) <strong>and</strong>the amount <strong>of</strong> superimposed strain on the folds positionon the graph. Folded incompetent layers plot in the left half<strong>of</strong> the diagram <strong>and</strong> the slope <strong>of</strong> the data points shows thestrong negative correlation between the b 1 <strong>and</strong> F parameters.Folded competent layers plot in the right half <strong>of</strong> thediagram <strong>and</strong> the steeper slope <strong>of</strong> the data point trend shows aless pronounced positive correlation between the b 1 <strong>and</strong> Fparameters (Fig. 6).We have used these plots as st<strong>and</strong>ards with which wecompared the plots representing the natural folds from thestudy area. In addition to the qualitative estimate <strong>of</strong> theviscosity contrast <strong>and</strong> the proportion <strong>of</strong> incompetent tocompetent fraction, we can deduce semi-quantitatively theactive amplification <strong>of</strong> the buckle fold <strong>and</strong> the amount <strong>of</strong>post-buckle shortening. The histograms <strong>of</strong> the F values <strong>and</strong>b 3 /b 1 ratios for each <strong>of</strong> the folds studied are given in Fig. 7.4.2. <strong>Quantitative</strong> analysis <strong>of</strong> fold styles in the garnet zoneThe post peak metamorphic folds in the eastern part <strong>of</strong>the massif, i.e. in the garnet zone, display only a smallvariation in shape. The post buckle flattening is small, asdocumented by relatively low positive values <strong>of</strong> F (0.8–3)(Fig. 7a), which approximates to a parallel or weaklyflattened parallel fold. Very few folds in this domain belongto class 3A, i.e. folds that exhibit relatively small negativevalues <strong>of</strong> F (Fig. 7a). The dominance <strong>of</strong> class 1B (i.e.parallel) folds is important as it documents that activebuckling played a major role during the fold amplification<strong>and</strong> that post-buckle flattening was subordinate. The shapesacquired from the harmonic Fourier analysis lie betweenchevron <strong>and</strong> sine wave, as can be seen from the histogram <strong>of</strong>b 3 /b 1 (Fig. 5a), which has an average value <strong>of</strong> b 3 /b 1 ZK0.05. The absolute values <strong>of</strong> both F <strong>and</strong> b 1 are low(F max Z3; b 1max Z3), suggesting a low amount <strong>of</strong> deformation<strong>of</strong> the analysed rocks. A comparison <strong>of</strong> the measuredfold shapes plotted on the b 1 vs. F diagram with the modelmultilayer sequences (Fig. 6) indicates that the fieldexamples show a close resemblance to model folds (type5, i.e. folds characterized by a high viscosity ratio (m 1 /m 2 )117


DTD 5ARTICLE IN PRESS10L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24Fig. 6. The position <strong>of</strong> representative multilayer sequences (modified after Ramsay <strong>and</strong> Huber (1987, pp. 415–416)) in the diagram b 1 vs. F. The relativelycompetent layers are dark in colour. Corresponding symbols are given to the left <strong>of</strong> the schematic drawings; open symbols represent the more deformed stage.The boxed area (F from 4 to K4 <strong>and</strong> b 1 from 0 to 3) has been enlarged. The properties <strong>of</strong> the multilayer sequences are (following Ramsay <strong>and</strong> Huber, 1987): (1)m 1 /m 2 is low, n (the number <strong>of</strong> layers) is high, _A (fold amplification rate) is low, _e is high; (2) m 1 /m 2 is low, n is moderate, _A is moderate, _e is moderate; (3) m 1 /m 2is low, n is low, no characteristic initial wavelength is established; (4) m 1 /m 2 is high, n is high, _A is high, _e is low; (5) m 1 /m 2 is high, n is moderate, _A is high, _e islow. (6) m 1 /m 2 is high, n is low, _A is high, _e is low. Multilayer sequence (3) is missing in our diagram because no folding occurs <strong>and</strong> the deformation ispredominantly layer parallel shortening. (a) <strong>and</strong> (b) in the text refer to the less <strong>and</strong> more deformed stages, respectively.<strong>and</strong> moderate n (d 1 /d 2 ) values). This fold assemblage hasgeometry close to that <strong>of</strong> chevron folds <strong>and</strong> is characterizedby high fold amplification <strong>and</strong> a low degree <strong>of</strong> post-buckleflattening.4.3. <strong>Quantitative</strong> analysis <strong>of</strong> fold styles at the staurolite zoneThe histogram <strong>of</strong> F indexes for the post peak metamorphicfolds in the staurolite zone (Fig. 2) shows that thefolds <strong>of</strong> class 1B (FZ1) are no longer present in the centralpart <strong>of</strong> the massif (Fig. 7b). Class 1C folds are the mostcommon with F values ranging between 1 <strong>and</strong> 5 with anaverage value around 4. There are a number <strong>of</strong> foldsbelonging to class 3A, with F values ranging from K10 toK1 with an average <strong>of</strong> around FZK4. These F valuesindicate that a high amount <strong>of</strong> post-buckle flattening occursin both the competent <strong>and</strong> incompetent layers. Three peakscan be distinguished in the histogram <strong>of</strong> the b 3 /b 1 ratio. Onemaximum is situated close to the chevron shape, a second isclose to the sine wave <strong>and</strong> the third one lies between theshapes <strong>of</strong> a parabola <strong>and</strong> a semi-ellipse (Fig. 7b). However,the majority <strong>of</strong> the folds are situated between the chevron<strong>and</strong> sine wave shapes. This shape diversity indicates thatfold geometries characteristic <strong>of</strong> both competent <strong>and</strong>incompetent materials occur. In the context <strong>of</strong> the mineralfabric in which these folds develop, this indicates that thereare compositional <strong>and</strong> modal differences implying competencecontrasts between layers. The diagram b 1 vs. F alsodemonstrates that the amphibolites in this zone havesuffered an important amount <strong>of</strong> deformation, which is118


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 11Fig. 7. Graphs <strong>of</strong> the fold analysis b 1 vs. F with the inset histograms <strong>of</strong> F <strong>and</strong> b 3 /b 1 coefficients: (a) garnet zone; (b) staurolite zone; (c) sillimanite zone. Typicalfold shapes (sine wave, parabola etc.) expressed by the ratio <strong>of</strong> b 3 /b 1 (Hudleston, 1973) are depicted as vertical lines. See text for further discussion.documented by higher values <strong>of</strong> the b 1 <strong>and</strong> F coefficients (Fig.7b). This fold assemblage corresponds to that <strong>of</strong> model 5b(Fig. 6) a multilayer sequence marked by high viscosity ratio(m 1 /m 2 ) <strong>and</strong> moderate n (d 1 /d 2 ) values. This implies that highfold amplification <strong>and</strong> an important post-buckle flatteningdeveloped in both competent <strong>and</strong> incompetent layers. Thesefolds can be regarded as a more deformed equivalent <strong>of</strong> thefold assemblage observed in the garnet zone (cf. Fig. 7a<strong>and</strong>bwith Fig. 6—model fold types 5a <strong>and</strong> 5b).4.4. <strong>Quantitative</strong> analysis <strong>of</strong> fold styles in the sillimanite zoneWithin the sillimanite zone (Fig. 7c) most <strong>of</strong> the folds are<strong>of</strong> class 1C <strong>and</strong> have F values that are concentrated between119


DTD 5ARTICLE IN PRESS12L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–241 <strong>and</strong> 5 with an average value <strong>of</strong> around 3. Very few class3A folds are present. This indicates the important fact thatthere are relatively few incompetent layers <strong>and</strong> that a highdegree <strong>of</strong> post buckle flattening has occurred (as indicatedby the absence <strong>of</strong> parallel folds). The broad histogram <strong>of</strong> theb 1 /b 3 ratios, with no unequivocal peaks, reveals that theaverage fold shape approximates to that <strong>of</strong> a sine curve. Thegraph <strong>of</strong> b 1 vs. F (Fig. 7c) confirms the low number <strong>of</strong> class3A folds <strong>and</strong> the dominance <strong>of</strong> class 1C. It can be seen fromthis diagram that class 1C folds (FO1) have a low range <strong>of</strong>b 1 values (between 1 <strong>and</strong> 4). In addition, we observe apositive correlation between the F <strong>and</strong> b 1 parameters. Theimplications <strong>of</strong> these observations are that although the foldamplification increases with the degree <strong>of</strong> post-buckleflattening, the maximum value <strong>of</strong> amplification is relativelylow. The measured fold assemblage may be compared withthe model folds <strong>of</strong> type 2b in Fig. 6, which indicate a lowviscosity ratio (m 1 /m 2 ) <strong>and</strong> a moderate value <strong>of</strong> n (d 1 /d 2 ).Compared with the folds in the staurolite zone where highamplification <strong>and</strong> low post-buckle flattening dominate, inthe sillimanite zone we observe limited amplification <strong>and</strong>pronounced post-buckle flattening. This indicates a changefrom folding dominated by active amplification (in thestaurolite zone) to dominantly passive fold amplification inthe sillimanite zone. We interpret this as reflecting animportant change in the rheological properties <strong>of</strong> the foldedmaterial rather than the result <strong>of</strong> differing amounts <strong>of</strong> finitestrain.4.5. Active buckling vs. post buckle flattening in the studiedfoldsThe fold geometries were examined in order to determinethe relative importance <strong>of</strong> active buckling <strong>and</strong> post-buckleflattening across the metamorphic zones <strong>of</strong> the studiedamphibolite unit. It has been shown that in the garnet zonefold amplification was dominated by active buckling withonly a small contribution from post-buckle flattening. In thestaurolite zone, although the folds have also experienced ahigh degree <strong>of</strong> amplification, this involved both activebuckling <strong>and</strong> post-buckle flattening. By comparing theseresults with those obtained from the model folds it can beargued that the folds in the staurolite zone are the equivalent<strong>of</strong> those in the garnet zone (model fold types 5a <strong>and</strong> 5b) buthave experienced a higher degree <strong>of</strong> finite strain. In thesillimanite zone the folds show a relatively lowamplification but important post-buckle flattening eventhough field observations suggest that the strainintensity <strong>of</strong> these folds <strong>and</strong> those <strong>of</strong> the staurolitezone are very similar. It is concluded that thedominance <strong>of</strong> flattening in the folds developed in thesillimanite zone, indicating that the folding wasdominated by passive amplification as opposed to thatwhich occurred in the garnet <strong>and</strong> staurolite zones.5. <strong>Quantitative</strong> analysis <strong>of</strong> rock anisotropyIn order to underst<strong>and</strong> more fully the results <strong>of</strong> themesoscopic fold <strong>analyses</strong> discussed above, it is useful tostudy the petr<strong>of</strong>abrics <strong>of</strong> the folded units. The main goals <strong>of</strong>this study are to evaluate the mechanical anisotropy <strong>of</strong>folded systems <strong>and</strong> to determine the micro-deformationalmechanisms associated with folding.5.1. Methods <strong>of</strong> quantitative micro<strong>structural</strong> analysisApproximately 100 thin sections, collected from all threemetamorphic zones, have been studied in an attempt to showthe relationship between the microstructures <strong>and</strong> thefolding. <strong>Quantitative</strong> micro<strong>structural</strong> analysis has beenapplied to eight representative samples collected from thegarnet, staurolite <strong>and</strong> sillimanite zones <strong>and</strong> from the graniteaureole. Two thin sections, cut perpendicular to the F 3 foldaxes (YZ sections), were taken from the fold hinge <strong>and</strong> foldlimb in each zone, respectively. As noted above, macroscopicfolds are only present in the garnet to sillimanitezones. The degree <strong>of</strong> transposition <strong>of</strong> the original metamorphicfabric within the contact aureole was so high thatthe macroscopic folds are preserved only in domains withhigh lithological contrast. However, relics <strong>of</strong> rootlessmicroscopic folds can be seen in massive amphibolites inthin sections <strong>and</strong>, in this zone, it is these that are analysedfrom a micro<strong>structural</strong> point <strong>of</strong> view.A quantitative micro<strong>structural</strong> analysis <strong>of</strong> grains <strong>and</strong>grain boundaries was carried out on the representativesamples by tracing <strong>and</strong> digitising the outlines <strong>of</strong> individualgrains using the ESRI ArcView 3.2 Desktop GIS environment.The map <strong>of</strong> grain boundaries was generated usingArcView extension Poly (Lexa, 2003). These data havebeen treated by MATLABe PolyLX Toolbox (Lexa, 2003)in which grain boundary <strong>and</strong> grain SPO were analysed usingthe moments <strong>of</strong> inertia ellipse fitting <strong>and</strong> eigen-analysis <strong>of</strong>the bulk orientation tensor techniques (Lexa, 2003).Digitised drawings <strong>of</strong> representative samples are shown inFig. 8 <strong>and</strong> the results from the quantitative micro<strong>structural</strong><strong>analyses</strong> are presented in Table 1. The grain size <strong>of</strong> theminerals was calculated in terms <strong>of</strong> their Ferret diameter <strong>and</strong>the resulting grain size distributions were statisticallyevaluated. The grain size statistics are summarized in Fig.9, which shows median values <strong>and</strong> the quartile difference <strong>of</strong>the Ferret diameters. In addition, a method <strong>of</strong> determiningthe orientation <strong>of</strong> grain boundaries <strong>and</strong> grain shapes,using the eigen-analysis technique, was applied in anattempt to quantify the bulk rock anisotropy. The results <strong>of</strong>the analysis <strong>of</strong> the bulk SPO <strong>of</strong> grains vs. their aspect ratio(R) are presented in Fig. 10. Grain boundary preferredorientation (GBPO) is presented using a diagram <strong>of</strong>eigenvalue ratios (rZe 1 /e 2 ) <strong>and</strong> the orientation <strong>of</strong> thelargest eigenvector <strong>of</strong> the different types <strong>of</strong> grain boundaries(Fig. 11).The grain size analysis is a powerful technique for120


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 13describing the degree <strong>of</strong> grain coarsening related to theintensity <strong>of</strong> metamorphism (Kretz, 1994). Moreover, inpolymineralic tectonites the relative grain size distributionfor different minerals may indicate the degree <strong>of</strong> strain–stress partitioning (H<strong>and</strong>y, 1990; Schulmann et al., 1996).The SPO <strong>and</strong> elongation <strong>of</strong> minerals with a low degree <strong>of</strong>crystallographic anisotropy is generally attributed to theamount <strong>of</strong> strain in a rock (recrystallized quartz, calcite <strong>and</strong>feldspar). However, the degree <strong>of</strong> elongation <strong>of</strong> stronglyanisotropic minerals, such as amphiboles, reflects the degree<strong>of</strong> metamorphism rather than the degree <strong>of</strong> strain. Consequentlya typical feature <strong>of</strong> amphiboles is the decrease <strong>of</strong>axial ratio with increasing metamorphic grade (Brodie <strong>and</strong>Rutter, 1985). Although the degree <strong>of</strong> GBPO is the factormost affiliated to the micro<strong>structural</strong> anisotropy, it isnecessary to discuss all <strong>of</strong> the above-mentioned parametersin order to evaluate the bulk mechanical anisotropy <strong>of</strong> therock.5.2. Grain size distributionThe grain size is expressed as the Ferret diameter <strong>of</strong> thegrain section without stereological corrections. The grainsizes <strong>of</strong> amphibole <strong>and</strong> plagioclase show slightly differentevolutionary trends (Fig. 9). In the limbs (full symbols),amphibole grain size decreases from 31 to 21 mm from thegarnet to the staurolite zone, respectively. The size thenincreases with increasing metamorphic grade via sillimanitezone reaching 80 mm in the contact aureole <strong>of</strong> the granite.The grain size distribution <strong>of</strong> amphiboles in the hinge zones(open symbols) shows a similar trend but the grain sizedifferences between the metamorphic zones are not as highas in the fold limbs. The amphiboles from the fold hinges inthe contact aureole (median value 43 mm) are slightlysmaller than those from fold hinges in the sillimanite zone(median value 46 mm).In the limbs, dynamically recrystallized plagioclase hasthe same grain size in both the garnet <strong>and</strong> staurolite zones(median value 17 mm). The grain size then increases withincreasing metamorphic grade up to a median value <strong>of</strong>65 mm in the contact aureole. As with the amphiboles, thegrain size is always smaller in the fold hinges reaching amaximum median value <strong>of</strong> 35 mm in the contact aureole.Both the average grain size <strong>of</strong> amphibole <strong>and</strong> the varianceare slightly higher than those <strong>of</strong> plagioclase for allmetamorphic zones.In the garnet zone, amphibole shows a larger grain size<strong>and</strong> grain size spread than plagioclase (Fig. 9). Moving tothe staurolite zone the grain size <strong>of</strong> the amphibolesdecreases, resulting in a decreasing difference in grain sizedistribution between the plagioclase <strong>and</strong> amphiboles. Themain characteristics <strong>of</strong> the sillimanite zone are the increasein grain size <strong>of</strong> both minerals with respect to the previouszones coupled with an increase in difference <strong>of</strong> grain sizedistributions. Within the contact aureole the grain sizedistributions <strong>of</strong> coarse-grained plagioclase <strong>and</strong> amphibolebecome equal.5.3. Aspect ratio <strong>and</strong> shape preferred orientationThe microstructures <strong>of</strong> the fold limbs are characterizedby a strong SPO <strong>of</strong> amphibole grains as well as a highaverage aspect ratio (RZ2.05, 2.17 <strong>and</strong> 2.03 in the garnet,staurolite, <strong>and</strong> sillimanite zones, respectively) (Fig. 10).Only in the highest metamorphic grade does the SPOdecrease, together with the aspect ratio (down to RZ1.57).Plagioclase displays the same trend as the amphibole but theabsolute values <strong>of</strong> their aspect ratios <strong>and</strong> SPO are noticeablylower (RZ1.46, 1.64 <strong>and</strong> 1.59 in the garnet, staurolite <strong>and</strong>sillimanite zones, respectively). As with the amphiboles, theaspect ratio <strong>and</strong> SPO <strong>of</strong> the plagioclase decreases in thecontact aureole (RZ1.49, i.e. similar to the R value <strong>of</strong>amphibole in this zone).The SPO <strong>of</strong> amphibole in the hinge zones shows acomplex evolution. In the garnet zone the SPO is still high,but the aspect ratio is rather low (RZ1.81). In the staurolitezone both the SPO <strong>and</strong> the aspect ratio (RZ1.89) decreasewith respect to the limb zone. In contrast the fold hinges inthe sillimanite zone are associated with rather high SPO <strong>and</strong>very high aspect ratio with respect to the previous zone (RZ2.36). Within the contact aureole it is found that the SPO <strong>of</strong>amphibole <strong>and</strong> the aspect ratio (RZ1.60) diminish in thehinge areas <strong>of</strong> the micr<strong>of</strong>olds. A similar trend can beobserved in the plagioclase (RZ1.48, 1.53, 1.76 <strong>and</strong> 1.41for the garnet, staurolite, sillimanite zones <strong>and</strong> the contactaureole, respectively).5.4. Grain boundary preferred orientationA study <strong>of</strong> the evolution <strong>of</strong> amphibole–amphibole grainboundary preferred orientation (GBPO) shows two importanttrends. The first is the GBPO in the fold limbs which,apart from the staurolite zone, decreases with increasingmetamorphic grade (rZ2.12, 2.39, 1.76 <strong>and</strong> 1.23 for thegarnet, staurolite, sillimanite zones <strong>and</strong> the contact aureole,respectively) (Fig. 11). The second trend shows that in thegarnet <strong>and</strong> staurolite zones the GBPO in the hinge domainsis at a high angle to that <strong>of</strong> the axial plane representing S 3 .Itcan be seen that in the garnet <strong>and</strong> staurolite zones the degree<strong>of</strong> GBPO decreases noticeably in the hinge areas <strong>of</strong> foldscompared with the limb areas (rZ1.47 <strong>and</strong> 1.43, respectively).In contrast, in the sillimanite zone, the GBPO in thehinge zone becomes parallel or sub-parallel to the axialplane (S 3 ) <strong>and</strong> shows the same intensity (rZ1.7) to that inthe limbs. Similarly, in the contact aureole, the GBPO in thehinge zones shows parallelism to the axial plane (S 3 ), <strong>and</strong>has a similar intensity (rZ1.22) to that on the limbs.The r values for plagioclase–plagioclase boundaries areweak for all the zones (1.06–1.21; see Fig. 11 <strong>and</strong> Table 1)indicating a very weak or even a lack <strong>of</strong> GBPO. In contrast,the amphibole–plagioclase GBPO, which is controlled by121


DTD 5 ARTICLE IN PRESS14L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24Fig. 8. Typical microstructures from the four metamorphic zones, redrawn <strong>and</strong> digitized in the ArcView GIS environment. The rose diagrams represent grain long axis distribution. Full symbols correspond t<strong>of</strong>old limbs, open symbols to fold hinges. These symbols have the same significance in Figs. 9 <strong>and</strong> 11. The minerals can be identified as follows: plagioclase—white; amphibole—medium grey; spheneCilmenomagnetite—black; quartz—vertically hatched; calcite—dotted; clinopyroxene—light grey.122


Table 1Statistical values <strong>of</strong> the quantitative textural analysisZone Grt St SilSil (plutonaureole)Sample 162 163 84 84 10 10 149 4Limb Hinge Limb Hinge Limb Hinge Limb HingeGBPO Pl–pl 1.06 1.10 1.21 1.21 1.25 1.12 1.10 1.22Eigenvalue rZ Amp–amp 2.12 1.47 2.39 1.43 1.76 1.70 1.23 1.22e1/e2Amp–pl 1.45 1.17 1.95 1.26 1.48 1.52 1.33 1.08Eigenvector Pl–pl K25 K74 22 88 K9 K65 28 83orientation (8) a Amp–amp K11 K22 K3 K76 K11 K77 74 K83Amp–pl K16 K2 K4 89 K14 K78 91 K73SPO Pl 1.21 1.06 1.57 1.25 1.42 1.44 1.23 1.16Eigenvalue r Amp 2.39 1.57 3.06 1.44 2.16 1.88 1.44 1.31Eigenvector Pl K16 K38 1 K90 K9 K81 89 K90orientation (8) a Amp K11 K3 K6 K88 K15 K79 79 K86Aspect ratio Pl 1.46 1.48 1.64 1.53 1.59 1.76 1.49 1.41(median)Amp 2.05 1.81 2.17 1.89 2.03 2.36 1.57 1.60Grain size—FerretdiameterMedian (mm) Pl 17 22 17 15 39 28 65 35Amp 31 33 21 20 64 46 80 43Q1 (mm) Pl 11 14 10 10 27 20 36 21Amp 16 18 13 13 38 29 45 29Q 3 (mm) Pl 25 34 25 21 52 39 114 53Amp 75 59 35 32 104 79 130 65Q3KQ1 (mm) Pl 14 20 15 12 25 19 79 33Amp 59 41 22 19 66 49 85 37a Positive values are oriented anticlockwise with respect to the horizontal.L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 15DTD 5 ARTICLE IN PRESS123


DTD 5ARTICLE IN PRESS16L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24Fig. 9. Plot <strong>of</strong> amphibole <strong>and</strong> plagioclase median <strong>of</strong> Ferret diameter <strong>and</strong> variances for all the metamorphic zones studied. Full symbols correspond to fold limbs<strong>and</strong> open symbols represent hinge areas. Samples from the pluton aureole are assigned as Sil*.the difference between the grain size <strong>of</strong> the large elongategrains <strong>of</strong> amphibole <strong>and</strong> small grains <strong>of</strong> plagioclase as wellas by the SPO <strong>of</strong> the amphibole, is higher. The highestGBPO for these boundaries is developed in the limb zones<strong>of</strong> the folds in the staurolite zone (rZ1.95). The degree <strong>of</strong>GBPO is low in the limb zones <strong>of</strong> folds in the garnet <strong>and</strong>sillimanite zones (rZ1.45 <strong>and</strong> 1.48, respectively) <strong>and</strong>further decreases to rZ1.33 in the contact aureole. In thehinge regions the GBPO <strong>of</strong> plagioclase–amphibole boundariesis very low in the garnet, staurolite zones <strong>and</strong> contactaureole (rZ1.17, 1.26 <strong>and</strong> 1.08, respectively; Fig. 11). It isonly in the hinge zones <strong>of</strong> the folds in the sillimanite zonethat the GBPO <strong>of</strong> this boundary exceeds 1.5 (rZ1.52).5.5. A comparison <strong>of</strong> the microstructures in the hinge <strong>and</strong>limb areas for folds from the different metamorphic gradesThe analysis <strong>of</strong> the limb areas <strong>of</strong> the folds in the garnetzone reveals a large grain size <strong>and</strong> aspect ratio differencebetween the amphiboles (elongate <strong>and</strong> large) <strong>and</strong> plagioclase(sub-equant <strong>and</strong> small) (Figs. 9 <strong>and</strong> 10). This workalso shows a relatively small SPO <strong>and</strong> GBPO for plagioclasecompared with strong SPO <strong>and</strong> GBPO for amphibole. Thismicrostructure represents a network <strong>of</strong> large amphibolecrystals surrounding pockets <strong>of</strong> fine-grained plagioclase. Inthe hinge regions <strong>of</strong> folds in the garnet zone the amphibolesare oriented at high angles to the axial plane. Compared withthe limb there is a decrease in the aspect ratio, grain size,SPO <strong>and</strong> GBPO <strong>of</strong> both minerals.In the limbs <strong>of</strong> the folds in the staurolite zone both theamphiboles <strong>and</strong> the plagioclase display a very intense SPO,high aspect ratio <strong>and</strong> a strong GBPO coupled with adecreasing difference in grain size between the two mineralscompared with the fold limbs in the garnet zone (Figs. 9–11). This reflects the fact that both minerals forminterconnecting aggregates. The hinge zone microstructure,like that in the garnet zone, is marked by rotation <strong>of</strong> thefabric into an orientation at a high angle to the fold axialplane <strong>and</strong>, compared with the limbs, the mineral aspectratio, grain size, SPO, <strong>and</strong> GBPO are all reduced.The sillimanite zone <strong>and</strong> the contact aureole both showan increase in grain size for both minerals compared withthe garnet <strong>and</strong> staurolite zones (Figs. 9–11). However, themain difference is the parallelism <strong>of</strong> the mineral preferredorientation in the hinge zones with the fold axial plane,which is expressed by both the SPO <strong>and</strong> the GBPO (Figs. 8<strong>and</strong> 11). This is connected with an increase in aspect rati<strong>of</strong>or both plagioclase <strong>and</strong> amphibole in the hinge comparedwith the limb. The other important feature is the similarGBPO for like–like <strong>and</strong> unlike boundaries for both the hinge<strong>and</strong> limb zones. The only difference between the sillimanitezone <strong>and</strong> contact aureole is a weakening <strong>of</strong> the SPO <strong>and</strong>GBPO in the latter where the aspect ratio <strong>of</strong> both mineralsapproaches 1. The distribution <strong>of</strong> amphibole <strong>and</strong> plagioclaseaggregates in the limbs is rather r<strong>and</strong>om in the sillimanitezone <strong>and</strong> tends to be layered in the contact aureole.In summary, we note that the rocks <strong>of</strong> the study area fallinto two distinct groups. Those in the garnet <strong>and</strong> staurolitezones, which show mineral preferred orientation in thehinge zone at a high angle to the fold axial plane <strong>and</strong> to the124


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 17Fig. 10. The plot <strong>of</strong> grain shape preferred orientation (SPO) <strong>of</strong> amphibole <strong>and</strong> plagioclase displaying the weighted ratio <strong>of</strong> the eigenvalues <strong>of</strong> inertia. The resultsare summarized in a boxplot-type diagram <strong>of</strong> aspect ratios vs. eigenvalue ratio <strong>of</strong> bulk matrix <strong>of</strong> inertia <strong>of</strong> the individual minerals. The individual boxes showthe median, first <strong>and</strong> third quartiles <strong>of</strong> the aspect ratio. The ‘whiskers’ represent statistical estimates <strong>of</strong> the data range. Outliers are not plotted. Boxes <strong>of</strong> thelimbs are shaded while those <strong>of</strong> hinge areas are white. The number <strong>of</strong> the analysed sample is indicated for each zone. Samples from the pluton aureole areassigned as Sil*.SPO <strong>of</strong> amphibole <strong>and</strong> plagioclase on the limbs, <strong>and</strong> those inthe sillimanite zone <strong>and</strong> contact aureole, which showmineral preferred orientation in the hinge zone sub-parallelto the axial plane.6. DiscussionIn this paper we have attempted to use the theory <strong>of</strong>buckling <strong>of</strong> multilayers <strong>and</strong> the theory <strong>of</strong> buckling <strong>of</strong>mechanically anisotropic materials to explain the foldingmechanisms <strong>of</strong> some natural folds, which show lateralvariations in their shape. The folds are developed in a rockunit <strong>of</strong> constant mineral composition but which exhibits alateral variation in rock microstructure related to metamorphicgrade. Systematic variations in fold shapes acrossthe whole pr<strong>of</strong>ile suggest that the original variations in rockprotolith are less important for folding mechanisms than themetamorphic recrystallization associated with development<strong>of</strong> mineral fabrics prior to the folding event.6.1. Competence contrast vs. mechanical anisotropyThe question arises as to whether the differences betweenthe fold styles in the garnet <strong>and</strong> staurolite zone, <strong>and</strong> folds inthe sillimanite zone, are the result <strong>of</strong> differences in themechanical anisotropy <strong>of</strong> the mineral fabric or <strong>of</strong> rheologicalvariations between adjacent layers. As is discussed inthe following section, an inspection <strong>of</strong> Figs. 6 <strong>and</strong> 7 can helpanswer this.The buckling behaviour <strong>of</strong> a multilayer is controlled bythe ratio <strong>of</strong> the competencies, e.g. the viscosities (m 1 /m 2 )<strong>of</strong>the competent <strong>and</strong> incompetent layers <strong>and</strong> their relativethicknesses (e.g. Ramberg, 1963). The most important125


DTD 5ARTICLE IN PRESS18L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24Fig. 11. Plot <strong>of</strong> grain boundary preferred orientation (GBPO) <strong>of</strong> amphibole–amphibole, amphibole–plagioclase <strong>and</strong> plagioclase–plagioclase boundariesdisplaying weighted ratio <strong>of</strong> eigenvalues <strong>of</strong> inertia. The numbers represent the orientation <strong>of</strong> the eigenvector V 1 <strong>of</strong> GBPO with respect to the horizontal.Positive values indicate anticlockwise deviation. Samples from the pluton aureole are labelled Sil*.factor controlling the mode <strong>and</strong> style <strong>of</strong> folding <strong>of</strong> amineral fabric is the mechanical anisotropy <strong>of</strong> thematerial <strong>and</strong> this is defined by the ratio <strong>of</strong> two moduli,one a measure <strong>of</strong> the resistance to layer or fabricparallel compression (M) <strong>and</strong> the other to the resistanceto shear in the same direction (L).It can be shown that there is a direct link betweenthe anisotropy (M/L) <strong>and</strong> the competence contrast <strong>of</strong> abilaminate (m 1 /m 2 )(seePrice <strong>and</strong> Cosgrove, 1990) <strong>and</strong>this allows one to use both theories to study foldingbehaviour. By analysing the geometries <strong>of</strong> the studiedfolds it is possible to determine which <strong>of</strong> the twotheoretical approaches is more appropriate. As shownearlier, the low F values <strong>of</strong> the folds from the garnet(low b 1 ) <strong>and</strong> staurolite zones (high b 1 ) are similar tothose in the range type 5a <strong>and</strong> 5b to type 4b (Fig. 6).The trend <strong>of</strong> the arrows in Fig. 7b (which indicates theratio <strong>of</strong> active buckling amplification <strong>and</strong> fold flatteningin the staurolite zone) is comparable with the trendsindicated in Fig. 6 for a multilayer (type 5b) with ahigh competence contrast between adjacent layers. Theconstruction <strong>of</strong> the dip isogons for the folds from thiszone shows alternations <strong>of</strong> layers with differentgeometries (classes 1B <strong>and</strong> 1C alternating with class3) <strong>and</strong> different curvature pr<strong>of</strong>iles <strong>of</strong> the hinge regions(Fig. 12), a pattern compatible with the folds <strong>of</strong> type5b. Using the classical theory <strong>of</strong> Ramberg (1963), thispattern can be interpreted in terms <strong>of</strong> folding <strong>of</strong> amultilayer sequence marked by alternation <strong>of</strong> layerswith a high competence contrast.However, as was shown above, the garnet zoneamphibolites are more or less compositionally homogeneous<strong>and</strong> do not show distinct compositional layering.Inspection <strong>of</strong> Fig. 7a shows that the folds in the garnet zonecould be considered to represent the early stages <strong>of</strong> active<strong>and</strong> passive amplification observed in the staurolite zone.However, there is a marked difference in the density <strong>of</strong> thedata relating to the ‘competent’ <strong>and</strong> ‘incompetent’ members<strong>of</strong> the multilayer. When the pattern <strong>of</strong> dip isogons for arepresentative fold pr<strong>of</strong>ile for these folds is examined (Fig.12), it can be seen that there is an important increase <strong>of</strong> foldflattening component resulting in a convergence <strong>of</strong> shapes(<strong>and</strong> therefore dip isogon patterns) <strong>of</strong> adjacent layers.A comparison <strong>of</strong> the natural data from folds in thesillimanite zone (Fig. 7c) with the model folds in Fig. 6shows that the best fit is with type 2b folds. Thisgeometry is traditionally interpreted as indicating abilaminate with a low competence contrast. However,we note the strong asymmetry in the density <strong>of</strong> pointsindicating a lack <strong>of</strong> incompetent ‘layers’. This isconfirmed by inspection <strong>of</strong> the dip isogon pattern <strong>of</strong>representative folds from this zone (Fig. 12). Althoughthe pattern shows alternations <strong>of</strong> fold shapes <strong>of</strong> class 1Cwith those <strong>of</strong> class 3, it is obtained by analysing‘layers’ that are composed <strong>of</strong> several units. The absence<strong>of</strong> alternations <strong>of</strong> clearly defined individual layerscombined with the high degree <strong>of</strong> flattening <strong>and</strong>relatively low fold amplification indicate that thissystem approximates more closely to the model analysedby Biot than to a bilaminate.126


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 19Fig. 12. Selected fold assemblages from the garnet, staurolite <strong>and</strong> sillimanite zones, respectively. (a) Dip isogon patterns. (b) Graphs showing the changes in F(degree <strong>of</strong> fold flattening) across successive folded layers for each fold assemblage. (c) b 1 vs. F plots constructed for the presented folds showing the relativeimportance <strong>of</strong> fold flattening <strong>and</strong> active fold amplification.6.2. Interpretation <strong>of</strong> the deformation micro-mechanismsassociated with foldingExperimental work on amphibole <strong>and</strong> plagioclase as wellas field observations have shown that the former mineral isstronger than the latter under the complete range <strong>of</strong>homologous temperatures (see Brodie <strong>and</strong> Rutter (1985)for a review). Observations <strong>of</strong> the microstructures in thehinge zone show that the micro-folding in the garnet zonehas been achieved by the bending <strong>and</strong> fracturing <strong>of</strong> thestrong amphiboles whilst the relatively weak plagioclaserecrystallizes <strong>and</strong> accommodates space modificationsassociated with the rigid body rotation <strong>of</strong> the amphibolecrystals in a weak matrix by mechanism modelled byArbaret et al. (2001), for example (Table 2). However, n<strong>of</strong>racturing or bending <strong>of</strong> amphiboles is observed in limbzones, which implies that the fabric on the limbs stillrepresents the original, i.e. pre-folding, fabric. Folding <strong>of</strong>amphiboles in the garnet zone is thus similar to thatdescribed from low-grade metasediments <strong>and</strong> which leadsto the occurrence <strong>of</strong> metamorphic differentiation (Cosgrove,1976). In this work, Biot’s (1961) theory was used toaccount for the buckling <strong>of</strong> a greenschist facies mineralfabric made up <strong>of</strong> a rigid stress supporting framework <strong>of</strong>mica containing interstitial weak quartz. In the hinge theframework <strong>of</strong> mica protects the quartz from the buckling127


DTD 5ARTICLE IN PRESS20L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24stress but on the limbs it does not. The resulting stressgradients cause the migration <strong>of</strong> quartz from the limb to thehinge <strong>and</strong> mica from the hinge to the limb.Micro<strong>structural</strong> examination <strong>of</strong> the rocks in the presentstudy area shows an analogous situation in the amphiboleplagioclase rock. The resulting stress gradients establishedin both the amphibole <strong>and</strong> the plagioclase cause a migration<strong>of</strong> plagioclase to the hinge <strong>and</strong> amphibole to the limbs,leading to metamorphic differentiation.We suggest that this process operated in the hinge zones<strong>of</strong> some macr<strong>of</strong>olds leading to their thickening. Because <strong>of</strong>strain compatibility, this hinge thickening develops inalternating layers, thus producing patterns similar to thefolding <strong>of</strong> bilaminate (Fig. 12).In the hinge zones <strong>of</strong> folds in the staurolite zone aconspicuous reduction in the grain size <strong>of</strong> the amphiboleswith respect to the limb zones can be observed. This islocally connected with the reorientation <strong>of</strong> amphibolefragments, <strong>and</strong> we interpret this microstructure as beingthe result <strong>of</strong> the fracturing <strong>and</strong> rigid body rotation <strong>of</strong>amphiboles as described by Nyman et al. (1992). Thedynamically recrystallized plagioclase in the hinge zonesalso shows a decrease in grain size with respect to the limb.In the hinges, both minerals have similar grain size, lowaspect ratio, very weak SPO <strong>and</strong> GBPO <strong>and</strong> exhibit animportant degree <strong>of</strong> mixing. We suggest that these featuresmay stimulate a change in deformation mechanism in thesedomains <strong>and</strong> that in the highly attenuated folds thefracturing <strong>of</strong> amphiboles <strong>and</strong> the dynamic recrystallization<strong>of</strong> plagioclase switch to a type <strong>of</strong> granular flow facilitatingthe development <strong>of</strong> high strain intensities in these areas(Table 2). The changeover from dislocation creep dominatedflow to granular flows connected with a reduction <strong>of</strong>grain size <strong>and</strong> the mechanical mixing <strong>of</strong> minerals has beendescribed by several authors in greenschist facies metabasitemylonites (e.g. Stünitz, 1993; Berger <strong>and</strong> Stünitz,1996). As with the folds in the garnet zone we argue that thefabric on the limbs <strong>of</strong> the folds in the staurolite zonerepresents the original, pre-fold fabric.In contrast to the garnet <strong>and</strong> staurolite zones no bendingor fracturing <strong>of</strong> the amphibole lattice was noted in the hingezones <strong>of</strong> folds in the sillimanite zone. Instead the fold hingeareas in the sillimanite zone display crossover growths <strong>of</strong>amphiboles <strong>and</strong> straight, well-equilibrated grain boundaries<strong>of</strong> all minerals. In addition, the amphibole grains show nosigns <strong>of</strong> internal deformation <strong>and</strong> the grain size is alwayshigher than that observed in the garnet <strong>and</strong> staurolite zones.All these criteria indicate that these grains developed by themechanism <strong>of</strong> nucleation <strong>and</strong> possibly syn-deformationalgrowth (Vernon, 1976; Rosenberg <strong>and</strong> Stünitz, 2003). Theplagioclase also shows well-equilibrated grain boundariesmeeting at triple point junctions <strong>and</strong> an almost uniformdistribution in the rock, features consistent with the highgradecrystallization <strong>of</strong> amphiboles <strong>and</strong> plagioclase (Brodie<strong>and</strong> Rutter, 1985). Comparison <strong>of</strong> the mineral microstructures<strong>of</strong> the hinge zones <strong>and</strong> the limbs shows that the hingezones display higher aspect ratios, smaller grain sizes,similar SPO <strong>and</strong> like–like <strong>and</strong> unlike GBPOs (Table 2). Thisimplies that, during the development <strong>of</strong> the folds, recrystallizationoccurred simultaneously in both the hinge zones<strong>and</strong> the limbs, producing an increase in elongation <strong>of</strong> theoriginal crystals in the hinge zone. This is because the grainsin both the limb regions <strong>and</strong> the hinge regions are parallel tothe axial plane, i.e. are perpendicular to the largest principalcompressive stress axis. This resulted in an increase inaspect ratio by the process <strong>of</strong> heterogeneous dissolutionaccompanied by recrystallization <strong>and</strong> grain growth, themechanism <strong>of</strong> schistosity transposition well known in highgrade schists <strong>and</strong> described by a number <strong>of</strong> authors (seePasschier <strong>and</strong> Trouw (1996, fig. 4.17) for a review). Unlikethe folds in the garnet <strong>and</strong> staurolite zones where weconsider that the fabric on the fold limbs represents the prefoldingmicrostructure, in the sillimanite zone it is modifiedby dissolution <strong>and</strong> growth process.In the contact aureole a similar type <strong>of</strong> micro<strong>structural</strong>pattern to the sillimanite zone occurs but it is marked by acomplete loss <strong>of</strong> SPO <strong>and</strong> GBPO coupled with an importantgrain size increase (Table 2). These features are consistentwith there having been an important contribution from highgrade “static” recrystallization in both the hinge <strong>and</strong> limbdomains. However, we emphasise that one can still observesmaller grain sizes <strong>and</strong> aspect ratios for both amphibole <strong>and</strong>plagioclase in the hinge zones compared with the limbs.7. ConclusionsA summary <strong>of</strong> the fold analysis <strong>and</strong> the micro<strong>structural</strong>studies supporting the fold mechanics models is given inTable 2.The garnet grade region. The quantitative micro<strong>structural</strong>analysis <strong>of</strong> the limb zones in the folds in the garnetzone reveals the possible existence <strong>of</strong> a stress-supportingnetwork. This type <strong>of</strong> structure implies a relativelyhomogeneous stress distribution in the rock (Jordan,1988), which is controlled by the rheologically resistantamphiboles, with the plagioclase representing only weakpockets that deform to accommodate the deformationimposed by the strong amphibole. This is supported by thelarge grain size difference between the recrystallizedplagioclase <strong>and</strong> amphibole crystals, which can be interpretedin terms <strong>of</strong> a stress-supporting network with a highrheological contrast between the plagioclase <strong>and</strong> amphibole(H<strong>and</strong>y, 1990). In addition, the micro<strong>structural</strong> analysisshows that the amphibole stress-supporting frameworkcollapses in the hinge zones by brittle failure. This brittledeformation, which represents an extreme example <strong>of</strong> strainlocalization, tends to produce sharp-hinged chevron folds ona grain scale (Fig. 3a <strong>and</strong> b).The observed folding mechanism also gives rise tochevron folds on a macroscopic scale indicating that themechanical anisotropy <strong>of</strong> the rock was high. To interpret the128


Table 2Summary <strong>of</strong> parameters derived from fold shape <strong>and</strong> micro<strong>structural</strong> <strong>analyses</strong> in all zones. Used symbols are: A, Amp—amphibole; P, Pl—plagioclase; S—grain size; R—aspect ratio; SPO—shape preferredorientation; GBPO—grain boundary preferred orientationFoldingmechanismsDegree <strong>of</strong>mechanicalanisotropyQuantitavivemicrostructureparametersMicrostructureDeformationmechanismsGarnetzoneStaurolitezoneSillimanitezone <strong>and</strong>pluton aureoleActive amplification Active amplification <strong>and</strong> post-buckle flattening Passive amplificationLow to medium b1. Low positive F Medium to high b1. High F Low to medium b1. High positive FHigh mechanical anisotropy Bilaminate Low mechanical anisotropyS R SPO GBPO S R SPO GBPO S R SPO GBPOLimb AOP AOP AOP AOP AZP ARP ARP ARP AZP AZP AZP AZPHinge AOP AOP AOP AOP AZP ARP ARP AOP AZP AZP AZP AZPAmphibole supported LBF (large elongate <strong>and</strong> interconnectedIWL microstructure with low-viscosity contrast (alter-Amphibole supported LBF structure with low-viscosityAmp grains surrounding pockets <strong>of</strong> finenation<strong>of</strong> Pl-rich <strong>and</strong> Amp-rich aggregates)contrast (mixture <strong>of</strong> Amp <strong>and</strong> Pl <strong>of</strong> equal size <strong>and</strong> aspectgrained Pl)ratio)Hinge Limb Hinge Limb Hinge <strong>and</strong> limbAmp fracturing Brittle reactivation <strong>of</strong> S 2fabricAmp fracturing <strong>and</strong> granularflow <strong>of</strong> Amp–PlmatrixDynamic recrystallization<strong>of</strong> Pl <strong>and</strong> ductile flow inPl-rich layersNucleation <strong>and</strong> syndeformational growthL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 21DTD 5 ARTICLE IN PRESS129


DTD 5ARTICLE IN PRESS22L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24isogon patterns (Fig. 12) we propose an explanation <strong>of</strong>alternations <strong>of</strong> hinge zones thickened by this mechanism <strong>of</strong>micr<strong>of</strong>olding with those without any thickening <strong>and</strong> distinctmicro<strong>structural</strong> changes.The staurolite grade region. In the limbs <strong>of</strong> foldsgenerated in the staurolite zone, the pre-folding fabric ismarked by fairly well developed, alternating elongateaggregates <strong>of</strong> plagioclase <strong>and</strong> amphibole. The intensity <strong>of</strong>mineral preferred orientation is very strong, as documentedby the high aspect ratio <strong>and</strong> strong GBPO <strong>and</strong> SPO values.Inspection <strong>of</strong> the fabric in thin sections (Fig. 3c <strong>and</strong> d)suggests that this structure approximates to the so-called‘interconnected weak layer structure’ (IWL <strong>of</strong> H<strong>and</strong>y(1990)) characterized by an alternation <strong>of</strong> relatively strongamphibole rich domains <strong>and</strong> relatively weak domains rich inplagioclase. Because the alternating domains differ only insmall modal differences in amphiboles <strong>and</strong> plagioclase thisstructure represents an IWL structure with a low viscositycontrast as defined by H<strong>and</strong>y (1994). Such a systempossesses the geometrical characteristics <strong>of</strong> a bilaminatewith diffuse boundaries between the layers. In the hingezones the micr<strong>of</strong>abric shows areas <strong>of</strong> distributed deformation.During fold development these zones <strong>of</strong> distributeddeformation (granular flow in the hinge zones) wouldcontribute to strain s<strong>of</strong>tening in this region leading tocontinuous amplification <strong>of</strong> the folds. We note that even ifthe deformation <strong>of</strong> the amphiboles is by brittle failure, itresults in distributed ductile flow in the highly deformedhinges. Unlike the folds in the garnet zone where probablyno slip on the limbs occurred, in the staurolite zone the slipis distributed through the relatively weak plagioclase richzones increasing the tendency for active amplification <strong>of</strong> thefold. In addition, on the limbs, because <strong>of</strong> the presence <strong>of</strong>relatively weak ‘layers’, fold amplification can be furtherassisted by flattening. This is well documented by the dipisogons pattern <strong>and</strong> the b 1 vs. F graph <strong>of</strong> Fig. 12.The sillimanite grade zone. In the sillimanite zone theamphibole <strong>and</strong> plagioclase show a relatively high aspectratio connected with a low degree <strong>of</strong> GBPO <strong>of</strong> like–like <strong>and</strong>unlike boundaries. In this zone, unlike the staurolite zonewhere the plagioclase was interconnected, isolated elongategrains or aggregates <strong>of</strong> plagioclase surrounded by highlyelongate <strong>and</strong> well-oriented crystals <strong>of</strong> amphibole occur(Figs. 3e <strong>and</strong> f <strong>and</strong> 8). This structure <strong>and</strong> the flattening <strong>of</strong>both minerals can be interpreted in terms <strong>of</strong> a stresssupportingnetwork with a low viscosity contrast betweenweaker plagioclase <strong>and</strong> stronger amphibole (H<strong>and</strong>y, 1990).Thus, in the sillimanite zone, it was the strength <strong>of</strong> ‘weaker’plagioclase that dominated the rheological properties <strong>of</strong> thesystem, which therefore acted as a relatively weak,homogeneous material.In contrast to the garnet zone, where buckling wascontrolled by localized micr<strong>of</strong>olding <strong>and</strong> to the staurolitezone where it was controlled by ductile shearing alongweak, plagioclase rich zones, no such zones <strong>of</strong> weakness,which facilitate the active amplification <strong>of</strong> the folds, areobserved in the sillimanite zone. Instead the fold shapeanalysis shows an importance <strong>of</strong> post-buckle flattening overactive fold amplification. The micro <strong>structural</strong> analysis <strong>of</strong>both the limb <strong>and</strong> hinge areas shows features consistent withhomogeneous flattening (i.e. higher aspect ratio <strong>and</strong> smallergrain size in the hinge domains than in the limbs) anobservation entirely consistent with passive amplification <strong>of</strong>a material with a low mechanical anisotropy.The contact metamorphic aureole. The deformation inthe contact aureole (Figs. 3g <strong>and</strong> h <strong>and</strong> 8) is an extremeexample <strong>of</strong> flattening dominated deformation as shown bydifferences in grain size <strong>and</strong> grain shapes in the hinge <strong>and</strong>limb areas. The lack <strong>of</strong> macroscopic folds in theamphibolites within the aureole is taken as further evidencethat the amplification was almost entirely passive.AcknowledgementsThe project was funded by grants <strong>of</strong> Czech NationalGrant Agency No. 42-201-204 to K.S. <strong>and</strong> 42-201-318 to P.Štípská, by Czech Geological Service assignment No. 6327to P. Mixa, <strong>and</strong> by a Ph.D. financial support attributed by theFrench Government to L.B. We thank R. Vernon <strong>and</strong> D.Grujic for constructive reviews <strong>and</strong> J. Hippertt for editorialhelp.Appendix AA polar graph aimed to represent the variation <strong>of</strong>orthogonal thickness t around folded layers was first utilizedby Lisle (1997). A fold can be represented by a series <strong>of</strong>points with polar coordinates (1/t 0 , a), where 1/t 0 is thereciprocal normalized thickness (t 0 Zt/t h , where t h is theextreme value <strong>of</strong> t, which is generally situated in the foldhinge), <strong>and</strong> a is the orientation <strong>of</strong> the layer tangent. Usingthis technique, Ramsay’s (1967) fold types give rise tovarious conic sections (ellipses, hyperbolas) in the polargraph with horizontal semiaxis equal to unity (see Lisle,1997).Flattening index F, or axial ratio <strong>of</strong> strain ellipse, expressthe amount <strong>of</strong> post-buckle flattening superimposed on theparallel fold. We used a <strong>numerical</strong>ly stable direct leastsquaremethod (Halír` <strong>and</strong> Flusser, 1998) to fit either ellipsesor hyperbolas onto points in a polar graph <strong>of</strong> normalizedthickness. This technique <strong>of</strong> evaluation <strong>of</strong> the flatteningindex F poses a robust estimate <strong>and</strong> is preferred in this work.The method <strong>of</strong> analysing fold shapes in terms <strong>of</strong> theharmonic coefficients <strong>of</strong> a Fourier series was originallydevised by Stabler (1968) <strong>and</strong> subsequently elaborated byHudleston (1973).The most basic <strong>and</strong> suitable segment <strong>of</strong> a folded surfacefor analysis is a ‘quarter-wavelength’ unit between adjacenthinge <strong>and</strong> inflexion points. Such a choice <strong>of</strong> unit leads to aharmonic series consisting only <strong>of</strong> the odd terms <strong>of</strong> a sine130


DTD 5ARTICLE IN PRESSL. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24 23series <strong>and</strong> the fold pr<strong>of</strong>ile is thus approximated by theequation: f ðxÞ Z XN ð2n K1Þpxb 2nK1 sin2LnZ1The first few harmonic coefficients are sensitiveparameters <strong>of</strong> fold shape <strong>and</strong> most information about thefold shape can be gained from the first two coefficients, b 1<strong>and</strong> b 3 .To obtain these coefficients, a ‘quarter-wavelength’ unit<strong>of</strong> length L <strong>of</strong> the fold is divided into equal sectors on a line,which is perpendicular to the axial surface <strong>and</strong> passesthrough the inflection point. Pairs <strong>of</strong> f(x n ) <strong>and</strong> x n coordinatesare measured at each <strong>of</strong> these points. A system <strong>of</strong> linearequations is assembled from pairs {x n , f(x n )} <strong>and</strong> solved forunknown coefficients, so2 sin px 1sin 3px 1/ sin ð2n K1Þpx 312L 2L2L2 3sin px 2sin 3px 2/ sin ð2n K1Þpx b 122L 2L2Lb 3« « 1 «6«74 56 sin px nsin3px nsinð2n K1Þpx 745nb 2nK122L3 2L2Lf ðx 1 Þf ðx 2 ÞZ64 «75f ðx n ÞReferencesArbaret, L., Mancktelow, N.S., Burg, J.-P., 2001. Effect <strong>of</strong> shape <strong>and</strong>orientation on rigid particle rotation <strong>and</strong> matrix deformation in simpleshear flow. Journal <strong>of</strong> Structural Geology 23, 113–125.Baratoux, L., 2004. Petrology, deformation mechanisms, <strong>and</strong> fabricanisotropy <strong>of</strong> metabasites deformed at natural strain <strong>and</strong> metamorphicgradient. Ph.D. thesis, Charles University, Prague.Berger, A., Stünitz, H., 1996. Deformation mechanisms <strong>and</strong> reaction <strong>of</strong>hornblende: examples from the Bergell tonalite (central Alps).Tectonophysics 257, 149–174.Biot, M.A., 1961. Theory <strong>of</strong> folding <strong>of</strong> stratified viscoelastic media <strong>and</strong> itsimplication in tectonics <strong>and</strong> orogenesis. Bulletin <strong>of</strong> the GeologicalSociety <strong>of</strong> America 72, 1595–1620.Biot, M.A., 1963a. Stability <strong>of</strong> multilayer continua including the effect <strong>of</strong>gravity <strong>and</strong> visco elasticity. Journal <strong>of</strong> Franklin Institute 276, 231–252.Biot, M.A., 1963b. Theory <strong>of</strong> stability <strong>of</strong> multilayered continua in finite,anisotropic elasticity. Journal <strong>of</strong> Franklin Institute 276, 128–153.Biot, M.A., 1967. Rheological stability with couple stresses <strong>and</strong> itsimplication to geological folding. Proceedings <strong>of</strong> the Royal Society <strong>of</strong>London, Series A 2298, 402–423.Brodie, K.H., Rutter, E.H., 1985. On the relationship between deformation<strong>and</strong> metamorphism, with special reference to the behavior <strong>of</strong> basicrocks. In: Thompson, A.B., Rubie, D.C. (Eds.), Metamorphic Reactions;Kinetics, Textures, <strong>and</strong> Deformation. Springer-Verlag, NewYork, pp. 138–179.Cháb, J., Žáček, V., 1994. Geology <strong>of</strong> the Žulová pluton mantle (Bohemianmassif, Central Europe). Věstník ÚÚG 69, 1–12.Chlupáč, I., 1994. Facies <strong>and</strong> biogeographic relationships in Devonian <strong>of</strong>the Bohemian massif. Corier Forschungsinstitut Senckenberg 169, 299–317.Cosgrove, J.W., 1976. The formation <strong>of</strong> crenulation cleavage. Journal <strong>of</strong> theGeological Society <strong>of</strong> London, 132(Part 2, 155–178.Dubey, A.K., 1980. Late stages in the development <strong>of</strong> folds as deducedfrom model experiments. Tectonophysics 65, 311–322.Dubey, A.K., Cobbold, P.R., 1977. Noncylindrical flexural slip folds innature <strong>and</strong> experiment. Tectonophysics 38, 223–239.Dudek, A., 1980. The crystalline basement block <strong>of</strong> the outer Carpathians inMoravia—Brunovistulicum. Rozpravy Československé Akademie věd,Rada matematika—prírodní vědy 90, 85pp.Fletcher, R.C., 1995. Three-dimensional folding <strong>and</strong> necking <strong>of</strong> a powerlawlayer: are folds cylindrical <strong>and</strong> do we underst<strong>and</strong> why?Tectonophysics 247, 65–83.Fowler, T.J., Winsor, C.N., 1996. Evolution <strong>of</strong> chevron folds by pr<strong>of</strong>ileshape changes; comparison between multilayer deformation experiments<strong>and</strong> folds <strong>of</strong> the Bendigo–Castlemaine goldfields, Australia.Tectonophysics 258, 125–150.Ghosh, S.K., 1968. Experiments <strong>of</strong> buckling <strong>of</strong> multilayers which permitinterlayer gliding. Tectonophysics 6, 207–249.Halír, R., Flusser, J., 1998. Numerically stable direct least squares fitting <strong>of</strong>ellipses. In: Proceedings <strong>of</strong> the 6th International Conference in CentralEurope on Computer Graphics <strong>and</strong> Visualization, WSCG ’98. CZ,Plzeň, pp. 125–132.H<strong>and</strong>y, M.R., 1990. The solid-state flow <strong>of</strong> polymineralic rocks. Journal <strong>of</strong>Geophysical Research, B 95, 8647–8661.H<strong>and</strong>y, M.R., 1994. Flow laws for rocks containing two non-linear viscousphases; a phenomenological approach. Journal <strong>of</strong> Structural Geology16, 287–301.Hobbs, B., 1971. The analysis <strong>of</strong> strain in folded layers. Tectonophysics 11,329–375.Hudleston, P.J., 1973. Fold morphology <strong>and</strong> some geometrical implications<strong>of</strong> theories <strong>of</strong> fold development. Tectonophysics 16, 1–46.Jehlička, J., 1995. The Žulová Massif in Silesian—its geochemistry <strong>and</strong>petrogenesis. PhD thesis, Charles University, Prague.Jordan, P.G., 1988. The rheology <strong>of</strong> polymineralic rocks—an approach.Geologische Rundschau 77, 285–294.Kretz, R., 1994. Metamorphic Crystallization. Wiley, New York.Lexa, O., 2003. Numerical approach in <strong>structural</strong> <strong>and</strong> micro<strong>structural</strong><strong>analyses</strong>. Ph.D. thesis, Charles University, Prague.Lisle, R.J., 1997. A fold classification scheme based on a polar plot <strong>of</strong>inverse layer thickness. In: Sengupta, S. (Ed.), Evolution <strong>of</strong> GeologicalStructures in Micro- to Macro-scales. Chapman & Hall, London,pp. 323–339.Maluski, H., Rajlich, P., Souček, J., 1995. Pre-Variscan, Variscan, <strong>and</strong>Early Alpine thermo-tectonic history <strong>of</strong> the north-eastern BohemianMassif: 40 Ar/ 39 Ar study. Geologishe Rundschau 84, 345–358.Mancktelow, N.S., Abbassi, M.R., 1992. Single layer buckle folding in nonlinearmaterials; II. Comparison between theory <strong>and</strong> experiment.Journal <strong>of</strong> Structural Geology 14, 105–120.Nyman, M.W., Law, R.D., Smelik, E.A., 1992. Cataclastic deformationmechanism for the development <strong>of</strong> core-mantle structures in amphibole.Geology 20, 455–458.Oertel, G., Ernst, W.G., 1978. Strain <strong>and</strong> rotation in a multilayered fold.Tectonophysics 48, 77–106.Panozzo, R.H., 1983. Two-dimensional analysis <strong>of</strong> shape-fabric usingprojections <strong>of</strong> digitized lines in a plane. Tectonophysics 95, 279–294.Panozzo, R.H., 1984. Two-dimensional strain from the orientation <strong>of</strong> linesin a plane. Journal <strong>of</strong> Structural Geology 6, 215–221.Passchier, C.W., Trouw, R.A.J., 1996. Microtectonics. Springer-Verlag,Berlin.Patočka, F., Valenta, J., 1990. Geochemistry <strong>of</strong> metatrachytes <strong>and</strong>metarhyolites <strong>of</strong> the southern part <strong>of</strong> the Devonian Vrbno Group inthe Horní Město area <strong>and</strong> tectonic setting <strong>of</strong> the origin <strong>of</strong> themetavolcanic protholith. Časopis pro mineralogii a geologii 35, 41–64 (in Czech).131


DTD 5ARTICLE IN PRESS24L. Baratoux et al. / Journal <strong>of</strong> Structural Geology xx (xxxx) 1–24Poirier, J.P., Guillopé, M., 1979. Deformation induced recrystallization <strong>of</strong>minerals. Bulletin de Minéralogie 102, 67–74.Pouba, Z., 1962. Vysvětlivky k prehledné geologické mapě ČSSR1:200,000, list M-33-XVIII, Jeseník (Legend <strong>of</strong> the geological map1:200,000, Jeseník). ČSAV, Prague.Price, N.J., Cosgrove, J.W., 1990. Analysis <strong>of</strong> Geological Structures.Cambridge University Press, Cambridge.Ramberg, H., 1961. Relationship between concentric longitudinal strain<strong>and</strong> concentric shearing strain during folding <strong>of</strong> homogeneous sheets <strong>of</strong>rock. American Journal <strong>of</strong> Science 259, 382–390.Ramberg, H., 1963. Fluid dynamics <strong>of</strong> viscous buckling applicable t<strong>of</strong>olding <strong>of</strong> layered rocks. Bulletin <strong>of</strong> the American Association <strong>of</strong>Petrologists <strong>and</strong> Geologists 47, 484–505.Ramberg, H., 1964. Selecting buckling <strong>of</strong> composite layers with contrastedrheological properties, a theory <strong>of</strong> simultaneous formation <strong>of</strong> severalorders <strong>of</strong> folds. Tectonophysics 1, 307–341.Ramsay, J.G., 1967. Folding <strong>and</strong> Fracturing <strong>of</strong> Rocks. McGraw-Hill, NewYork.Ramsay, J.G., Huber, M., 1987. The Techniques <strong>of</strong> Modern StructuralGeology. Volume 2: Folds <strong>and</strong> Fractures. Academic Press, London.Rosenberg, C.L., Stünitz, H., 2003. Deformation <strong>and</strong> recrystallization <strong>of</strong>plagioclase along a temperature gradient: an example from the Bergelltonalite. Journal <strong>of</strong> Structural Geology 25, 389–408.Roth, Z., 1962. Vysvětlivky k prehledné geologické mapě ČSSR 1:200,000,list M-33-XXIV, Olomouc (Legend <strong>of</strong> the geological map 1:200,000,Olomouc). ČSAV, Prague.Rozkošný, I., Souček, J., 1989. Contribution to the petrology <strong>of</strong> the Žulovámassif mantle. Acta Universitatis Carolinae, Geologica 2, 165–197.Schulmann, K., Gayer, R., 2000. A model for an obliquely developedcontinental accretionary wedge developed by oblique collision: the NEBohemian Massif. Journal <strong>of</strong> the Geological Society London 157, 401–416.Schulmann, K., Mlčoch, B., Melka, R., 1996. High-temperature microstructures<strong>and</strong> rheology <strong>of</strong> deformed granite, Erzgebirge, BohemianMassif. Journal <strong>of</strong> Structural Geology 18, 719–733.Souček, J., 1978. Metamorphic zones <strong>of</strong> the Vrbno <strong>and</strong> Rejvíz Series, theHrubý Jeseník Mts, Czechoslovakia. Tschermaks MineralogischePetrographische Mitteilungen 25, 195–217.Stabler, C.R., 1968. Simplified Fourier analysis <strong>of</strong> fold shapes. Tectonophysics6, 343–350.Štípská, P., Schulmann, K., Hoeck, V., 2000. Complex metamorphiczonation <strong>of</strong> the Thaya Dome; result <strong>of</strong> buckling <strong>and</strong> gravitationalcollapse <strong>of</strong> an imbricated nappe sequence. In: Cosgrove, J.W.,Ameen, M.S. (Eds.), Forced Folds <strong>and</strong> Fractures Geological SocietySpecial Publication, 169, pp. 197–211.Štípská, P., Schulmann, K., Kröner, A., 2004. Vertical extrusion <strong>and</strong> middlecrustal spreading <strong>of</strong> omphacite granulite: a model <strong>of</strong> syn-convergentexhumation (Bohemian Massif, Czech Republic). Journal <strong>of</strong> MetamorphicGeology 22, 179–198.Stünitz, H., 1993. Transition from fracturing to viscous flow in a naturallydeformed metagabbro. In: Bol<strong>and</strong>, J.N., Fitz Gerald, J.D. (Eds.),Defects <strong>and</strong> Processes in the Solid State; Geoscience Applications; theMcLaren Volume Developments in Petrology, 14. Elsevier, Amsterdam-NewYork, International, pp. 121–150.Vernon, R.H., 1976. Metamorphic Processes. Murby, London; Wiley, NewYork.Yund, R., Tullis, J., 1991. Compositional changes <strong>of</strong> minerals associatedwith dynamic recrystallization. Contributions to Mineralogy <strong>and</strong>Petrology 108, 346–355.132


J. metamorphic Geol., 2008, 26, 273–297 doi:10.1111/j.1525-1314.2007.00755.xVertical extrusion <strong>and</strong> horizontal channel flow <strong>of</strong> orogenic lowercrust: key exhumation mechanisms in large hot orogens?K. SCHULMANN, 1 O. LEXA, 1,3 P. ŠTÍPSKÁ, 1 M. RACEK, 2 L. TAJČMANOVÁ, 2,4 J. KONOPÁSEK, 2,3J.-B. EDEL, 1 A. PESCHLER 1 AND J. LEHMANN 11 Université Louis Pasteur, EOST, UMR 7516 - 7517, 1 Rue Blessig, Strasbourg 67 084, France2 Czech Geological Survey, Klárov 3, 118 21 Praha 1, Czech Republic3 Institute <strong>of</strong> Petrology <strong>and</strong> Structural geology, Charles University, Albertov 6, 128 43, Prague, Czech Republic4 Dipartimento di Mineralogia e Petrologia, Università di Padova, Corso Garibaldi 37, I-35 137 Padova, ItalyABSTRACTA large database <strong>of</strong> <strong>structural</strong>, geochronological <strong>and</strong> petrological data combined with a Bougueranomaly map is used to develop a two-stage exhumation model <strong>of</strong> deep-seated rocks in the eastern sector<strong>of</strong> the Variscan belt. An early sub-vertical fabric developed in the orogenic lower <strong>and</strong> middle crustduring intracrustal folding followed by the vertical extrusion <strong>of</strong> the lower crustal rocks. These eventswere responsible for exhumation <strong>of</strong> the orogenic lower crust from depths equivalent to 18)20 kbar todepths equivalent to 8)10 kbar, <strong>and</strong> for coeval burial <strong>of</strong> upper crustal rocks to depths equivalent to 8–9 kbar. Following the folding <strong>and</strong> vertical extrusion event, sub-horizontal fabrics developed at mediumto low pressure in the orogenic lower <strong>and</strong> middle crust during vertical shortening. Fabrics that record theearly vertical extrusion originated between 350 <strong>and</strong> 340 Ma, during building <strong>of</strong> an orogenic root inresponse to SE-directed Saxothuringian continental subduction. Fabrics that record the later subhorizontalexhumation event relate to an eastern promontory <strong>of</strong> the Brunia continent indenting into therheologically weaker rocks <strong>of</strong> the orogenic root. Indentation initiated thrusting or flow <strong>of</strong> the orogeniccrust over the Brunia continent in a north-directed sub-horizontal channel. This sub-horizontal flowoperated between 330 <strong>and</strong> 325 Ma, <strong>and</strong> was responsible for a heterogeneous mixing <strong>of</strong> blocks <strong>and</strong>boudins <strong>of</strong> lower <strong>and</strong> middle crustal rocks <strong>and</strong> for their progressive thermal re-equilibration. Theerosion depth as well as the degree <strong>of</strong> reworking decreases from south to north, pointing to an outflow <strong>of</strong>lower crustal material to the surface, which was subsequently eroded <strong>and</strong> deposited in a forel<strong>and</strong> basin.Indentation by the Brunia continental promontory was highly noncoaxial with respect to the SEorientedSaxothuringian continental subduction in the Early Visean, suggesting a major switch <strong>of</strong> plateconfiguration during the Middle to Late Visean.Key words: Bohemian Massif; channel flow; exhumation; orogenic lower crust; Variscan belt.INTRODUCTIONCurrent concepts <strong>of</strong> exhumation <strong>of</strong> deep-seated rocksin convergent orogens are generally based on the style<strong>of</strong> the pressure–temperature–time (P–T–t) pathretrieved from high-pressure (HP) to ultra-highpressure(UHP) rocks (e.g. Duchene et al., 1997). Onegroup <strong>of</strong> exhumation mechanisms for these rocks hasbeen inferred from conceptual or <strong>numerical</strong> modelsdriven by subduction–accretion processes that result ineither corner flow circulation within an accretionarywedge (Platt, 1986, 1993; Allem<strong>and</strong> & Lardeaux, 1997;Gerya & Stockhert, 2006) or buoyancy-driven exhumation<strong>of</strong> subducted continental crust (Chemendaet al., 1995). Another group <strong>of</strong> conceptual models hasbeen developed for gravity-driven exhumation <strong>of</strong> HProcks in thickened orogenic root systems. In thesemodels, processes such as convective removal <strong>of</strong> atectospheric root (Engl<strong>and</strong> & Houseman, 1988;Andersen et al., 1991) or lateral variations in gravitationalpotential energy <strong>of</strong> thickened continentallithosphere (Milnes & Koyi, 2000; Rey et al., 2001;V<strong>and</strong>erhaeghe & Teyssier, 2001) drive the exhumation.Recently, an alternative model has been developed forthe large-scale horizontal movement <strong>of</strong> melt-bearingmiddle crust based on channel flow (Beaumont et al.,2001, 2006; Godin et al., 2006), <strong>and</strong> has been appliedto explain the ductile extrusion <strong>of</strong> medium-pressuremetamorphic rocks along the Himalayan front (Grujicet al., 1996; Jamieson et al., 2002, 2004). Whereas thefirst group <strong>of</strong> models focuses on explaining verticaldisplacements <strong>of</strong> UHP <strong>and</strong> HP rocks, the second group<strong>of</strong> models emphasizes the importance <strong>of</strong> large-scalehorizontal displacements in orogens.Monocyclic continuous models <strong>of</strong> orogenic lowercrust exhumation may be supported by <strong>structural</strong> <strong>and</strong>kinematic field studies that emphasize the 2D character<strong>of</strong> the exhumation process (e.g. Milnes & Koyi, 2000).However, more commonly there are complexities inthe 3D character <strong>of</strong> the tectonic evolution <strong>of</strong> largeÓ 2007 Blackwell Publishing Ltd 273133


274 K. SCHULMANN ET AL.orogenic root systems that arise from the polyphasenature <strong>of</strong> vertical <strong>and</strong> horizontal material <strong>and</strong> heattransfer during long-lasting orogenic events that makethese systems more difficult to unravel. The polyphase<strong>and</strong> discontinuous character <strong>of</strong> orogeny may be due tomajor changes in plate configurations (e.g. Deweyet al., 1989) or to the existence <strong>of</strong> inherited rheologicalheterogeneities <strong>and</strong> variations in mechanical anisotropy(e.g. Burg, 1999). Therefore, detailed regional<strong>structural</strong>, petrological <strong>and</strong> geochronological studies<strong>of</strong> orogenic fabrics may provide a key to decipheringthe succession, <strong>and</strong> length <strong>and</strong> time scales <strong>of</strong> processesresponsible for material <strong>and</strong> heat transfer within theselarge orogenic root systems.Classically, the Palaeozoic Variscan orogen inWestern <strong>and</strong> Central Europe (Fig. 1) is a large, hot,bivergent orogen that is interpreted to have developed50S.ArmoricanF.BrayF.Cadomian0TBBav. F.ElbeF.hercynianRhenohercynianthuringianSaxothuringianMoldanubian14 1Bruni runia48-40N50 kmMOLDANUBIANMonotonous <strong>and</strong>Varied GroupGföhl unitTEPLA-BARRANDIANProterozoicEarly Palaeozoic4WEST SUDETES52 oLUGIANPRAHA50 o 48 o12 oSAXOTHURINGIANduring prolonged convergence between Laurussia <strong>and</strong>Gondwana (e.g. Ziegler, 1986; Matte et al., 1990).Traditionally, burial <strong>and</strong> exhumation <strong>of</strong> UHP <strong>and</strong> HProcks are thought to be the result <strong>of</strong> a kinematic continuum<strong>of</strong> coaxial subduction <strong>and</strong> subsequent collisionprocesses (e.g. OÕBrien & Carswell, 1993; Konopa´ sek& Schulmann, 2005), <strong>and</strong> as a consequence petrological<strong>and</strong> geochronological data may fit in one <strong>of</strong> the2D conceptual models discussed above.In this study, we present <strong>structural</strong>, petrological <strong>and</strong>geochronological data acquired during the last twodecades from the eastern termination <strong>of</strong> the EuropeanVariscan front within an area <strong>of</strong> about 20 000 km 2(Figs 1 & 2). It is shown that exhumation <strong>of</strong> the orogeniclower crust occurred during two distinct periodsrelated to a major change in the configuration <strong>of</strong>lithospheric plates during the Visean. The first exhu-TEPLA-BARRANDIANELBE -ZONESaxothuringiansubductionSouth Central Bohemian Bohemian pluton plutonC retaceousBasinMOLDANUBIAN DOMAINMORAVO-SILESIANBRUNIA14 o 16 o BrnoFig. 1. Outline geological map <strong>of</strong> the Bohemian Massif with major units shown schematically (modified after Franke, 2000).Upper inset is the position <strong>of</strong> the Bohemian Massif in the framework <strong>of</strong> the European Variscides (after Edel et al., 2003). The areadiscussed in this review is marked by a rectangular box.Ó 2007 Blackwell Publishing Ltd134


EXHUMATION IN LARGE HOT OROGEN 275mation event was related to almost vertical materialtransfer, which was driven by S-E directed oceanic <strong>and</strong>continental (Armorican–Saxothuringian) subductiondynamics during the Devonian <strong>and</strong> Tournasian toEarly Visean (Fig. 1). Following this early evolution,during the Middle to Late Visean indentation from theeast by a promontory <strong>of</strong> the Brunia continent at a highangle to the Saxothuringian subduction directionresulted in horizontal flow <strong>and</strong> transport <strong>of</strong> theextruded orogenic lower crust over the rigid continentas a hot fold nappe.GEOTECTONIC SETTINGSuess (1912, 1926) described the geology <strong>of</strong> the easternVariscan front <strong>and</strong> divided the crystalline complexes <strong>of</strong>the Bohemian Massif into two parts, an internalMoldanubian–Lugian domain <strong>and</strong> an external Moravo–SilesianZone (Figs 1 & 2). Dudek (1980) completedthis subdivision <strong>and</strong> defined a Brunia continentwith the Moravo–Silesian Zone as its westerndeformed margin. The eastern segment <strong>of</strong> the Bruniacontinent is built up <strong>of</strong> unmetamorphosed to weaklymetamorphosed Neoproterozoic granites, high-gradeschists <strong>and</strong> migmatites. This basement is unconformable,covered by Lower Carboniferous forel<strong>and</strong> basinsedimentary rocks, by Devonian shallow marine sedimentaryrocks, <strong>and</strong> locally by Cambro-Ordovicianclastic, pelitic metasediment rocks <strong>and</strong> bimodal metavolcanicrocks (Franke, 2000; Hartley & Otava, 2001).The Moravo–Silesian Zone represents a NE–SWtrendingbelt <strong>of</strong> sheared <strong>and</strong> metamorphosed rocksderived from the Brunia continent. This 300 km long,30)50 km wide belt consists <strong>of</strong> three NE–SW-elongatedtectonic windows emerging through <strong>structural</strong>lyoverlying high-grade rocks <strong>of</strong> the Moldanubian–Lugian domain: a southern Thaya window; a centralSvratka window; <strong>and</strong> a northern Silesian domain(Fig. 2). Schulmann et al. (2005) identified the Moldanubian–Lugi<strong>and</strong>omain in the Bohemian Massif asthe deep orogenic root system <strong>of</strong> the Variscan orogen(Fig. 2). The Elbe zone (Fig. 2) divides rocks <strong>of</strong> theMoldanubian–Lugian domain into two: a larger highgradeMoldanubian domain to the south <strong>and</strong> a smallerhigh-grade Lugian domain to the north (Suess, 1926).The Moldanubian domain (Figs 1 & 2) has beensubdivided into three major lithotectonic units (e.g.Fuchs, 1986), namely the amphibolite facies Monotonous<strong>and</strong> Varied Groups, which together with gneissesmake up the orogenic middle crust, <strong>and</strong> the predominantlygranulite facies Gfo¨ hl Unit, which is inferred tobe the orogenic lower crust (Fig. 2). The Varied Groupoutcrops <strong>structural</strong>ly above the Monotonous Group,with a contact that is commonly marked by bodies <strong>of</strong>granite gneiss. The Monotonous Group consists <strong>of</strong>migmatitic paragneisses (metagraywacke) interlayeredwith granite orthogneisses, quartzites <strong>and</strong> rare eclogites(Dudek & Fediukova´ , 1974; Petrakakis, 1997). TheVaried Group consists <strong>of</strong> a thick sequence <strong>of</strong> paragneissesinterlayered with calcsilicate rocks, marbles,quartzites, graphite schists, amphibolites <strong>and</strong> felsicmetavolcanic rocks. The Gfo¨ hl Unit is composed <strong>of</strong>large areas <strong>of</strong> migmatitic granite gneiss, called theGfo¨ hl gneiss, <strong>and</strong> <strong>of</strong> areas <strong>of</strong> various highly anatecticmigmatites <strong>and</strong> paragneisses that are in placesaccompanied by migmatitic amphibolites at the base.The Gfo¨ hl Unit includes numerous bodies <strong>of</strong> Ky–Kfsfelsic granulite as well as tectonic lenses <strong>of</strong> eclogite <strong>and</strong>garnet <strong>and</strong> ⁄ or spinel peridotite.The main part <strong>of</strong> the Lugian domain is composed <strong>of</strong>medium- to high-grade granite gneisses <strong>and</strong> metamorphosedvolcano-sedimentary rocks with Cambro–Ordovician protolith ages (Kro¨ ner et al., 2001, <strong>and</strong>references therein). The granite gneisses contain boudins<strong>of</strong> eclogite <strong>and</strong> a belt <strong>of</strong> garnet–omphacite granulite,<strong>and</strong> are considered an equivalent <strong>of</strong> the Gfo¨ hlUnit, whereas the belts <strong>of</strong> medium-grade schists <strong>of</strong>volcano-sedimentary origin are regarded as an equivalent<strong>of</strong> either the Monotonous or the Varied Groups(Fig. 2). An Ordovician leptyno-amphibolite lowercrustal complex (the Stare´ Město belt) occurs in theeastern part <strong>of</strong> the Lugian domain (Sˇtı´ pska´ et al.,2001).DEFINITION OF BASEMENT AND OROGENICCRUSTAL LEVELSIn this section, multiple criteria, such as the peakpressure conditions attained by individual units, thecharacter <strong>and</strong> the age <strong>of</strong> the protolith <strong>and</strong> chronology<strong>of</strong> metamorphic zircon, are used to decipher the relativevertical position <strong>of</strong> crustal units during the LowerPalaeozoic <strong>and</strong> the Devonian. This information is thenused to propose a model for the stratification <strong>of</strong> theorogenic crust along the eastern Variscan front.The Brunia continentBased on zircon protolith ages (Fig. 3a) <strong>and</strong> 40 Ar ⁄ 39 Arcooling ages ranging from 600 to 540 Ma, granites <strong>and</strong>migmatites forming the basement <strong>of</strong> the Brunia continentare inferred to have originated during the Pan-African orogenic events (Van Breemen et al., 1982;Fritz et al., 1996; Finger et al., 2000; Friedl et al.,2004). Tectonic imbrication <strong>of</strong> Devonian metasedimentaryrocks <strong>and</strong> basement, <strong>and</strong> metamorphism tolower greenschist facies occurred at the western margin<strong>of</strong> the Brunia continent during the Carboniferous(Francu˚ et al., 2002). In the northern Silesian domain(Fig. 2), the metamorphic pattern <strong>of</strong> nappes derivedfrom the Brunia continent shows an inverted Barrovianzonation ranging from chlorite grade in the east tokyanite ⁄ sillimanite grade in the west (Sˇtı´ pska´ &Schulmann, 1995; Schulmann & Gayer, 2000). Thewestern termination <strong>of</strong> the Moravo–Silesian Zone inthe north reached eclogite facies conditions (Fig. 4;Sˇtı´ pska´ et al., 2006), <strong>and</strong> it is likely that the western tip<strong>of</strong> the Moravo–Silesian Zone in the south has beenÓ 2007 Blackwell Publishing Ltd135


276 K. SCHULMANN ET AL.Moldanubian-LugianOROGENIC ROOTOrogenic upper crustProterozoic <strong>and</strong>Paleozoic sedimentsOrogenic middle crustMonotonous groupVaried groupGneissesOrogenic lower crust(Gföhl Unit in Moldanubian)Granulites, gneissesMetasedimentsOrdovician mafic lower crustLUGIANBELBE ZONEC′Fig. 7SilesianB’ B′MOLDANUBIAN0 20 40 60MAGMATIC ROCKSSyenitesTonalitesCMPAS outhernC entralNorthernCSvratk vratka aThay haya aA′Moravo-SilesianBRUNIABRUNIAFig. 6Fig. 5Forel<strong>and</strong> basinCulmPara-autochtonous unitsCoverBasementAlochtonous unitsCoverBasementMoravianMicaschist ZonePermian coverLate granitesLetovice ophioliteFig. 2. Map <strong>of</strong> the eastern margin <strong>of</strong> the Bohemian Massif to show inferred orogenic crustal levels <strong>and</strong> the locations <strong>of</strong> Figs 5, 6 & 7,which are indicated by rectangular boxes. Also shown are the positions <strong>of</strong> interpretative cross sections A–A¢, B–B¢ <strong>and</strong> C–C¢ presentedin Fig. 10.Ó 2007 Blackwell Publishing Ltd136


Probabi l irequencyProbabi l irequencyEXHUMATION IN LARGE HOT OROGEN 277underthrust to eclogite conditions as well (Konopa´ seket al., 2002).The orogenic rootThe Gföhl Unit <strong>and</strong> Lugian high-grade rocksThe Gfo¨ hl Unit (Fig. 2) is composed predominantly <strong>of</strong>felsic granulites (11–20 kbar, 800–1000 °C) <strong>and</strong> highgradegneisses (e.g. OÕBrien & Ro¨ tzler, 2003; Sˇtı´ pska´ &Powell, 2005b) that contain eclogite lenses <strong>of</strong> basalticcomposition, suggesting a crustal origin (18–19 kbar,800)900 °C, Fig. 4; e.g. Medaris et al., 1998; Sˇtípska´& Powell, 2005a <strong>and</strong> references therein). The eclogites<strong>and</strong> granulites, as well as other lithologies <strong>of</strong> the Gfo¨ hlUnit, were overprinted at amphibolite facies conditions(4)12 kbar, 750)850 °C, Fig. 4; e.g. Petrakakis,1997; Hasalova´ et al., 2008b). The Gfo¨ hl Unit containsnumerous tectonically emplaced bodies <strong>of</strong> garnet <strong>and</strong>spinel peridotite (28–44 kbar, 900–1300 °C, Fig. 4),associated with lenses <strong>of</strong> eclogite (16–20 kbar, 800–950 °C) <strong>and</strong> clinopyroxenite, representing partial meltcrystallized in the upper mantle (Medaris et al., 1995).The Lugian eclogites <strong>and</strong> granulites (18–20 kbar, 800–900 °C, Fig. 4; e.g. Steltenpohl et al., 1993; Sˇtípska´et al., 2004) are considered to reflect UHP conditions(Kryza et al., 1996), but this has not been confirmed;they were re-equilibrated at variable P–T conditions(4)11 kbar, 700)800 °C, Fig. 4).Multigrain zircon <strong>and</strong> monazite fractions fromgranulites <strong>and</strong> high-grade gneisses <strong>of</strong> the Gfo¨ hl Unityielded conventional upper intercept U–Pb agesbetween 550 <strong>and</strong> 510 Ma, interpreted to date the formation<strong>of</strong> the protoliths (Fig. 3b; e.g. Van Breemenet al., 1982; Schulmann et al., 2005). In addition,xenocrystic cores <strong>and</strong> prismatic zircon from felsicgranulites <strong>and</strong> Gfo¨ hl gneiss yield Silurian to DevonianU–Pb ages also interpreted to date the formation <strong>of</strong> theprotoliths (Fig. 3b; Friedl et al., 2004). However, amajority <strong>of</strong> U–Pb zircon ages fall in the range360)340 Ma with a prominent peak at c. 340 Ma,which is interpreted to record the age <strong>of</strong> Variscanmetamorphism (Fig. 3d). Ultra-potassic syenites (durbachites)spatially related to Ky–Kfs granulites yieldages around 338 Ma <strong>and</strong> 325 Ma (Fig. 3e; Holubet al., 1997; Janousˇek & Holub, 2007).The Monotonous <strong>and</strong> Varied GroupsIn the Moldanubian domain, paragneisses <strong>of</strong> theMonotonous <strong>and</strong> Varied Groups generally recordmedium-pressure metamorphism (8)9 kbar, 610)660 °C) associated at higher temperatures with widespreadanatexis ( ~ 9 kbar, 700)800 °C), followed byre-equilibration in the sillimanite stability field atconditions around 4)6 kbar <strong>and</strong> 600)800 °C (Fig. 4;e.g. Petrakakis, 1997). The metamorphic conditions<strong>of</strong> the Monotonous Group paragneisses aret y fBrunia <strong>and</strong> Moravo-Silesian0.7(a)503 Ma0.6580 Ma0.50.4684 Ma0.30.2Protolith0.1U–Pb0350 400 450 500 550 600 650 700 7504(b)3.532.521.510.5Orogenic middle crust503 MaLugian (60)Moldanubian (5)ProtolithU–Pb0350 400 450 500 550 600 650 700 750Orogenic lower crust0.7(c)0.6413 Ma0.5389 Ma0.4550 Ma0.3487 Ma430 Ma0.20.1Moldanubian (14)Lugian (9)ProtolithU–Pb0350 400 450 500 550 600 650 700 750t y f3.5(d)3 340 Ma2.520.510.5359 MaMoldanubian (22)Lugian (15)1.4(e)1.2MetamorphicMagma.0.2U–PbU–Pb00300 320 340 360 380 400 420 280 300 320 340 360 380 40010.80.60.4335 Ma324 Ma340 Ma353 MaFig. 3. Probability curves <strong>of</strong> age histograms to show existing zircon U–Pb ages. For the Brunia continent <strong>and</strong> the Moravo–SilesianZone (a), Moldanubian <strong>and</strong> Lugian domain (b, c & d) <strong>and</strong> intrusive rocks (e). BB1, Brunia basement south <strong>of</strong> the Elbe zone; BB2,Brunia basement north <strong>of</strong> the Elbe zone; MB1, Moravian domain basement rocks; MB2, Silesian domain basement rocks; MSZ,Moravian Micaschist Zone; ID, Mg-rich syenites (durbachites) <strong>of</strong> the orogenic lower crust; IMC, intrusive rocks <strong>of</strong> the middle crust;IT, tonalite–granodiorite intrusions; IUC, granites in upper crustal units. Numbers in parentheses st<strong>and</strong> for number <strong>of</strong> available ages.Ó 2007 Blackwell Publishing Ltd137


278 K. SCHULMANN ET AL.Fig. 4. P–T conditions for (a) peak metamorphism in the orogenic root <strong>and</strong> Brunia continent south <strong>of</strong> the Elbe Zone; (b) peakmetamorphism in the orogenic root <strong>and</strong> Brunia continent north <strong>of</strong> the Elbe Zone; (c) retrograde conditions in the orogenic root<strong>and</strong> Brunia continent south <strong>of</strong> the Elbe Zone; <strong>and</strong> (d) retrograde conditions in the orogenic root <strong>and</strong> Brunia continent north <strong>of</strong> theElbe Zone. Source <strong>of</strong> data: Carswell & O¢Brien (1993), Steltenpohl et al. (1993), Medaris et al. (1995), Sˇtı´ pska´ & Schulmann (1995),Bro¨ cker & Klemd (1996), Kryza et al. (1996), Pitra & Guiraud (1996), Parry et al. (1997), Petrakakis (1997), Medaris et al. (1998),Klemd & Bro¨ cker (1999); Romanova´ &Sˇtı´ pska´ (2001); Konopa´ sek et al. (2002); OÕBrien & Ro¨ tzler (2003); Sˇtı´ pska´ et al. (2004); Lexaet al. (2005); Sˇtı´ pska´ & Powell (2005a,b); Racek et al. (2006); Sˇtı´ pska´ et al. (2006); Tajcˇmanová et al. (2006); also shown are steadystategeotherms calculated for various mantle heat flows (as indicated in the figure) <strong>and</strong> exponential radioactive heat production.commonly re-equilibrated in the stability field <strong>of</strong>cordierite at conditions around 4.5–6 kbar <strong>and</strong> 600–720 °C (Petrakakis, 1997), inferred to be due to thethermal effect <strong>of</strong> Carboniferous intrusions emplacedat 330)310 Ma (Fig. 3e). Despite the overall medium-pressurecharacter, in the Moldanubian domainthe Varied <strong>and</strong> Monotonous Group rocks includerare lenses <strong>of</strong> eclogite that register conditions <strong>of</strong> 13–16 kbar at 600–680 °C (Fig. 4; Medaris et al., 1995).The peak metamorphism <strong>of</strong> the volcano-sedimentaryrock series in the Lugian domain varies from garnetto kyanite grade, yielding 7)9 kbar at 550)650 °C(Romanova´ & Sˇtı´ pska´ , 2001; Jastrzębski, 2005;Murtezi, 2006) with re-equilibration in the sillimanitestability field at sub-solidus conditions <strong>of</strong> 5–6 kbar ataround 600 °C.In the Monotonous <strong>and</strong> Varied Groups <strong>of</strong> theMoldanubian domain, scarce U)Pb zircon data fromorthogneisses <strong>and</strong> metavolcanic rocks yield Neoproterozoic<strong>and</strong> Cambrian protolith ages (Fig. 3c;Schulmann et al., 2005). U)Pb zircon data frommetavolcanic rocks <strong>and</strong> granites <strong>of</strong> the Lugian domainÓ 2007 Blackwell Publishing Ltd138


EXHUMATION IN LARGE HOT OROGEN 279yield predominantly Cambrian protolith ages (Fig. 3c;e.g. Kro¨ ner et al., 2000 <strong>and</strong> references therein).Neoproterozoic <strong>and</strong> Lower Palaeozoic low-grade sequencesA sequence <strong>of</strong> weakly metamorphosed Neoproterozoic<strong>and</strong> Silurian to Devonian sedimentary rocks is separatedby a crustal-scale detachment from the underlyingmedium- <strong>and</strong> high-grade rocks <strong>of</strong> the Moldanubian<strong>and</strong> Lugian domains (Pitra et al., 1994; Mazur et al.,2005). The NE–SW-trending contact <strong>of</strong> the uppercrustal sedimentary rocks <strong>and</strong> middle crustal metamorphicrocks is marked by granodiorite sills, whichyield U–Pb zircon ages <strong>of</strong> 350–340 Ma (Fig. 3e).Intrusion <strong>of</strong> the sills was accompanied by metamorphism<strong>of</strong> the hangingwall rocks at conditions around4 kbar <strong>and</strong> 550)590 °C (Fig. 4). Locally, granite plutons,which yield U–Pb zircon ages around 330 Ma,intrude these low-grade rocks (e.g. Schulmann et al.,2005).define these units as representing the orogenic middlecrust. The Monotonous <strong>and</strong> Varied Groups <strong>and</strong> lowgradeunits yield only Neoproterozoic <strong>and</strong> Ordovicianprotolith ages, they lack a Devonian thermal reworking<strong>and</strong> they are interpreted as forming shallowercrustal levels during the Devonian <strong>and</strong> Carboniferous.Using the same reasoning, the weakly metamorphosedNeoproterozoic <strong>and</strong> Lower Palaeozoic rocks that showpeak pressures around 4 kbar, <strong>and</strong> no evidence <strong>of</strong>Devonian thermal reworking are defined as the orogenicupper crust.During the Carboniferous history <strong>of</strong> building theorogenic root, the rocks occupying <strong>structural</strong>ly deeperpositions were transported upwards, whereas shallowerunits were moved downwards. The result is thatthe orogenic lower <strong>and</strong> middle crusts now form subparallelbelts at the surface. Exceptionally, the Varied<strong>and</strong> Monotonous Groups reached lower orogeniccrustal conditions during their burial, <strong>and</strong> in this case,they become part <strong>of</strong> the orogenic lower crust.Definition <strong>of</strong> crustal levels within the Moldanubian–Lugianorogenic rootThe lithologies <strong>and</strong> metamorphic conditions <strong>of</strong> themain units <strong>of</strong> the Moldanubian <strong>and</strong> Lugian domainsare used to define a lithotectonic zonation that correspondsto different crustal levels that existed duringmaximum thickening <strong>of</strong> the orogenic root. In turn, thepresent map distribution <strong>of</strong> the different crustal levelsdefines the spatial pattern <strong>of</strong> crustal units that experienceddifferent vertical displacements. Interpretation<strong>of</strong> these different crustal levels associated with thickening<strong>and</strong> the differences in vertical displacement, togetherwith the geometry <strong>and</strong> <strong>structural</strong> record withineach unit may allow us to unravel the mechanism(s) <strong>of</strong>exhumation <strong>of</strong> these units.According to the lithologies <strong>and</strong> peak P–T conditionsdescribed above, the Gfo¨ hl Unit is a granulite–eclogite unit representing the lower part <strong>of</strong> a thickenedcrust that reached lithostatic pressures <strong>of</strong> 18–20 kbar,corresponding to a depth <strong>of</strong> 60–70 km, which wedefine as the orogenic lower crust. In addition, thegeochronological <strong>and</strong> petrological data show a systematiccorrelation <strong>of</strong> protolith age, metamorphic age<strong>and</strong> peak pressures (Figs 3 & 4). Only the high-graderocks <strong>of</strong> the Gfo¨ hl Unit <strong>and</strong> the Lugian granulitesrecord Devonian protolith ages, Early Carboniferousmetamorphic ages <strong>and</strong> peak pressures <strong>of</strong> 18–20 kbar.For these reasons, Schulmann et al. (2005) suggestedthat the orogenic lower crust in the Bohemian Massifrepresents Neoproterozoic–Cambro-Ordovician continentalcrust that experienced a major thermalreworking during Devonian intra-continental (backarc)rifting.The Monotonous <strong>and</strong> Varied Groups yield peakpressure conditions not exceeding 12 kbar <strong>and</strong> areinterpreted to have reached maximum crustal depthsaround 40 km under an elevated thermal gradient. WeTECTONICSOFTHEEASTERNPARTOFTHEBOHEMIAN MASSIFStructural evolution <strong>of</strong> the Brunia continental margin: theMoravo–Silesian ZoneThe Brunia continent forms a pile <strong>of</strong> internallyimbricated basement- <strong>and</strong> cover-derived thrust sheets,termed the Moravo–Silesian Zone, which developedduring Carboniferous dextral–oblique thrusting <strong>of</strong> theMoldanubian–Lugian domain (Schulmann et al.,1991). As a result, the whole Moravo–Silesian Zoneexperienced intense non-coaxial NE-directed deformation(Schulmann et al., 1994; Fritz et al., 1996).Crustal-scale folds with west-plunging hinges refoldthe nappe pile during the final stages <strong>of</strong> NE-directeddeformation (Schulmann, 1990; Sˇtı´ pska´ et al., 2000).Finally, the Svratka dome <strong>and</strong> the Silesian domainwere folded by west-facing folds that also affected theCulm forel<strong>and</strong> basin, suggesting the age <strong>of</strong> latestshortening to be around 310 Ma (Hartley & Otava,2001).Orogenic structure <strong>of</strong> the crustal rootThe eastern branch <strong>of</strong> the Moldanubian domainlocated between Central Moldanubian pluton <strong>and</strong>Moravo–Silesian Zone (Fig. 1) is examined here. Inorder to characterize <strong>structural</strong> <strong>and</strong> petrological variationsalong-strike <strong>of</strong> the orogenic margin we introducethe southern, central <strong>and</strong> northern Moldanubi<strong>and</strong>omains.The large-scale structure <strong>of</strong> the Moldanubi<strong>and</strong>omain is characterized from west to east as follows:(i) first, there is a sequence <strong>of</strong> east-dipping schists <strong>and</strong>marbles <strong>of</strong> the Monotonous <strong>and</strong> Varied groups, whichis intruded by the Central Moldanubian pluton; (ii) tothe east, this sequence is overlain by high-gradeÓ 2007 Blackwell Publishing Ltd139


280 K. SCHULMANN ET AL.gneisses, migmatites <strong>and</strong> granulites containingnumerous inclusions <strong>of</strong> peridotites <strong>and</strong> eclogites; (iii)further east, this orogenic lower crust is thrust overwest-dipping middle crustal rocks that are composed<strong>of</strong> sillimanite micaschists, orthogneiss or leptyniticamphibolite <strong>of</strong> Cambro-Ordovician age in the south(Figs 5c & 6c); <strong>and</strong> (iv) finally, the easternmostextremity <strong>of</strong> the Moldanubian sequence is characterizedby a lower crustal segment <strong>of</strong> mafic <strong>and</strong> felsicgranulites associated with migmatitic biotite paragneiss<strong>and</strong> a few mantle slivers (Fig. 5c). All these unitstrend parallel to the continental margin represented byMoravian nappes <strong>of</strong> the Thaya <strong>and</strong> Svratka domes,respectively (Fig. 2). This sequence <strong>of</strong> rocks is generallyvalid for the northern <strong>and</strong> southern parts <strong>of</strong> theMoldanubian domain, but in contrast the central partis dominated by orogenic lower crust rocks intruded byMg–K-rich syenite intrusions (figs 4 & 5 in Schulmannet al., 2005).The geological structure <strong>of</strong> the Lugian domain(Fig. 7) is characterized by alternations <strong>of</strong> orthogneisseswith metasedimentary <strong>and</strong> metavolcanic rocksthat comprise the orogenic middle crust. In the centralpart <strong>of</strong> the orthogneiss belt there is a narrow NEtrendingbelt <strong>of</strong> garnet–omphacite granulite <strong>and</strong> migmatitethat represents the orogenic lower crust. ACarboniferous granodiorite sill marks the easternboundary <strong>of</strong> the Lugian domain, with an Ordovicianleptyno-amphibolite unit extending further to the east(Fig. 7c). Directly in the footwall <strong>of</strong> the Lugi<strong>and</strong>omain there is an eclogite unit derived from the lowercrust <strong>of</strong> the Brunia continent (Sˇtı´ pska´ et al., 2006).Imbricated Brunia basement <strong>and</strong> cover thrust sheets <strong>of</strong>the Moravo–Silesian domain represent the <strong>structural</strong>lydeepest unit <strong>of</strong> the section. Superposed orogenic fabricsobserved in the field <strong>and</strong> their petrological characteristicsin four representative areas are describedbelow, three from the Moldanubian domain <strong>and</strong> onefrom the Lugian domain, where the spatial relationshipsbetween the orogenic lower crust <strong>and</strong> the adjacentorogenic middle crust may be critically evaluated.Structural development <strong>of</strong> the Moldanubian domainThe most characteristic <strong>structural</strong> feature <strong>of</strong> the Moldanubi<strong>and</strong>omain is the widespread flat-lying foliation.In the southern Moldanubian domain (Figs 2 & 5), theflat fabric gently dips to the east in the western part<strong>and</strong> to the west in the eastern part. This geometry ledAustrian geologists to define a large-scale Gfo¨ hl napperesting on middle crustal rocks in synformal structures(Tollmann, 1982; Fuchs, 1986). However, recentstudies have shown that the sub-horizontal fabrics (S 3 )were superimposed on vertical fabrics (S 2 ), <strong>and</strong>although the superposition is <strong>of</strong> variable intensity, itoccurs in almost all the rock types <strong>of</strong> the lower <strong>and</strong>middle crust.For the Moldanubian domain, we review three areaswhere the relationship <strong>of</strong> the steep to the flat fabricshas been studied in detail (Kolenovska´ et al., 1999;Racek et al., 2006; Tajcˇmanova´ et al., 2006; P. Sˇtı´ pska´& K. Schulmann, unpublished data). First, the <strong>structural</strong>succession is described on the macroscopic scale,where the S 2 fabric, defined mainly by steep lithologicalalternation <strong>and</strong> ⁄ or steep internal rock layering, wasreworked by multiple mechanisms into a flat-lying S 3fabric <strong>of</strong> variable intensity (Fig. 8). Later a detaileddiscussion is given <strong>of</strong> the metamorphic characterization<strong>of</strong> the steep S 2 fabric <strong>and</strong> metamorphic conditions<strong>of</strong> the flat D 3 reworking in individual lithologies.The first area is located in the northernmost termination<strong>of</strong> the southern Moldanubian domain (Figs 2 &5; Racek et al., 2006). Here, exceptionally well-preservedS 2 fabrics occur as steep alternations <strong>of</strong> paragneisses,marbles, felsic volcanic rocks, amphibolites<strong>and</strong> quartzites <strong>of</strong> the middle crust that are surroundedby migmatites <strong>and</strong> granulites <strong>of</strong> the orogenic lowercrust (Fig. 5, stereoplot E). The S 2 fabric was the result<strong>of</strong> isoclinal steep folding <strong>and</strong> from transposition <strong>of</strong>compositional b<strong>and</strong>ing S 1 in amphibolites (Fig. 8d).The S 2 foliation was refolded by open to closerecumbent F 3 folds in the central part <strong>of</strong> the middlecrustal domain (Fig. 8f), <strong>and</strong> approaching the contactswith surrounding orogenic lower crust the degree <strong>of</strong>flat D 3 reworking increases in conjunction with development<strong>of</strong> S 3 axial planar schistosity (Fig. 8e). Themineral stretching lineation trends NE–SW over thewhole area, in accordance with the orientation <strong>of</strong> thehinges <strong>of</strong> the F 3 folds (Fig. 5, stereoplot F). The S 3foliation clearly changes its orientation <strong>and</strong> dip, followingthe form <strong>of</strong> the orogenic middle crust, <strong>and</strong>similarly, the L 3 lineation plunges either to the SSW orFig. 5. (a) Structural map <strong>of</strong> a critical area <strong>of</strong> the southern Moldanubian domain with S 2 <strong>and</strong> S 3 fabrics (after Racek et al., 2006). SeeFig. 2 for regional location. The map shows a large area <strong>of</strong> orogenic middle crust, called the Drosendorf window, that is entirelysurrounded by orogenic lower crustal rocks. Structural trends (thin lines – S3 fabrics, thick lines – S 2 fabrics) indicate extrapolations<strong>of</strong> the major orientation <strong>of</strong> <strong>structural</strong> fabrics in the field. The density <strong>of</strong> trend lines indicates the homogeneity <strong>of</strong> fabric elements inthe field. (b) Stereograms: A – stereogram showing poles to S 3 fabrics in the area <strong>of</strong> orogenic lower crust north <strong>of</strong> the orogenicmiddle crustal domain; B – poles to S 3 fabrics from the orogenic lower crust <strong>and</strong> orogenic middle crust close to the northerntermination <strong>of</strong> the orogenic middle crust domain; C – poles to S 3 fabrics from the orogenic lower crust <strong>and</strong> orogenic middle crustalong the western border <strong>of</strong> the orogenic middle crust domain; D – poles to S 3 fabrics from the orogenic lower crust underlying theorogenic middle crust domain; E – stereogram showing poles to relicts <strong>of</strong> S 2 preserved within central part <strong>of</strong> the orogenic middlecrust domain; F – stereogram showing L 3 mineral <strong>and</strong> stretching lineations from the whole area. Equal area projection, lowerhemisphere, contoured at multiples <strong>of</strong> uniform distribution. (c) Interpretative cross-section shows main <strong>structural</strong> features describedin the text. The numbers on the cross-section <strong>and</strong> map show locations <strong>of</strong> samples used for P–T calculations shown in Fig. 9.Ó 2007 Blackwell Publishing Ltd140


EXHUMATION IN LARGE HOT OROGEN 281NNE according to the lenticular shape <strong>of</strong> the orogenicmiddle crustal (Fig. 5, stereoplots B, C & D).The second area is the southernmost part <strong>of</strong> the3000 km 2 region <strong>of</strong> orogenic lower crust <strong>of</strong> the centralMoldanubian domain (central part <strong>of</strong> Fig. 2, <strong>and</strong>northern part <strong>of</strong> Fig. 5; Kolenovska´ et al., 1999). Here,the orogenic lower crust is a vast domain typicallycomposed <strong>of</strong> felsic migmatitic orthogneisses <strong>and</strong>migmatites containing large bodies <strong>of</strong> Ky–Kfsgranulite, eclogite <strong>and</strong> peridotite. The whole area is(a)(b)ABDPaPaMiAmFeMiMi(c)PaPaMiAmMiFeAmMiOrÓ 2007 Blackwell Publishing Ltd141


282 K. SCHULMANN ET AL.characterized by a flat migmatitic S 3 foliation (Fig. 5,stereoplot A) that is axial planar to rare isoclinal foldswith hinges generally trending NNE–SSW, parallel tostretching <strong>and</strong> mineral lineations. Numerous shearb<strong>and</strong>s filled with leucosome indicate a dominant topto-the-NEshearing (Urban, 1992; Schulmann et al.,1994). The eclogites <strong>and</strong> granulites form lozengeshapedboudins surrounded by migmatitic fabric.These rocks locally contain relics <strong>of</strong> early NE-trendingsteep S 2 fabric, but because <strong>of</strong> their scarce occurrence,(a)(b)PaMiMiOrFeSy(c)PaPaOrFeÓ 2007 Blackwell Publishing Ltd142


EXHUMATION IN LARGE HOT OROGEN 283we cannot demonstrate the regional extent <strong>of</strong> the S 2fabric in this area (Fig. 8h).The third area is located in the northern Moldanubi<strong>and</strong>omain (Figs 2 & 6) <strong>and</strong> is characterized bythrusting <strong>of</strong> a NNW–SSE-elongated granulite bodyover the middle crust (Fig. 6a, c). The granulite showsan exceptionally well-preserved steep NNW–SSEtrendinggranulite-to-amphibolite facies S 2 foliation(Fig. 6, stereoplot A). D 3 shallow to moderately steep,south-dipping shear zones, which show top to theNNE shear sense, cut the S 2 fabric heterogeneously.The surrounding migmatites <strong>and</strong> migmatitic orthogneissesexhibit predominantly S 3 fabric, which dipsgently to moderately to the south (Fig. 6, stereoplotB). This foliation, formed essentially by migmatiticb<strong>and</strong>ing, contains numerous close to isoclinal F 3 foldswith hinges trending NNE–SSW parallel to minerallineation. Within the adjacent middle crust, thedeformation is also polyphase, comprising an earlyfoliation that dips steeply to the SW, which was reworkedby a schistosity that dips gently in the samedirection (Fig. 6, stereoplot C).Structural development <strong>of</strong> the Lugian domainThe northern part <strong>of</strong> the Lugian domain is characterizedby a central granulite belt surrounded on thewestern <strong>and</strong> eastern sides by strongly migmatized orthogneisses(Figs 7c & 8b). The granulite <strong>and</strong> theadjacent orthogneisses reveal the NNE–SSW-trendingvertical S 2 foliation parallel to the trend <strong>of</strong> the orogeniclower crust (Fig. 7, stereoplot A). This S 2 foliationcontains rootless folds formed by layers <strong>of</strong> metabasitethat preserve evidence <strong>of</strong> an early S 1 foliation (Sˇtípska´et al., 2004).East <strong>of</strong> the granulite belt, the orthogneisses <strong>and</strong>migmatite show reworking <strong>of</strong> the vertical S 2 fabric bymoderately west-dipping S 3 foliation (Fig. 8g). The dip<strong>of</strong> the S 3 fabric progressively decreases eastwards towardsthe underlying Ordovician lower crustal leptyno-amphiboliteunit (Fig. 7, stereoplot B); agranodiorite sill emplaced syntectonic with the D 3deformation marks the boundary between this unit<strong>and</strong> the Lugian orthogneiss, dating the D 3 structures atc. 340 Ma (Parry et al., 1997; Sˇtípska´ et al., 2001).Importantly, the structure <strong>of</strong> the leptyno-amphiboliteunit is discordant with respect to D 3 fabrics <strong>of</strong> thewestern Lugian orthogneisses; the leptyno-amphiboliteunit is inferred to be <strong>of</strong> Early Ordovician age, based onU–Pb zircon dates that yield an age <strong>of</strong> c. 510 Ma(Fig. 7, cross-section; Sˇtı´ pska´ et al., 2001; Lexa et al.,2005).To the NW from the granulite belt, the S 2 fabric isaffected by sub-horizontal, variably dipping amphibolitefacies S 3 fabric (Fig. 8a,b). This S 3 fabric hasintensely reworked the steep S 2 foliation <strong>of</strong> the adjacentorogenic middle crust to the point that it progressivelypasses into a several hundred metre widenormal-sense shear zone dipping gently to the NE(Fig. 8c).P–T–t history <strong>of</strong> the Moldanubian–Lugian rootThe relationship between the prograde <strong>and</strong>retrograde P–T paths <strong>and</strong> orogenic fabrics is a keypiece <strong>of</strong> information necessary to underst<strong>and</strong> thethermo-mechanical processes that operated withinthe orogenic root. Micro<strong>structural</strong>, petrological <strong>and</strong>thermodynamic <strong>modelling</strong> studies during the lastfive years have revealed a systematic pattern <strong>of</strong> P–Tevolution related to the early steep fabrics in boththe orogenic lower <strong>and</strong> middle crust. In addition,combination <strong>of</strong> the P–T evolution with Sm–Nd <strong>and</strong>40 Ar ⁄ 39 Ar cooling ages reveals important differencesin the thermo-chronological evolution <strong>of</strong> theorogen.Micro<strong>structural</strong> <strong>and</strong> petrographic characterization <strong>of</strong> S 2 <strong>and</strong>S 3 fabricsIn the orogenic lower crust, the steep fabric is characterizedby compositional b<strong>and</strong>ing formed by mafic<strong>and</strong> felsic granulite (Sˇtı´ pska´ et al., 2004). More commonlyS 2 is defined by a mineral fabric, marked byoriented kyanite, biotite <strong>and</strong> recrystallized ribbons <strong>of</strong>quartz <strong>and</strong> feldspar in the case <strong>of</strong> the felsic granulite(Tajcˇmanova´ et al., 2006), <strong>and</strong> by the elongated shape<strong>of</strong> garnet <strong>and</strong> aligned omphacite grains in the case <strong>of</strong>the mafic granulite (Sˇtı´ pska´ et al., 2004). In the orogeniclower crust composed <strong>of</strong> migmatitic orthogneisses,the S 2 fabric is defined by alternation <strong>of</strong> infinitemonomineralic recrystallized b<strong>and</strong>s <strong>of</strong> quartz, plagioclase<strong>and</strong> K-feldspar, in an assemblage with kyanite,biotite <strong>and</strong> garnet that results from a high-temperatureFig. 6. (a) Structural map <strong>of</strong> an area showing thrusting <strong>of</strong> the orogenic lower crust over the orogenic middle crustal in the NE part <strong>of</strong>the Moldanubian domain, with regional S 2 <strong>and</strong> S 3 fabrics as shown in Fig. 5 (after Tajčmanova´ et al., 2006). See Fig. 2 for regionallocation. The map shows a large body <strong>of</strong> orogenic lower crustal granulites that preserve relicts <strong>of</strong> the S 2 fabrics. The <strong>structural</strong>trends (thin layers – S 3 fabrics, thick layers – S 2 fabrics) indicate extrapolations <strong>of</strong> major orientations <strong>of</strong> <strong>structural</strong> fabrics in the field.The density <strong>of</strong> trend-lines indicates the homogeneity <strong>of</strong> fabric elements in the field. (b) Stereograms: A – stereogram shows poles to S 2fabrics in the orogenic lower crustal granulites; B – stereogram <strong>of</strong> poles to S 3 fabrics from the orogenic middle crust. Squares instereogram show integrated directions <strong>of</strong> L 3 lineations from the area. C – Stereogram <strong>of</strong> poles to S 3 fabrics from the orogeniclower crustal granulites <strong>and</strong> associated migmatites. Equal area projection, lower hemisphere, contoured at multiples <strong>of</strong> uniformdistribution. (c) Simplified cross-section shows well-preserved S 2 fabrics in competent granulites surrounded by cordierite migmatitescompletely reworked by S 3 fabric. The migmatites <strong>and</strong> granulites <strong>of</strong> the orogenic lower crust are thrust over orogenic middle crustto the east. The numbers on the cross-section <strong>and</strong> map show locations <strong>of</strong> samples used for P–T calculations in Fig. 9.Ó 2007 Blackwell Publishing Ltd143


284 K. SCHULMANN ET AL.(a)(b)OrMeOmLaGrMeFe(c)OrOrMeOmLaGrMeFeÓ 2007 Blackwell Publishing Ltd144


EXHUMATION IN LARGE HOT OROGEN 285<strong>and</strong> medium- to high-pressure recrystallization <strong>of</strong>coarse-grained granite (Hasalova´ et al., 2008a). Inareas inferred to be a part <strong>of</strong> the orogenic lower crust,based on the presence <strong>of</strong> HP relicts <strong>and</strong> the high degree<strong>of</strong> anatexis, the S 2 fabric is defined by steep compositionallayering in amphibolites, felsic metavolcanicrocks, orthogneisses <strong>and</strong> paragneisses.In the orogenic middle crust the S 2 fabric is characterizedby a steep lithological alternation <strong>of</strong> paragneisseswith amphibolites, felsic metavolcanic rocks,quartzites, marbles <strong>and</strong> calcsilicate rocks (Romanova´&Sˇtípska´ , 2001; Racek et al., 2006). In the amphibolites,the steep fabric is characterized by alignedhornblende <strong>and</strong> recrystallized plagioclase ribbons inassociation with garnet that indicates medium-pressureconditions for formation <strong>of</strong> this fabric. Evidence <strong>of</strong> theprograde fabrics in the paragneisses <strong>of</strong> the Moldanubi<strong>and</strong>omain occurs only as rare oriented inclusions <strong>of</strong>staurolite <strong>and</strong> kyanite in garnet <strong>and</strong> remnants <strong>of</strong> theseminerals in the matrix (Racek et al., 2006). Even insome areas <strong>of</strong> the Moldanubian domain where themacroscopic fabric <strong>of</strong> the paragneisses remained steep,the higher-grade assemblages were commonly overprintedpassively by lower-grade assemblages with sillimaniteor cordierite, which we ascribe to the highreactivity <strong>of</strong> the paragneisses in the presence <strong>of</strong> melt.As the other lithologies do not show variation <strong>of</strong> theassemblage with changing P–T conditions, the mainguide in characterizing the steep fabric in the orogenicmiddle crust <strong>of</strong> the Moldanubian domain is the steeplithological alternation combined with peak P–T conditions.In the Lugian domain, in the orogenic middle crustthe characterization <strong>of</strong> the steep fabric is much betterconstrained because <strong>of</strong> the absence <strong>of</strong> anatexis in themetasedimentary rocks <strong>and</strong> the absence or limitedextent <strong>of</strong> anatexis in the orthogneisses. In the orthogneisses,the S 2 fabric is marked by steep alternation <strong>of</strong>recrystallized augen <strong>and</strong> ribbons <strong>of</strong> quartz <strong>and</strong> feldspar<strong>and</strong> by the orientation <strong>of</strong> biotite <strong>and</strong> muscovite. Steeplocalized ultramylonitic zones are developed locally.The prograde <strong>and</strong> peak P–T conditions <strong>of</strong> the steepfabrics in Lugian metapelites are constrained by thesuccessive growth <strong>of</strong> garnet, staurolite <strong>and</strong> kyanitein the micaschists (Romanova´ &Sˇtípska´ , 2001; Jastrzębski,2005; Murtezi, 2006).As described above, the flat fabric in the orogeniclower <strong>and</strong> middle crust commonly originates throughhorizontal folds affecting the steep fabric <strong>and</strong> is onlyrarely associated with sub-horizontal or gently dipping,thrust-related shear zones. Highly anisotropicrocks, such as b<strong>and</strong>ed orthogneisses, show strong axialplanar crenulation cleavage, which, in places, passesgradually into a sub-horizontal foliation with evidence<strong>of</strong> the early vertical anisotropy occurring only as a fewrootless folds (Fig. 8h). The intensity <strong>and</strong> characteristicmineralogy <strong>of</strong> the flat reworking is variabledepending on the specific area, lithology <strong>and</strong> degree <strong>of</strong>metamorphism. However, some mineralogical featuresare common for the Ky–Kfs granulites, the orthogneisses<strong>and</strong> the paragneisses. These include thewidespread development <strong>of</strong> the assemblages garnet–sillimanite–K-feldspar, in the Moldanubian domain,<strong>and</strong> garnet–muscovite ± sillimanite, in the Lugi<strong>and</strong>omain, showing the overall moderate pressure character<strong>of</strong> the S 3 reworking <strong>and</strong> the higher temperature inthe Moldanubian domain compared with the Lugi<strong>and</strong>omain. Because <strong>of</strong> the variability <strong>of</strong> the S 3 structuresalong the orogen, the detailed mineralogical characteristics<strong>and</strong> P–T conditions for both the S 2 <strong>and</strong> S 3fabrics for the four areas identified above are describedbelow.P–T evolution <strong>of</strong> the Moldanubian domainRacek et al. (2006) examined petrological relationshipsbetween the steep <strong>and</strong> flat fabrics in the orogenic lower<strong>and</strong> middle crust <strong>of</strong> the southern Moldanubian domain(Figs 5 & 9a). Modelling <strong>of</strong> garnet zoning in eclogitefrom the western part <strong>of</strong> the orogenic lower crust (no. 5in Figs 5 & 9a) indicates burial process to have takenplace from around 10 kbar <strong>and</strong> 700 °C to around 15–16 kbar <strong>and</strong> 800 °C. Similar peak conditions aredetermined for kyanite–sillimanite migmatite <strong>structural</strong>lyabove the eclogite (around 14 kbar at 750 °C;no. 6 in Figs 5 & 9a). However, these prograde <strong>and</strong>peak conditions are associated with the steep S 2 fabriconly as the early <strong>structural</strong> relations <strong>of</strong> relict mineralsFig. 7. (a) Structural map <strong>of</strong> a critical area showing thin vertical belt <strong>of</strong> the orogenic lower crust surrounded by orogenic middlecrust in the eastern part <strong>of</strong> the Lugian domain with regional S 2 <strong>and</strong> S 3 fabrics as in Fig. 5 (after Sˇtı´ pska´ et al., 2004). See geological mapin Fig. 2 for regional location. The map shows a narrow body <strong>of</strong> the orogenic lower crustal granulites, preserving relicts <strong>of</strong> the S 2fabrics, <strong>and</strong> the surrounding fabric <strong>of</strong> the orogenic middle crust. The fabric in the mafic lower crust <strong>and</strong> the syntectonic tonalite is alsoshown. The <strong>structural</strong> trends (thin layers – S 3 fabrics, thick layers – S 2 fabrics) indicate extrapolations <strong>of</strong> major orientations <strong>of</strong><strong>structural</strong> fabrics in the field. Density <strong>of</strong> trend lines indicates the homogeneity <strong>of</strong> fabric elements in the field. Stereograms:A – stereogram <strong>of</strong> poles to S 2 fabrics in the garnet–omphacite-bearing granulites; B – stereogram <strong>of</strong> poles to S 3 fabrics from theorogenic lower crust <strong>and</strong> surrounding orogenic middle crust rocks represented by thin dots <strong>and</strong> contoured. Thick dots in the stereogramsshow S 3 foliations typical for the Ordovician metagabbros <strong>and</strong> Carboniferous tonalite sill <strong>of</strong> the Lugian mafic complex.C – Stereogram <strong>of</strong> L 3 directions from the granulite belt <strong>and</strong> surrounding orogenic middle crust (thin dots <strong>and</strong> contoured data).Black squares show mineral <strong>and</strong> stretching lineations from mylonitized gabbros <strong>and</strong> granodiorite sill. Note strong discrepancy inorientations <strong>of</strong> S 3 foliations <strong>and</strong> L 3 lineations between granulite belt <strong>and</strong> adjacent Ordovician mafic complex. Equal area projection,lower hemisphere, contoured at multiples <strong>of</strong> uniform distribution. (c) Simplified cross-section shows the vertical central granulitebelt showing thrusting <strong>of</strong> the orogenic lower crust over the Ordovician leptyno-amphibolite complex. Note the syntectonic D 3intrusion <strong>of</strong> the 340 Ma granodiorite sill <strong>and</strong> discordant fabrics in the 510 Ma leptyno-amphibolite complex.Ó 2007 Blackwell Publishing Ltd145


286 K. SCHULMANN ET AL.Fig. 8. Field photographs <strong>of</strong> the S 2 –S 3 relationships. (a) Well preserved S 2 fabric in the garnetiferous amphibolite in the orogenicmiddle crust <strong>of</strong> the Lugian domain. (b) Strongly reworked S 2 fabric in the amphibolite in the orogenic middle crust <strong>of</strong> the Lugi<strong>and</strong>omain. (c) Complete transposition <strong>of</strong> early fabric in new S 3 in epidote amphibolite in the orogenic middle crust <strong>of</strong> the Lugian domain.(d) Excellent preservation <strong>of</strong> the steep S 2 fabric in garnetiferous amphibolite in the orogenic middle crust <strong>of</strong> the southern Moldanubi<strong>and</strong>omain. (e) S 2 folded by F 3 folds in the felsic granulite <strong>of</strong> the orogenic lower crust in the southern Moldanubian domain.(f) Asymmetrical folding <strong>of</strong> the S 2 fabric in the S 3 channel-flow fabric <strong>of</strong> the orogenic lower crust in southern Modanubi<strong>and</strong>omain. (g) Character <strong>of</strong> S 2 foliation in high-grade b<strong>and</strong>ed orthogneiss <strong>of</strong> the Lugian domain. (h) Micr<strong>of</strong>olding <strong>and</strong> partial transposition<strong>of</strong> high-grade S 2 fabric in the Gfo¨ hl gneiss <strong>of</strong> the southern Moldanubian domain.Ó 2007 Blackwell Publishing Ltd146


EXHUMATION IN LARGE HOT OROGEN 287were obliterated by strong S 3 reworking. High-pressureconditions are also reported from the Ky–Kfs granulitebody to the east (15 kbar, 800 °C, no. 2 in Figs 5 &9a), also part <strong>of</strong> the orogenic lower crust, where theretrieved P–T is clearly related to the S 2 fabric.The retrograde path <strong>of</strong> rocks from the orogeniclower crust situated far from the Brunia margin ischaracterized by an almost isothermal decompressionto about 7 kbar, as recorded by the S 3 assemblagegarnet-hornblende-plagioclase in eclogite <strong>and</strong> garnet–sillimanite–biotite in migmatite (no. 5, 6 in Fig. 9a).Unlike the central part <strong>of</strong> the Moldanubian, the Ky–Kfs granulite forming the orogenic lower crust at theboundary with the Brunia margin shows cooling withdecompression (no. 2 in Fig. 9a). Modelling <strong>of</strong> zoningin garnet that is syntectonic with the S 2 fabrics inparagneisses <strong>of</strong> the orogenic middle crust is consistentwith increase <strong>of</strong> P–T conditions to about 9 kbar <strong>and</strong>700 °C (nos 3 & 4 in Figs 5 & 9a). The retrograde P–Tpath <strong>of</strong> the sillimanite-bearing S 3 fabric in rocks <strong>of</strong> theorogenic middle crust is associated with decrease <strong>of</strong>pressure coupled with increase <strong>of</strong> temperature, toabout 7 kbar <strong>and</strong> 700–750 °C (nos 3 & 4 in Fig. 9a).Kolenovska´ et al. (1999), Sˇtı´ pska´ & Powell (2005a)<strong>and</strong> P. Sˇtípska´ & K. Schulmann (unpublished data)examined the P–T evolution <strong>of</strong> rocks <strong>and</strong> relics <strong>of</strong> boththe orogenic lower <strong>and</strong> middle crust over a large areadominated by flat S 3 fabric in the central Moldanubi<strong>and</strong>omain (Fig. 2). Here, Ky–Kfs granulites register peakconditions around 18 kbar <strong>and</strong> 850 °C with no clearindication <strong>of</strong> the prograde path (no. 8 in Figs 5 & 9b).In contrast, a MORB-type eclogite provides evidence<strong>of</strong> the prograde path from about 10 kbar <strong>and</strong> 750 °Cto 17–18 kbar in the form <strong>of</strong> inclusions <strong>of</strong> hornblende<strong>and</strong> plagioclase in a prograde garnet (no. 7 in Fig. 9b).Both samples from the orogenic lower crust showalmost isothermal decompression <strong>and</strong> strong D 3reworking at 7)10 kbar <strong>and</strong> 750–850 °C (nos 7 & 8 inFig. 9b). Associated paragneisses from the orogeniclower crust contain early staurolite <strong>and</strong> kyanite inclusionsin garnet that preserves prograde zoning, indicatingburial to about 10 kbar <strong>and</strong> 700 °C (no. 9 inFigs 5 & 9b; P. Sˇtípska´ & K. Schulmann, unpublisheddata). The matrix was completely transposed by the D 3deformation, as marked by sillimanite <strong>and</strong> biotitedeveloped during decompression (no. 9 in Figs 5& 9b). In another paragneiss the flat sillimanite–biotite-bearingfabric was overgrown by garnet, indicatingheating at around 7)8 kbar <strong>and</strong> 700)850 °C (no. 10 inFig. 9b; P. Sˇtípska´ & K. Schulmann, unpublisheddata). The prograde conditions <strong>and</strong> peak pressuresfrom samples <strong>of</strong> this area could not be directly linkedto the steep fabric because <strong>of</strong> almost complete transposition<strong>of</strong> rocks by the S 3 fabric. However, the progradeinclusions in garnet from the eclogite <strong>and</strong> fromthe paragneiss form straight, sub-vertical internalfabrics oriented at a high angle to the external S 3foliation, indicating that these garnet have probablygrown in the steep fabric.The steep S 2 fabric in the granulite body rimming theNE margin <strong>of</strong> the Moldanubian domain is marked bygarnet, kyanite <strong>and</strong> K-feldspar indicating conditions <strong>of</strong>about 18 kbar at 850 °C, whereas the crystallization <strong>of</strong>biotite in the same fabric occurred during decompressionaccompanied by slight cooling (no. 11 in Figs 6& 9c; Tajcˇmanova´ et al., 2006). Within the S 2 fabric,sillimanite–biotite intergrowths replacing garnet indicatesignificant decompression <strong>and</strong> cooling (no. 11 inFig. 9c). The S 3 fabric is associated with hercyniterimmingmetastable kyanite in the granulite <strong>and</strong> withthe development <strong>of</strong> cordierite, sillimanite <strong>and</strong> biotite inadjacent gneisses <strong>and</strong> migmatites, indicating conditions<strong>of</strong> around 4 kbar <strong>and</strong> 700 °C (nos 12, 25 in Figs 6& 9c). Muscovite–biotite schists <strong>of</strong> the adjacent orogenicmiddle crust contain sillimanite that overgrowsrelicts <strong>of</strong> kyanite in the matrix. The peak P–T conditionsestimated by Pitra & Guiraud (1996) are 8–9 kbarat 610–660 °C, which was followed by near-isothermaldecompression, as recorded in the whole middle crustalcomplex, to 4–6 kbar (no. 13 in Figs 5 & 9c).P–T evolution <strong>of</strong> the Lugian domainEarly petrological studies <strong>of</strong> granulite from the orogeniclower crust suggested peak conditions <strong>of</strong> about28 kbar at 1000 °C (Bro¨ cker & Klemd, 1996; Kryzaet al., 1996; Klemd & Bro¨ cker, 1999), based on tw<strong>of</strong>eldsparthermometry <strong>and</strong> Grt–Ky–Qtz–Pl barometry.However, Sˇtípska´ et al. (2004) questioned these resultsbecause <strong>of</strong> the possibility <strong>of</strong> non-equilibrium compositions,which may have been involved in the calculations.Instead, these authors proposed that peakpressures were around 18–20 kbar at 800)900 °C (no.14 in Figs 7 & 9d). Sˇtípska´ et al. (2004) further proposedthat these conditions are characteristic for thesteep S 2 foliation. These authors also constrainedconditions to around 10 kbar <strong>and</strong> 700 °C for theamphibolite facies retrogression within the S 3 fabric(no. 14 in Fig. 9d). Petrological studies <strong>and</strong> <strong>modelling</strong><strong>of</strong> prograde garnet zoning from rocks <strong>of</strong> the Ky–St–Grt micaschists revealed a prograde path up to about10 kbar <strong>and</strong> 650 °C (no. 15 in Figs 7 & 9d; Romanova´&Sˇtı´ pska´ , 2001; Jastrzębski, 2005). A micro<strong>structural</strong>study confirmed that the growth <strong>of</strong> prograde garnetwas syntectonic with the steep S 2 fabric. The mainreworking in the horizontal S 3 fabrics occurred in thefield <strong>of</strong> sillimanite stability at about 7 kbar <strong>and</strong> 650 °C(no. 15 in Fig. 9d).The Ordovician metamorphic fabric in the leptynoamphiboliteunit developed at about 10 kbar <strong>and</strong>800 °C (no. 18 in Figs 7 & 9d; Sˇtípska´ et al., 2001;Lexa et al., 2005). Carboniferous metamorphismrelated to the S 3 reworking <strong>of</strong> rocks at the westernmargin <strong>of</strong> this unit, close to the Carboniferous sillintruded at around 7 kbar (no. 17 in Figs 7 & 9d; Lexaet al., 2005), occurred under similar conditions <strong>of</strong>around 8 kbar <strong>and</strong> 750 °C (no. 16 in Figs 7 & 9d;Baratoux et al., 2005).Ó 2007 Blackwell Publishing Ltd147


288 K. SCHULMANN ET AL.(d)(c)PrRe(b)RePrPrPe(a)PePrReRe(e)(f)AAFig. 9. P–T diagrams showing prograde <strong>and</strong> retrograde P–T paths <strong>of</strong> samples studied in the <strong>structural</strong> context discussed in the text.(a) Southern Moldanubian area in Fig. 5: Source <strong>of</strong> data: 1 – kyanite micaschist (Sˇtı´ pska´ & Schulmann, 1995); 2 – Ky–Kfsgranulite (Racek et al., 2006); 3 – kyanite micaschist (Racek et al., 2006); 4 – staurolite-kyanite paragneiss (Racek et al., 2006);5 – retrogressed eclogite (Racek et al., 2006); 6 – sillimanite migmatite (Racek et al., 2006). (b) Central Moldanubian area in Fig. 5:7 – eclogite (Sˇtı´ pska´ & Powell, 2005a); 8 – Ky–Kfs granulite (P. Sˇtı´ pska´ & K. Schulmann, unpublished data), 9 – migmatitic paragneiss(P. Sˇtı´ pska´ & K. Schulmann, unpublished data); 10 – migmatitic paragneiss (P. Sˇtı´ pska´ & K. Schulmann, unpublished data).(c) Central Moldanubian area in Fig. 6: 11 – Ky–Kfs granulite (Tajcˇmanova´ et al., 2006); 12 – cordierite gneiss (Tajcˇmanová et al.,2006); 13 – metapelite (Pitra & Guiraud, 1996). (d) Lugian domain in Fig. 7: 14 – garnet–omphacite-bearing granulite (Sˇtı´ pskáet al., 2004); 15 – staurolite–kyanite micaschist (Romanova´ &Sˇtı´ pska´ , 2001); 16 – amphibolite (Lexa et al., 2005); 17 – granodiorite(Parry et al., 1997); 18 – Ordovician metapelitic granulite (Lexa et al., 2005). (e, f): Probability curves <strong>of</strong> age histograms dealingwith existing Moldanubian <strong>and</strong> Lugian Sm–Nd <strong>and</strong> 40 Ar ⁄ 39 Ar cooling ages. Bt – biotite 40 Ar ⁄ 39 Ar cooling ages; Grt–Cpx <strong>and</strong>Grt–whole rock, Sm–Nd ages; Hbl – hornblende <strong>and</strong> Mu – muscovite, Ar 40 ⁄ Ar 39 cooling ages.Ó 2007 Blackwell Publishing Ltd148


EXHUMATION IN LARGE HOT OROGEN 289Thermochronology <strong>of</strong> Moldanubian <strong>and</strong> Lugian domainsThe distribution <strong>of</strong> isotopic ages reflects a two-stagecooling, related to exhumation processes. An olderstage is marked by cooling ages from minerals withhigh blocking temperatures (Sm–Nd system in garnet,700)750 °C, Hensen & Zhou, 1995;40 Ar ⁄ 39 Ar inhornblende, around 480 °C, Harrison et al., 1985).Systematically, younger cooling ages are retrievedfrom minerals with low blocking temperatures( 40 Ar ⁄ 39 Ar in muscovite, around 350 °C, <strong>and</strong> in biotite,around 280 °C, Harrison et al., 1985). Peridotites<strong>and</strong> associated eclogites yield Sm–Nd whole-rock–garnet ages between 350 <strong>and</strong> 325 Ma (Brueckner et al.,1991; Beard et al., 1992; Medaris et al., 1995). Thepeak for Sm–Nd ages for both types <strong>of</strong> rocks is around336 Ma (Fig. 9e). The 40 Ar ⁄ 39 Ar hornblende data arecompatible with the distribution <strong>of</strong> Sm–Nd ages,whereas the 40 Ar ⁄ 39 Ar muscovite data record systematicallyyounger ages ranging from 331 to 325 Ma(Fig. 9e; Matte et al., 1990; Dallmeyer et al., 1992;Fritz et al., 1996).The Variscan cooling history <strong>of</strong> the Lugian eclogitesis documented by Sm–Nd garnet–clinopyroxene–whole-rock ages <strong>of</strong> 340)330 Ma (Brueckner et al.,1991). Recently, Lange et al. (2005) provided Sm–Ndisochrons for the Lugian granulites that define agesbetween 357 <strong>and</strong> 337 Ma, which fit the existing40 Ar ⁄ 39 Ar hornblende <strong>and</strong> biotite cooling ages fromthis region (Schneider et al., 2006). The 40 Ar ⁄ 39 Armuscovite ages from all lithologies <strong>of</strong> the Lugi<strong>and</strong>omain are similar to Sm–Nd garnet,40 Ar ⁄ 39 Arhornblende <strong>and</strong> 40 Ar ⁄ 39 Ar biotite data from granulites<strong>and</strong> eclogites, indicating a rapid <strong>and</strong> monocycliccooling history (Fig. 9f).Polyphase fabric <strong>and</strong> metamorphic evolution <strong>of</strong> theorogenic beltThe <strong>structural</strong> pattern in the Lugian domain reveals awell-preserved D 2 vertical fabric in the orogenic lower<strong>and</strong> middle crust as well as coherency <strong>of</strong> these units,defined by sub-parallel alternations <strong>of</strong> continuous belts<strong>of</strong> orogenic lower <strong>and</strong> middle crust on the geologicalmap. Here, the geometry <strong>of</strong> the geological units <strong>and</strong>map patterns are fully controlled by the D 2 deformation.The D 3 deformation was generally weak <strong>and</strong>heterogeneous, being mostly concentrated close to thethrust <strong>of</strong> the Lugian domain over the Ordovician leptyno-amphiboliteunit to the east <strong>and</strong> close to the area<strong>of</strong> flat fabric <strong>and</strong> the normal-sense shear zone boundingthe Lugian domain in the west (Figs 7 & 8).Normal-sense non-coaxial shearing developed continuouslyfrom the earlier pure shear deformation <strong>and</strong>probably this was responsible for displacement <strong>of</strong>schists <strong>of</strong> the orogenic middle crust to the NW <strong>and</strong>unro<strong>of</strong>ing <strong>of</strong> the orogenic lower crust.In the orogenic lower <strong>and</strong> middle crust <strong>of</strong> the Lugi<strong>and</strong>omain, petrology indicates that retrograde conditions<strong>of</strong> omphacite-bearing granulites <strong>and</strong> peak pressureconditions <strong>of</strong> adjacent micaschists are similar. In theorogenic middle crust, the prograde mineral growth isclearly related to the steep S 2 fabric (Romanova´ &Sˇtı´ pska´ , 2001), whereas in the granulites from theorogenic lower crust, development <strong>of</strong> the S 2 fabric isrelated to their retrogression (Sˇtípska´ et al., 2004). Theremarkable differences in prograde <strong>and</strong> retrogradeP–T paths <strong>of</strong> rocks from the orogenic middle <strong>and</strong>lower crust, respectively, related to steep S 2 foliationsmakes this region the best example <strong>of</strong> vertical material<strong>and</strong> heat transfer during D 2 .Sˇtípska´ et al. (2001) further discussed the significance<strong>of</strong> similar Carboniferous <strong>and</strong> Ordovicianmetamorphic conditions for the easterly lower crustalleptyno-amphibolite complex. They concluded thatthis complex cooled after Ordovician rifting <strong>and</strong>remained stable in the crust before the onset <strong>of</strong>Variscan deformation. Heterogeneous Variscanamphibolite facies deformation was localized in thewestern margin <strong>of</strong> the Ordovician block <strong>and</strong> sharesthe same P–T conditions <strong>and</strong> kinematics withamphibolite facies retrograde fabric <strong>of</strong> the easternmostorogenic lower crust. Consequently, Sˇtı´ pska´et al. (2004) suggested that the transition from thesteep-to-the-west moderately dipping fabrics in theorogenic lower crust was linked to thrusting <strong>of</strong> theserocks over a rigid lower crustal block made up <strong>of</strong>leptyno-amphibolite unit rocks at a depth equivalentto about 10 kbar (Fig. 5c).In contrast, the horizontal S 3 fabrics affectingrocks <strong>of</strong> the orogenic lower crust <strong>and</strong> the normal-senseshear zone developed in the west were linked to thedevelopment <strong>of</strong> retrograde <strong>and</strong> syntectonic mineralassemblages indicating an important decrease <strong>of</strong> temperature<strong>and</strong> pressure typical for detachment zones(e.g. V<strong>and</strong>erhaeghe & Teyssier, 2001). The Lugian rootrecords a single-stage cooling history characterized bytelescoping <strong>of</strong> ages for minerals with different blockingtemperatures.The <strong>structural</strong> development <strong>of</strong> the NE termination<strong>of</strong> the Moldanubian domain shares a number <strong>of</strong>features with the Lugian domain. In this area, the S 2foliation was well preserved in felsic granulites inconjunction with a weak D 3 reworking, whichemphasizes a coherency <strong>of</strong> the lower <strong>and</strong> middleorogenic crust <strong>and</strong> indicates that the map patterndeveloped during D 2 vertical movements, similar tothat in the Lugian domain. The steep fabric in theorogenic lower crust records retrogression to the sillimanitestability field, suggesting that the orogeniclower crust was exhumed to middle crustal conditionsalong the S 2 fabric. The HP characteristics <strong>of</strong> thisfabric are not necessarily always preserved, preservationbeing dependent mainly on whether the crustavoids re-hydration during exhumation. However, incontrast with the Lugian domain, the heterogeneousD 3 deformation does not show any normal-sensecomponent <strong>of</strong> shear <strong>and</strong> is exclusively associated withÓ 2007 Blackwell Publishing Ltd149


290 K. SCHULMANN ET AL.top-to-the-NE-oriented thrusting in the stability field<strong>of</strong> cordierite at low pressures <strong>of</strong> 3–4 kbar.Structural <strong>and</strong> petrological studies <strong>of</strong> the southern<strong>and</strong> central Moldanubian domains show that similarP–T conditions around 7 kbar at 750 °C are characteristicfor all crustal levels during development <strong>of</strong> theflat S 3 fabric. Rocks from different depths were mixed<strong>and</strong> reworked together during the D 3 deformation, asevidenced by contrasting prograde P–T paths withdifferent pressure peaks. The increase <strong>of</strong> temperaturein the orogenic middle crust <strong>and</strong> the slight decrease <strong>of</strong>temperature in the orogenic lower crust during retrogressionsuggest a mutual thermal equilibration duringdevelopment <strong>of</strong> the S 3 flat fabric.Structural observations show that all rocks werestrongly deformed during D 3 horizontal flow, althoughmore competent lithologies, mostly middle crust <strong>and</strong>some granulites, retained their original steep S 2 fabric,whereas in the less competent, weak lower crust, thehorizontal top-to-the-NE D 3 ductile shearing dominates.During this process, the originally coherentorogenic middle <strong>and</strong> lower crust was disaggregated t<strong>of</strong>orm boudins <strong>and</strong> rootless folds in a pervasivelyflowing migmatitic matrix. We suggest that providingthe volume <strong>of</strong> orogenic middle crustal is high enough,the mineral assemblages from the D 2 vertical fabricsare preserved (Figs 6 & 9a). In contrast, the smallerboudins were completely re-equilibrated <strong>and</strong> evenheated in the flowing mass <strong>of</strong> hot orogenic lower crust(Figs 6 & 9b). The metamorphic conditions associatedwith the D 3 flow show clearly that pressure, temperature<strong>and</strong> intensity <strong>of</strong> D 3 reworking decreases fromsouth to north across the whole continental margin.Based on dating <strong>of</strong> metamorphic zircon from granulites,the timing <strong>of</strong> HP metamorphism <strong>and</strong> S 2 fabricformation in the Moldanubian domain was estimatedby Schulmann et al. (2005) <strong>and</strong> Tajcˇmanova´ et al.(2006) to occur between 350 <strong>and</strong> 340 Ma. This agespan corresponds with cooling ages <strong>of</strong> minerals withhigh blocking temperatures in the northern Moldanubi<strong>and</strong>omain (Matte et al., 1990; Macintyre et al.,1992), <strong>and</strong> indicates that already during the D 2 stagean important cooling <strong>and</strong> exhumation <strong>of</strong> this area wastaking place, similar to the Lugian domain. Theyounger 330–325 Ma 40 Ar ⁄ 39 Ar cooling ages obtainedfor muscovite <strong>and</strong> biotite imply a second distinct period<strong>of</strong> thermal reworking <strong>and</strong> isotopic resetting in theregion, correlated with the strong D 3 reworking in thesouth.Geophysical imagery <strong>of</strong> the subsurface shape <strong>of</strong> the Bruniacontinental promontoryIt is apparent that the thrust-related horizontal flow,which dominates the deformation <strong>of</strong> the Moldanubi<strong>and</strong>omain at around 325 Ma, does not exist in the Lugi<strong>and</strong>omain. This indicates a significant difference in thebulk exhumation processes between the two crustalsegments <strong>and</strong> the important involvement <strong>of</strong> the Bruniacontinent in the Moldanubian exhumation history. Inorder to discuss the relative contribution <strong>of</strong> the Bruniabasement in the tectonic evolution <strong>of</strong> the orogenic rootit is necessary to know the sub-surface extent <strong>of</strong> thebasement promontory underneath the orogenic rootrocks. Therefore, after the compilation <strong>of</strong> gravity data,a Bouguer anomaly map <strong>of</strong> the area (Fig. 10) wasproduced.At a large scale, the Bouguer anomaly map (Fig. 10)shows two main domains: (i) to the west, a domaincharacterized by low- <strong>and</strong> intermediate-gravity anomalies,inferred to be associated with rocks that have lowto intermediate densities; <strong>and</strong> (ii) to the east, a domainwith a succession <strong>of</strong> gravity highs inferred to be associatedwith significantly denser rocks. In the NE, thelow- to intermediate-gravity anomalies coincide withthe Lugian domain, whereas the gravity highs areassociated with the Brunia continent. The steep horizontalgradient in gravity reflects a steep boundarybetween the Lugian domain <strong>and</strong> the Brunia continent.In the SW, the Moldanubian domain west <strong>of</strong> theNNE–SSW-striking Central Moldanubian pluton ischaracterized by low- <strong>and</strong> intermediate-gravity anomaliessimilar to the Lugian domain. Similar to thenorth, we infer that the gravity highs over the Bruniacontinent in the SE represent mostly dense rocks.However, these gravity highs continue to the west, intothe eastern part <strong>of</strong> Moldanubian domain as far as theCentral Moldanubian pluton. Therefore, the observationmade in the north, where the gravity boundarycoincides with the geological boundary between theLugian domain <strong>and</strong> Brunia, is not valid in the south,where the geophysical boundary is located about 50–70 km west <strong>of</strong> the mapped geological boundary.Within the eastern Moldanubian domain, the rocksat outcrop are similar west <strong>and</strong> east <strong>of</strong> the SouthBohemian pluton, <strong>and</strong> rocks with high densities thatcould explain the gravity high in the eastern part <strong>of</strong> theMoldanubian domain are not represented at outcrop.Consequently, the dense rocks must be located at agreater depth. The gentle gradient <strong>of</strong> the gravity highsin the west indicates that the dense Brunia continentalpromontory dips towards the west, beneath the easternMoldanubian rocks. This interpretation is confirmedby preliminary 3D <strong>modelling</strong>, which shows that theboundary between the dense Brunia basement <strong>and</strong> thelow- to intermediate-density rocks <strong>of</strong> the Moldanubi<strong>and</strong>omain has a gentle dip down to a depth <strong>of</strong> some 2–3 km near the Central Moldanubian pluton, where itdips more steeply beneath the pluton (60–70°).Implication <strong>of</strong> Bouguer anomaly patternThis analysis <strong>of</strong> the Bouguer anomaly map in Fig. 10reveals striking differences between the Moldanubian<strong>and</strong> Lugian domains, which may be interpreted interms <strong>of</strong> the presence <strong>of</strong> Brunia basement under a thinlayer <strong>of</strong> Moldanubian rocks. This interpretation isconsistent with the suggestion that the S 3 fabric in theÓ 2007 Blackwell Publishing Ltd150


EXHUMATION IN LARGE HOT OROGEN 291OrdovicianBruniaMoldanubianBruniaLugian mafic lower SilesianCulm basincrust(a) A′ (b) B′CMPOMCOLCOLCCretaceous(c) basin C′CMPMg-K syeniteOMCThaya domeA-A′Southern Central NorthernBruniaMoldanubianMg-K syeniteOMCOLCOLCOMCUpper crustFig. 10. Interpretative cross-sections located in Fig. 2, showing distribution <strong>of</strong> the orogenic lower, middle <strong>and</strong> upper crust <strong>of</strong> theorogenic root <strong>and</strong> the Brunia continent. (a) East to west cross-section <strong>of</strong> southern Moldanubian showing synformal structure. TheS 3 fabric associated with channel flow is dominant in the orogenic lower crust, whereas in the orogenic middle crust the steep S 2fabric is dominant. Dismembered orogenic middle crust forms large-scale boudins surrounded by migmatitic S 3 fabric. (b) Crosssection<strong>of</strong> Lugian–Silesian domain, to show relationships between the D 2 extrusion <strong>of</strong> the orogenic lower crust that was thrust overthe Ordovician leptyno-amphibolite complex <strong>and</strong> the S 3 fabric associated with ductile thinning in the Lugian section. (c) North tosouth cross-section <strong>of</strong> the Moldanubian domain. Figure shows strong D 3 deformation <strong>of</strong> the rear part <strong>of</strong> the system with importantmixing <strong>of</strong> orogenic lower crust <strong>and</strong> orogenic middle crust in the channel flow fabric <strong>and</strong> weak D 3 deformation in the northernpart <strong>of</strong> the section.Moldanubian domain may result from underthrusting<strong>of</strong> the Brunia continental promontory. In contrast, theabsence <strong>of</strong> a gravity high over the Lugian domain,coupled with the <strong>structural</strong> <strong>and</strong> geochronological criteriadiscussed above, indicate that the Brunia continentwas not underthrust beneath the orogenic root <strong>of</strong>the Lugian domain. In other words, the horizontalfabrics in the Lugian domain cannot be explained bythe mechanical interaction between the Brunia continent<strong>and</strong> the Lugian root system.DISCUSSIONTectonic significance <strong>of</strong> steep orogenic fabricsThe zone <strong>of</strong> vertically foliated Lugian domain granulitesthat are surrounded by migmatized <strong>and</strong> highlysheared orthogneisses that also preserve a steep metamorphicS 2 fabric represents a key area for underst<strong>and</strong>ingthe early stage <strong>of</strong> exhumation <strong>of</strong> the lowerorogenic crust. Thrusting or normal faulting exhumationmechanisms cannot explain the steep S 2 fabric inthe granulite <strong>and</strong> orthogneisses, which we have interpretedinstead as recording a vertical channel alongwhich upward extrusion <strong>of</strong> the orogenic lower crustoccurred.Another important observation is that the S 2 foliationresults from macroscopic refolding <strong>of</strong> an earlysub-horizontal fabric at the scale <strong>of</strong> both the mesoscopicsub-horizontal compositional anisotropy <strong>and</strong>the orogenic crust (Sˇtípska´ et al., 2004; Racek et al.,2006). In addition, petrological studies show that theS 2 fabric in rocks <strong>of</strong> the orogenic middle crust adjacentto the granulites is associated with the progrademetamorphic evolution to a metamorphic peak ataround 9 kbar, whereas in the orogenic lower crustthe S 2 fabric is related with retrogression from highpressures around 18–20 kbar to pressures around7–10 kbar.The <strong>structural</strong> pattern <strong>and</strong> opposite metamorphicevolution developed in the kinematically similar S 2fabrics indicates vertical material transfer compatiblewith large-scale folding. We suggest that folding <strong>of</strong> thecrustal layers was responsible for the exhumation <strong>of</strong>the orogenic lower crust in cores <strong>of</strong> large antiforms <strong>and</strong>burial <strong>of</strong> the orogenic middle <strong>and</strong> upper crust in synformalregions. Furthermore, we suggest that thealternations <strong>of</strong> vertical belts <strong>of</strong> orogenic lower <strong>and</strong>middle crust that parallel the continental marginoriginated during this D 2 event. At the end <strong>of</strong> thisprocess most <strong>of</strong> the orogenic middle <strong>and</strong> lower crusthad already cooled below the blocking temperaturesfor the 40 Ar ⁄ 39 Ar system in hornblende <strong>and</strong> partly alsothe 40 Ar ⁄ 39 Ar system in muscovite.In the Polish part <strong>of</strong> the Lugian domain large-scalefolding <strong>of</strong> crustal layers was recognized by Don (1964)<strong>and</strong> Dumicz (1979). Moreover, a theoretical backgroundfor exhumation <strong>of</strong> lower crust by folding hasbeen provided by (Burg & Podladchikov, 1999) <strong>and</strong>applied to the exhumation <strong>of</strong> lower crustal rocks forÓ 2007 Blackwell Publishing Ltd151


292 K. SCHULMANN ET AL.example in the area <strong>of</strong> the Namche Barwa syntaxis(Burg et al., 1997).Based on shape analysis <strong>of</strong> macroscopic folds, Franeˇket al. (2006) proposed that the folding <strong>of</strong> lowercrustal layers occurred at granulite facies conditions bypassive amplification mechanisms. Consequently, theseauthors suggested that the folding started by thedevelopment <strong>of</strong> cuspate structures at the boundarybetween the lower <strong>and</strong> middle crust because <strong>of</strong> a lowviscositycontrast at this interface (cf. Kisters et al.,1996). Growth <strong>of</strong> the upward-pointing cusps wasreplaced by vertical extrusion as the whole domainbecame sufficiently shortened <strong>and</strong> the horizontalanisotropy was replaced by a vertical anisotropy. Atthat stage, the difference in the integrated vertical bulkviscosity between a weak lower crustal cusp <strong>and</strong> anadjacent stronger middle crustal lobe increased. Theseviscosity differences lead to strain partitioning <strong>and</strong>vertical extrusion <strong>of</strong> the weak orogenic lower crust butslower exhumation <strong>of</strong> adjacent middle crust (Jezˇeket al., 1998).To make the mechanism <strong>of</strong> vertical extrusion possiblea rigid floor is required (Schulmann et al., 2003),which is represented by a strong sub-root mantle in themodel proposed here. The strength <strong>of</strong> the sub-crustalmantle was modelled for a given thermal gradient byThompson et al. (2001) <strong>and</strong> Schulmann et al. (2002),who showed that at least 40 km <strong>of</strong> rigid mantle lithospherewas likely to have existed at the onset <strong>of</strong>exhumation <strong>of</strong> the thermally weakened lower crust.The occurrence <strong>of</strong> slivers <strong>of</strong> mantle peridotite, whichhave sharp boundaries <strong>and</strong> discordant internal fabricswith respect to the flow fabric <strong>of</strong> the surroundinggranulites, within the orogenic lower crust providesfurther evidence for the existence <strong>of</strong> a strong mantle<strong>and</strong> implies an important strength contrast between themantle <strong>and</strong> the orogenic lower crust (Medaris et al.,2006). Although the granulite extrusion structuresdescribed in this study resemble extrusions <strong>of</strong> lowercrustdriven by density inversion (Martinez et al.,2001), they were controlled mostly by lateral shorteningforces in this case, as shown in Fig. 11a.Horizontal flow <strong>of</strong> orogenic lower crustA transition from vertical S 2 fabrics to horizontal S 3fabrics occurs in almost all examples <strong>of</strong> lower crustalvertical structures in the Moldanubian <strong>and</strong> Lugi<strong>and</strong>omains. Therefore, the main questions that arise are:is there a causal relationship between vertical extrusion<strong>and</strong> the development <strong>of</strong> horizontal flow? <strong>and</strong> what arethe driving forces for the deformation?Ductile thinning <strong>and</strong> collapse <strong>of</strong> vertical fabric in the Lugi<strong>and</strong>omainA satisfactory tectonic model to explain the <strong>structural</strong>,petrological <strong>and</strong> geochronological data from theLugian domain is one involving folding <strong>of</strong> a layeredFig. 11. Bouguer anomaly map <strong>of</strong> the eastern margin <strong>of</strong> theBohemian Massif (provided by the Czech Geological Survey).Thick lines superimposed on the Bouguer anomaly map arelimits <strong>of</strong> geological unit boundaries <strong>and</strong> orogenic crustal levelsfrom Fig. 2. Br, Brunia continent; MDE, eastern branch <strong>of</strong> theMoldanubian domain; CMP, the Central Moldanubian pluton;MDW, western branch <strong>of</strong> the Moldanubian domain; LD, theLugian domain.orogenic crust followed by the asymmetrical northeastwardextrusion <strong>of</strong> the orogenic lower crust causedby the indentation <strong>of</strong> an Ordovician mafic lowercrustal block at c. 340 Ma. Extrusion <strong>of</strong> the orogeniclower crust was associated with the development <strong>of</strong> thehorizontal D 3 fabric accompanied with detachment <strong>of</strong>the western units <strong>and</strong> complete reworking <strong>of</strong> the orogenicmiddle crust in adjacent synforms. Consequently,the horizontal fabric in the Lugian domain cannothave been created in response to intracrustalflow because <strong>of</strong> lateral variations in lithostaticpressure ⁄ gravitational potential energy <strong>of</strong> thickenedcrust ⁄ lithosphere (Milnes & Koyi, 2000; V<strong>and</strong>erhaeghe& Teyssier, 2001).A more likely explanation is a model in which thevertically moving material experiences a reversal in theprincipal strain-rate directions (Feehan & Br<strong>and</strong>on,1999), which was expressed as a switch from the verticalfabric at depth to the sub-horizontal fabric atshallow crustal levels (Ring & Br<strong>and</strong>on, 1999). However,such a model requires the presence <strong>of</strong> a thickcontinental accretionary wedge, as proposed by Platt(1986), which has a mixed flow field involving verticalthickening at depth <strong>and</strong> vertical thinning near thesurface.Previously Schulmann & Gayer (2000) interpretedthe Lugian domain as an obliquely convergent wedgedeveloped above the Saxothuringian subduction zone.In this model, the Ordovician mafic lower crustal blockÓ 2007 Blackwell Publishing Ltd152


EXHUMATION IN LARGE HOT OROGEN 293represents a relict <strong>of</strong> the rigid backstop <strong>and</strong> the normal-senseshear zone in the west reflects the collapse <strong>of</strong>vertically extruded material (Fig. 11b). Importantly,the U–Pb zircon data show Ordovician protolith agesfor both the felsic orogenic root (orogenic lower <strong>and</strong>middle crust) <strong>and</strong> the Ordovician mafic lower crustalblock to the east. These data, together with the coincidence<strong>of</strong> the gravity <strong>and</strong> geological boundariesbetween the Lugian domain <strong>and</strong> the Brunia continent(Fig. 10), indicate that the transition from vertical tohorizontal fabric results from an intra-Lugianmechanical interaction <strong>and</strong> that the Brunia continentwas not involved in this process (Fig. 12).Heterogeneous channel flow <strong>and</strong> hot fold nappe fabricsin the Moldanubian domainThere are a number <strong>of</strong> differences associated with thedevelopment <strong>of</strong> the flat fabric in the Moldanubi<strong>and</strong>omain compared with the development <strong>of</strong> the flatfabric in the Lugian domain. In the southern part <strong>of</strong>the Moldanubian domain, the horizontal S 3 fabricdeveloped at decreasing pressure from south to north(Figs 4 & 5d) in conjunction with decreasing intensity<strong>of</strong> the D 3 reworking. This is associated with an increasein temperature <strong>and</strong> decrease in pressure in the orogenicmiddle crust concomitant with a decrease in temperature<strong>and</strong> pressure <strong>of</strong> the orogenic lower crust. Inaddition, migmatized gneisses <strong>of</strong> the orogenic lowercrust commonly surround blocks <strong>and</strong> boudins <strong>of</strong> theorogenic middle crust that preserve evidence <strong>of</strong> theearly HP metamorphism. The D 3 deformation showsconstant NNE–SSW-oriented stretching <strong>and</strong> top-tothe-NNEthrust-related shear movement, which isconsistent with oblique transpressive deformation inthe adjacent Moravo–Silesian Zone. The commongeometry <strong>and</strong> oblique thrust kinematics <strong>of</strong> the D 3deformation are the chief features <strong>of</strong> the mechanicalinteraction between the Moldanubian domain, Moravo–SilesianZone <strong>and</strong> the Brunia continent. Thegeophysical observations confirm the hypothesis <strong>of</strong>large-scale displacement <strong>of</strong> the Brunia basementunderneath the Moldanubian domain, which interpretationlocates the sub-surface margin <strong>of</strong> the Bruniacontinental promontory far to the west from thepresent Moldanubian domain <strong>and</strong> Moravo–SilesianZone boundary.This highly non-coaxial deformation correlates withyoung cooling ages (330–325 Ma) for muscovite <strong>and</strong>biotite 40 Ar ⁄ 39 Ar systems compared with older (350–340 Ma) hornblende 40 Ar ⁄ 39 Ar data, Sm–Nd coolingages from HP metamorphic rocks <strong>and</strong> metamorphicages from zircon, preserved mainly in the north.Muscovite <strong>and</strong> biotite from the adjacent Moravo–Silesian Zone nappes <strong>and</strong> the deformed Brunia continentshow similar young 40 Ar ⁄ 39 Ar cooling ages,which correspond to those <strong>of</strong> detrital muscovite fractionsfrom the Culm forel<strong>and</strong> basin. The detritalmuscovite shows convergence <strong>of</strong> the stratigraphic age<strong>of</strong> sedimentation <strong>and</strong> the 40 Ar ⁄ 39 Ar cooling ages at330–325 Ma (F. Neubauer, pers. comm.). Pebbles <strong>of</strong>granulite <strong>and</strong> Mg-rich syenite (durbachite) in theforel<strong>and</strong> conglomerates have yielded similar informationfrom U–Pb dating <strong>of</strong> zircon, where two groups <strong>of</strong>ages occur at c. 340 <strong>and</strong> c. 325 Ma, the latter beingconsistent with the stratigraphic age <strong>of</strong> 330–320 Ma(Kotkova´ et al., 2007).We suggest that the D 3 deformation, which involvedsubhorizontal top-to-the-NE thrusting <strong>of</strong> the Moldanubi<strong>and</strong>omain over the Brunia continent, occurred(a)(c)Fig. 12. Sequence <strong>of</strong> block diagrams toshow the principal exhumation mechanisms<strong>and</strong> progressive tectonic evolution <strong>of</strong> theeastern margin <strong>of</strong> the Bohemia Massif. (a)Early stage <strong>of</strong> crustal folding <strong>and</strong> extrusionassociated with vertical material <strong>and</strong> heattransfer. (b) Second stage (c. 340 Ma) <strong>of</strong>development <strong>of</strong> a flat fabric due to collapse<strong>of</strong> the vertical anisotropy – probably due toductile thinning mechanisms that accompanydetachments <strong>of</strong> the upper crust. (c)Weak orogenic root deformed by the Bruniacontinental promontory <strong>and</strong> the development<strong>of</strong> large hot nappes during D 3 at 330–325 Ma. The section shows a northwardattenuation <strong>of</strong> the crust together with anorthward-facing topographic slope withcoeval surface erosion in the North. Thelatter is consistent with the geometry <strong>of</strong> theBrunia continent.(b)Ó 2007 Blackwell Publishing Ltd153


294 K. SCHULMANN ET AL.at 330–325 Ma <strong>and</strong> was associated with exhumation <strong>of</strong>HP metamorphic rocks to the surface, as demonstratedby sedimentation <strong>of</strong> granulite pebbles <strong>and</strong> metamorphicmuscovite in the forel<strong>and</strong> basin. The earlier agesclustering around 340 Ma correspond to D 2 event, i.e.to the vertical material transfer associated with crustalscalefolding, evidence <strong>of</strong> which is partially preservedin flat-lying migmatites.The structure <strong>of</strong> the Moldanubian domain is consistentwith the Ôstage 3Õ <strong>of</strong> the channel-flow model – thehot fold nappe (Beaumont et al., 2006; Culshaw et al.,2006). In these channel-flow models, stage 3 coincideswith the arrival <strong>of</strong> a lower crustal block that forces weakmiddle <strong>and</strong> lower crust into large-scale gently inclinedfold nappes rooted in the thickened Moho.The hot fold nappe model can be applied to theMoldanubian root system, keeping in mind that themodel represents the result <strong>of</strong> a continuous 2D historywhereas the Moldanubian root system comprises anexhumation history in two stages that are kinematicallyindependent. Nevertheless, the hot fold nappemodel simulates well the relatively weakly deformedNE part <strong>of</strong> the Moldanubian system with well-preservedD 2 fabric, which reaches shallow crustal levels<strong>and</strong> ultimately the surface (Figs 8 & 11). To the south,the extruded nappes were derived from more internal<strong>and</strong> hotter parts <strong>of</strong> the orogen. Increasingly largevolumes <strong>of</strong> weak lower crust were gradually transportedsouthwards over the Brunia continent leadingto an increase in the amount <strong>of</strong> lower crustal materialat the surface.The channel flow is heterogeneous, as predicted byBeaumont et al. (2001, 2006), because <strong>of</strong> the previoushistory, which generates important crustal strengthvariations. The implication <strong>of</strong> this model is that theheterogeneous crust makes the geometry <strong>and</strong> composition<strong>of</strong> the channel flow similarly heterogeneous. Inthe Moldanubian case, the channel transportsdetached blocks <strong>and</strong> boudins <strong>of</strong> distinctly differentcompositions such as HP granulites <strong>and</strong> mid-crustalsegments that still preserve relicts <strong>of</strong> the D 2 fabric. Inthis southern part <strong>of</strong> the channel all isotopic systemsare re-equilibrated <strong>and</strong> the cooling ages are younger incomparison with the non-uniformly reset isotopicsystems in the north.CONCLUSIONSThis study has shown that exhumation <strong>of</strong> the orogeniclower crust in the eastern sector <strong>of</strong> the Variscan orogenicbelt is characterized by two independent stages.(1) The first stage is best preserved in the Lugi<strong>and</strong>omain, where it is characterized by an intra-crustalfolding that is responsible for vertical material transferassociated with exhumation <strong>of</strong> the deep orogenic lowercrust to shallower crustal levels corresponding topressures around 10 kbar at about 350 to 340 Ma. Thefolding <strong>and</strong> vertical extrusion events are followed by avertical shortening leading to development <strong>of</strong> subhorizontalfabrics at medium to low pressures. Theearly horizontal shortening was probably triggered bythe existence <strong>of</strong> a rigid Ordovician block, part <strong>of</strong> whichis preserved at the eastern boundary <strong>of</strong> the Lugi<strong>and</strong>omain. We suggest that this early exhumation eventwas related kinematically to Saxothuringian continentalsubduction to the east, creating a convergentcontinental accretionary wedge – the Lugian domain.Mechanical interactions with the large Brunia continentduring the exhumation process remain unconstrained.(2) The Moldanubian domain has steep fabrics insimilar orientations to those preserved in the Lugi<strong>and</strong>omain. Vertical material transfer during the LowerCarboniferous led to the development <strong>of</strong> alternations<strong>of</strong> steeply inclined domains <strong>of</strong> lower <strong>and</strong> middle orogeniccrust, similar to Lugian domain. However, thistectonic event was followed by a NNE-directed subhorizontalshearing at about 330–325 Ma resultingfrom subsurface indentation by the Brunia continentalpromontory into the Moldanubian domain. The arrival<strong>of</strong> the Brunia continent is responsible for the progressiveemplacement <strong>of</strong> hot fold nappes <strong>and</strong>heterogeneous channel flow in the rear (western) part<strong>of</strong> the system generating mixtures <strong>of</strong> middle <strong>and</strong> lowercrustal units with preserved early exhumation fabrics.ACKNOWLEDGEMENTSThis study was made possible thanks to the ANRproject ÔLFO in orogensÕ funding as well as to financialsupport <strong>of</strong> CNRS (UMRs 7516 <strong>and</strong> 7517) <strong>and</strong> theCzech Science Foundation (GACR 205 ⁄ 05 ⁄ 2187 <strong>and</strong>GACR 205 ⁄ 04 ⁄ 2065). Ondrej Lexa is indebted to theUniversite´ Louis Pasteur for covering his salary. Weare grateful to J. Platt, D. Grujic <strong>and</strong> an anonymousreviewer for constructive reviews. The editorial work<strong>of</strong> M. Brown is highly appreciated.REFERENCESAllem<strong>and</strong>, P. & Lardeaux, J. M., 1997. Strain partitioning <strong>and</strong>metamorphism in a deformable orogenic wedge: Applicationto the Alpine belt. Tectonophysics, 280, 157–169.Andersen, T. B., Jamtveit, B., Dewey, J. F. & Swensson, E.,1991. Subduction <strong>and</strong> eduction <strong>of</strong> continental crust: majormechanisms during continent–continent collision <strong>and</strong>orogenic extensional collapse, a model based on the southNorwegian Caledonides. Terra Nova, 3, 303–310.Baratoux, L., Schulmann, K., Ulrich, S. & Lexa, O., 2005.Contrasting microstructures <strong>and</strong> deformation mechanisms inmetagabbro mylonites contemporaneously deformed underdifferent temperatures (c. 650 °C <strong>and</strong> c. 750 °C). In: DeformationMechanisms, Rheology <strong>and</strong> Tectonics: From Minerals tothe Lithosphere, Special Publication 243 (eds Gapais, D., Brun,J.-P. & Cobbold, P. R.), pp. 97–125. Geological Society,London.Beard, B. L., Medaris, L. G., Johnson, C. M., Brueckner, H. K.&Mı´ sarˇ, Z., 1992. Petrogenesis <strong>of</strong> Variscan high-temperaturegroup–A eclogites from the Moldanubian zone <strong>of</strong> the BohemianMassif, Czechoslovakia. Contributions to Mineralogy <strong>and</strong>Petrology, 111, 468–483.Ó 2007 Blackwell Publishing Ltd154


EXHUMATION IN LARGE HOT OROGEN 295Beaumont, C., Jamieson, R. A., Nguyen, M. H. & Lee, B., 2001.Himalayan tectonics explained by extrusion <strong>of</strong> a low-viscositycrustal channel coupled to focused surface denudation.Nature, 414, 738–742.Beaumont, C., Nguyen, M. H., Jamieson, R. A. & Ellis, S., 2006.Crustal flow modes in large hot orogens. In: Channel Flow,Ductile Extrusion <strong>and</strong> Exhumation in Continental CollisionZones, Special Publication 268 (eds Law, R. D., Searle, M. P.& Godin, L.), pp. 91–145. Geological Society, London.Bro¨ cker, M. & Klemd, R., 1996. Ultrahigh-pressure metamorphismin the Snieznik Mountains (Sudetes, Pol<strong>and</strong>): P–Tconstraints <strong>and</strong> geological implications. Journal <strong>of</strong> Geology,104, 417–433.Brueckner, H. K., Medaris, L.G. & Bakun-Czubarow, N., 1991.Nd–Sm age <strong>and</strong> isotope patterns from Variscan eclogites <strong>of</strong>the eastern Bohemian Massif. Neues Jahrbuch fu¨r MineralogieAbh<strong>and</strong>lungen, 163, 169–196.Burg, J.-P., 1999. Ductile structures <strong>and</strong> instabilities: theirimplication for Variscan tectonics in the Ardennes. Tectonophysics,309, 1–25.Burg, J.-P. & Podladchikov, Y., 1999. Lithospheric scale folding:<strong>numerical</strong> <strong>modelling</strong> <strong>and</strong> application to the Himalayan syntaxes.International Journal <strong>of</strong> Earth Sciences, 88, 190–200.Burg, J. P., Davy, P., Nievergelt, P. et al., 1997. Exhumationduring crustal folding in the Namche-Barwa syntaxis. TerraNova, 9, 53–56.Carswell, D. A. & O¢Brien, P. J., 1993. Thermobarometry <strong>and</strong>geotectonic significance <strong>of</strong> high-pressure granulites: examplesfrom the Moldanubian Zone <strong>of</strong> the Bohemian Massif inLower Austria. Journal <strong>of</strong> Petrology, 34, 427–459.Chemenda, A. I., Mattauer, M., Malavieille, J. & Bokun, A. N.,1995. A mechanism for syn-collisional rock exhumation <strong>and</strong>associated normal faulting: results from physical <strong>modelling</strong>.Earth & Planetary Science Letters, 132, 225–232.Culshaw, N.G., Beaumont, C. & Jamieson, R.A., 2006. Theorogenic superstructure–infrastructure concept: revisited,quantified, <strong>and</strong> revived. Geology, 34, 733–736. doi:10.1130/G22793.1.Dallmeyer, R. D., Neubauer, F. & Hock, V., 1992. Chronology<strong>of</strong> late Paleozoic tectonothermal activity in the southeasternBohemian Massif, Austria (Moldanubian <strong>and</strong> Moravo–Silesianzones): 40 Ar ⁄ 39 Ar mineral age controls. Tectonophysics,210, 135–153.Dewey, J. F., C<strong>and</strong>e, S. & Pitman, W. C., 1989. Tectonic evolution<strong>of</strong> the India ⁄ Eurasia Collision Zone. Eclogae GeologicaeHelvetiae, 82, 717–734.Don, J., 1964. The Zlote <strong>and</strong> Krowiarki Mountains as <strong>structural</strong>elements <strong>of</strong> the Snieznik metamorphic massif. GeologiaSudetica, 1, 79–117.Duchene, S., Lardeaux, J. M. & Albarede, F., 1997. Exhumation<strong>of</strong> eclogites: insights from depth–time path analysis. Tectonophysics,280, 125–140.Dudek, A., 1980. The crystalline basement block <strong>of</strong> the OuterCarpathians in Moravia. Rozpravy Cˇeskoslovenske´ AkademieVeˇd, 90, 1–85.Dudek, A. & Fediukova´ , E., 1974. Eclogites <strong>of</strong> the BohemianMoldanubicum. Neues Jahrbbuch fu¨r Mineralogie Abh<strong>and</strong>lungen,121, 127–159.Dumicz, M., 1979. Tectogenesis <strong>of</strong> the metamorphosed series <strong>of</strong>the Klodzko district: a tentative explanation. GeologicaSudetica, 14, 29–44.Edel, J. B., Schulmann, K. & Holub, F. V., 2003. Anticlockwise<strong>and</strong> clockwise rotations <strong>of</strong> the Eastern Variscides accommodatedby dextral lithospheric wrenching: palaeomagnetic <strong>and</strong><strong>structural</strong> evidence. Journal <strong>of</strong> the Geological Society, 160,209–218.Engl<strong>and</strong>, P. C. & Houseman, G. A., 1988. The mechanics <strong>of</strong> theTibetan Plateau. Philosophical Transactions – Royal Society <strong>of</strong>London, Series A, 326, 301–320.Feehan, J. G. & Br<strong>and</strong>on, M. T., 1999. Contribution <strong>of</strong> ductileflow to exhumation <strong>of</strong> low-temperature, high-pressure metamorphicrocks: San Juan–Cascade nappes, NW WashingtonState. Journal <strong>of</strong> Geophysical Research B: Solid Earth, 104,10 883–10 902.Finger, F., Tichomirowa, M., Pin, C. & Hanzˇl, P., 2000. Relics<strong>of</strong> an early Panafrican metabasite–metarhyolite formation inthe Brno Massif, Moravia, Czech Republic. InternationalJournal <strong>of</strong> Earth Sciences, 89, 328–335.Francu˚, E., Francu˚, J., Kalvoda, J., Poelchau, H.S. & Otava, J.,2002. Burial <strong>and</strong> uplift history <strong>of</strong> the Palaeozoic Flysch in theVariscan forel<strong>and</strong> basin (SE Bohemian Massif, CzechRepublic). In: Continental Collision <strong>and</strong> the TectonosedimentaryEvolution <strong>of</strong> Forel<strong>and</strong>s, Stephen Mueller Special PublicationSeries (eds Bertotti, G., Schulmann, K. & Cloetingh, S.),pp. 167–179. European Geophysical Society, Berlin.Franeˇk, J., Schulman, K. & Lexa, O., 2006. Kinematic <strong>and</strong>rheological model <strong>of</strong> exhumation <strong>of</strong> high pressure granulites inthe Variscan orogenic root: example <strong>of</strong> the Blanský les granulite,Bohemian Massif, Czech Republic. Mineralogy <strong>and</strong>Petrology, 86, 253–276.Franke, W., 2000. The mid-European segment <strong>of</strong> the Variscides:tectonostratigraphic units, terrane boundaries <strong>and</strong> plate tectonicevolution. In: Orogenic Processes: Quatification <strong>and</strong>Modelling in the Variscan Belt, Special Publication 179 (edsFranke, W., Haak, W., Oncken, O. & Tanner, D.), pp. 35–63.Geological Society, London.Friedl, G., Finger, F., Paquette, J.L., von Quadt, A.,McNaughton, N. J. & Fletcher, I. R., 2004. Pre-Variscangeological events in the Austrian part <strong>of</strong> the Bohemian Massifdeduced from U–Pb zircon ages. International Journal <strong>of</strong> EarthSciences, 93, 802–823.Fritz, H., Dallmeyer, R. D. & Neubauer, F., 1996. Thick-skinnedversus thin-skinned thrusting: rheology controlled thrustpropagation in the Variscan collisional belt (the southeasternBohemian Massif, Czech Republic–Austria). Tectonics, 15,1389–1413.Fuchs, G., 1986. Zur Diskussion um den Deckenbau der Bo¨ h-mischen Masse. Jahrbuch der Geologischen Bundesanstalt, 129,41–49.Gerya, T. & Stockhert, B., 2006. Two-dimensional <strong>numerical</strong>modeling <strong>of</strong> tectonic <strong>and</strong> metamorphic histories at activecontinental margins. International Journal <strong>of</strong> Earth Sciences,95, 250–274.Godin, L., Grujic, D., Law, R.D. & Searle, M.P., 2006. Channelflow, ductile extrusion <strong>and</strong> exhumation in continental collisionzones: an introductionIn: Channel Flow, Ductile Extrusion <strong>and</strong>Exhumation in Continental Collision Zones, Special Publication268 (eds Law, R. D., Searle, M. P. & Godin, L.), pp. 1–23.Geological Society, London.Grujic, D., Casey, M., Davidson, C. et al., 1996. Ductile extrusion<strong>of</strong> the Higher Himalayan Crystalline in Bhutan: evidencefrom quartz micr<strong>of</strong>abrics. Tectonophysics, 260, 21–43.Harrison, T. M., Duncan, I. & McDougall, I., 1985. Diffusion <strong>of</strong>40 Ar in biotite; temperature, pressure <strong>and</strong> compositionaleffects. Geochimica et Cosmochimica Acta, 49, 2461–2468.Hartley, A. J. & Otava, J., 2001. Sediment provenance <strong>and</strong> dispersalin a deep marine forel<strong>and</strong> basin: The Lower CarboniferousCulm Basin, Czech Republic. Journal <strong>of</strong> the GeologicalSociety, 158, 137–150.Hasalova´ , P., Schulmann, K., Lexa, O. et al., 2008a. Origin <strong>of</strong>migmatites by deformation enhanced melt infiltration <strong>of</strong>orthogneiss: a new model based on quantitative micro<strong>structural</strong>analysis. Journal <strong>of</strong> Metamorphic Geology, 26, 29–53.Hasalova´ , P., Sˇtı´ pska´ , P., Powell, R., Schulmann, K., Janousˇek,V. & Lexa, O., 2008b. Transforming mylonitic metagranite byopen-system interactions during melt flow. Journal <strong>of</strong> MetamorphicGeology, 26, 55–80.Hensen, B.J. & Zhou, B., 1995. Retention <strong>of</strong> isotopic memory ingarnets partially broken down during an overprinting granulite-faciesmetamorphism: implications for the Sm–Nd closuretemperature. Geology, 23, 225–228.Holub, F. V., Cocherie, A. & Rossi, P., 1997. Radiometric dating<strong>of</strong> granitic rocks from the Central Bohemian Plutonic Complex(Czech Republic): constraints on the chronology <strong>of</strong>Ó 2007 Blackwell Publishing Ltd155


296 K. SCHULMANN ET AL.thermal <strong>and</strong> tectonic events along the Moldanubian-Barr<strong>and</strong>ianboundary. Comptes Rendus de lÕAcademie de Sciences –Serie IIa: Sciences de la Terre et des Planetes, 325, 19–26.Jamieson, R. A., Beaumont, C., Nguyen, M. H. & Lee, B., 2002.Interaction <strong>of</strong> metamorphism, deformation <strong>and</strong> exhumation inlarge convergent orogens. Journal <strong>of</strong> Metamorphic Geology,20, 9–24.Jamieson, R. A., Beaumont, C., Medvedev, S. & Nguyen, M. H.,2004. Crustal channel flows: 2. Numerical models withimplications for metamorphism in the Himalayan–Tibetanorogen. Journal <strong>of</strong> Geophysical Research B: Solid Earth, 109,B06407, 1–24.Janousˇek, V. & Holub, F. V., 2007. The causal link between HP–HT metamorphism <strong>and</strong> ultrapotassic magmatism in collisionalorogens: case study from the Moldanubian Zone <strong>of</strong> theBohemian Massif. Proceedings <strong>of</strong> the Geologists Association,118, 75–86.Jastrzębski, M., 2005. The tectonometamorphic evolution <strong>of</strong> themarbles in the Lądek–S´ niez_ nik Metamorphic Unit, WestSudetes. Geologica Sudetica, 37, 1–26.Jezˇek, J., Schulmann, K. & Thompson, A. B., 1998. Modelling <strong>of</strong>strain partitioning in transpression zones. In: Proceedings <strong>of</strong>the Fourth Annual Conference <strong>of</strong> the International Associationfor Mathematical Geology (eds Buccianti, A., Nardi, G. &Potenza, R.), pp. 827–832. De Frede, Naples.Kisters, A. F. M., Charlesworth, E. G., Gibson, R. L. &Anhaeusser, C. R., 1996. Steep structure formation in theOkiep copper district, South Africa: bulk inhomogeneousshortening <strong>of</strong> a high-grade metamorphic granite–gneisssequence. Journal <strong>of</strong> Structural Geology, 18, 735–751.Klemd, R. & Bro¨ cker, M., 1999. Fluid influence on mineralreactions in ultrahigh-pressure granulites: a case study in theSnieznik Mts (West Sudetes, Pol<strong>and</strong>). Contributions to Mineralogy<strong>and</strong> Petrology, 136, 358–373.Kolenovska´ , E., Schulmann, K., Kla´ pova´ , H. & Sˇtı´ pska´ , P.,1999. Tectonometamorphic evolution <strong>of</strong> the Moldanubianzone near Jemnice (south Moravia, Bohemian Massif).Beihefte zum European Journal <strong>of</strong> Mineralogy, 11, 91–110.Konopa´ sek, J. & Schulmann, K., 2005. Contrasting Early Carboniferousfield geotherms: evidence for accretion <strong>of</strong> a thickenedorogenic root <strong>and</strong> subducted Saxothuringian crust(Central European Variscides). Journal <strong>of</strong> the GeologicalSociety, 162, 463–470.Konopa´ sek, J., Schulmann, K. & Johan, V., 2002. Eclogite-faciesmetamorphism at the eastern margin <strong>of</strong> the Bohemian Massif– subduction prior to continental underthrusting? EuropeanJournal <strong>of</strong> Mineralogy, 14, 701–713.Kotkova´ , J., Gerdes, A., Parrish, R.R. & Nova´ k, M., 2007.Clasts <strong>of</strong> Variscan high-grade rocks within Upper Vise´ anconglomerates – constraints on exhumation history frompetrology <strong>and</strong> U–Pb chronology. Journal <strong>of</strong> MetamorphicGeology, 25, 781–801.Kro¨ ner, A., Sˇtı´ pska´ , P., Schulmann, K. & Jaeckel, P., 2000.Geochronological constraints on the pre-Variscan evolution <strong>of</strong>the northeastern margin <strong>of</strong> the Bohemian Massif, CzechRepublic. In: Quantification <strong>and</strong> Modelling in the VariscanBelt, Special Publication 179 (eds Franke, W.H.V., Oncken, O.& Tanner, D.), pp. 175–197. Geological Society, London.Kro¨ ner, A., Jaeckel, P. & Hegner, E., 2001. Single zircon ages<strong>and</strong> whole-rock Nd isotopic systematics <strong>of</strong> Early Palaeozoicgranitoid gneissess from the Czech <strong>and</strong> Polish Sudetes.(Jizerske´ Hory, Krkonosˇe mountains <strong>and</strong> Orlice-Sneˇzˇnı´ kcomplex). International Journal <strong>of</strong> Earth Sciences, 90, 304–324.Kryza, R., Pin, C. & Vielzeuf, D., 1996. High-pressure granulitesfrom the Sudetes (south-west Pol<strong>and</strong>): evidence <strong>of</strong> crustalsubduction <strong>and</strong> collisional thickening in the Variscan Belt.Journal <strong>of</strong> Metamorphic Geology, 14, 531–546.Lange, U., Bro¨ cker, M., Armstrong, R., Trapp, E. & Mezger,K., 2005. Sm–Nd <strong>and</strong> U–Pb dating <strong>of</strong> high-pressure granulitesfrom the Zlote <strong>and</strong> Rychleby Mts (Bohemian Massif, Pol<strong>and</strong><strong>and</strong> Czech Republic). Journal <strong>of</strong> Metamorphic Geology, 23,133–145.Lexa, O., Sˇtı´ pska´ , P., Schulmann, K., Baratoux, L. & Kro¨ ner,A., 2005. Contrasting textural record <strong>of</strong> two distinct metamorphicevents <strong>of</strong> similar P–T conditions <strong>and</strong> different durations.Journal <strong>of</strong> Metamorphic Geology, 23, 649–666.Macintyre, R. M., Bowes, D.R., Hamidullah, S. & Onscott,T.C., 1992. K–Ar <strong>and</strong> Ar–Ar study <strong>of</strong> amphibolites frommeta-ophiolite complexes, eastern Bohemian Massif. In: Proceedings<strong>of</strong> 1st International Conference on the BohemianMassif, Czech Geological Survey, Prague, pp. 195–199.Martinez, F., Goodliffe, A. M. & Taylor, B., 2001. Metamorphiccore complex formation by density inversion <strong>and</strong> lower crustextrusion. Nature, 411, 930–934.Matte, P., Maluski, H., Rajlich, P. & Franke, W., 1990. Terraneboundaries in the Bohemian Massif: result <strong>of</strong> large-scaleVariscan shearing. Tectonophysics, 177, 151–170.Mazur, S., Aleks<strong>and</strong>rowski, P. & Szczepanski, J., 2005. Thepresumed Tepla´ –Barr<strong>and</strong>ian ⁄ Moldanubian terrane boundaryin the Orlica Mountains (Sudetes, Bohemian Massif): <strong>structural</strong><strong>and</strong> petrological characteristics. Lithos, 82, 85–112.Medaris, G., Jelínek, E. & Mı´ sarˇ, Z., 1995. Czech eclogites:terrane settings <strong>and</strong> implications for Variscan tectonic evolution<strong>of</strong> the Bohemian Massif. European Journal <strong>of</strong> Mineralogy,7, 7–28.Medaris, L. G., Fournelle, J. H., Ghent, E. D., Jelínek, E. &Mı´ sarˇ, Z., 1998. Prograde eclogite in the Gfo¨ hl Nappe, CzechRepublic: new evidence on Variscan high-pressure metamorphism.Journal <strong>of</strong> Metamorphic Geology, 16, 563–576.Medaris, L. G., Beard, B. L. & Jelínek, E., 2006. Mantle-derived,UHP garnet pyroxenite <strong>and</strong> eclogite in the MoldanubianGfo¨ hl nappe, Bohemian Massif: a geochemical review, new P–T determinations, <strong>and</strong> tectonic interpretation. InternationalGeology Review, 48, 765–777.Milnes, A. G. & Koyi, H. A., 2000. Ductile rebound <strong>of</strong> anorogenic root: case study <strong>and</strong> <strong>numerical</strong> model <strong>of</strong> gravitytectonics in the Western Gneiss Complex, Caledonides,southern Norway. Terra Nova, 12, 1–7.Murtezi, M., 2006. The acid metavolcanic rocks <strong>of</strong> the Orlica–S´ niez_ nik Dome (Sudetes): their origin <strong>and</strong> tectono-metamorphicevolution. Geologica Sudetica, 38, 1–38.OÕBrien, P. J. & Carswell, D. A., 1993. Tectonometamorphicevolution <strong>of</strong> the Bohemian Massif: evidence from highpressure metamorphic rocks. Geologische Rundschau, 82,531–555.OÕBrien, P. J. & Ro¨ tzler, J., 2003. High-pressure granulites:formation, recovery <strong>of</strong> peak conditions <strong>and</strong> implications fortectonics. Journal <strong>of</strong> Metamorphic Geology, 21, 3–20.Parry, M., Sˇtı´ pska´ , P., Schulmann, K., Hrouda, F., Ježek, J. &Kro¨ ner, A., 1997. Tonalite sill emplacement at an obliqueplate boundary: northeastern margin <strong>of</strong> the Bohemian Massif.Tectonophysics, 280, 61–81.Petrakakis, K., 1997. Evolution <strong>of</strong> Moldanubian rocks in Austria:review <strong>and</strong> synthesis. Journal <strong>of</strong> Metamorphic Geology,15, 203–222.Pitra, P. & Guiraud, M., 1996. Probable anticlockwise P–Tevolution in extending crust: Hlinsko region, BohemianMassif. Journal <strong>of</strong> Metamorphic Geology, 14, 49–60.Pitra, P., Burg, J.-P., Schulmann, K. & Ledru, P., 1994. Lateorogenic extension in the Bohemian Massif: petro<strong>structural</strong>evidence in the Hlinsko region. Geodinamica Acta, 7, 15–30.Platt, J. P., 1986. Dynamics <strong>of</strong> orogenic wedges <strong>and</strong> the uplift <strong>of</strong>high-pressure metamorphic rocks. Geological Society <strong>of</strong>America Bulletin, 97, 1037–1053.Platt, J. P., 1993. Exhumation <strong>of</strong> high-pressure rocks: a review <strong>of</strong>concepts <strong>and</strong> processes. Terra Nova, 5, 119–133.Racek, M., Sˇtı´ pska´ , P., Pitra, P., Schulmann, K. & Lexa, O.,2006. Metamorphic record <strong>of</strong> burial <strong>and</strong> exhumation <strong>of</strong> orogeniclower <strong>and</strong> middle crust: a new tectonothermal model forthe Drosendorf window (Bohemian Massif, Austria). Mineralogy<strong>and</strong> Petrology, 86, 221–251.Rey, P., V<strong>and</strong>erhaeghe, O. & Teyssier, C., 2001. Gravitationalcollapse <strong>of</strong> the continental crust: definition, regimes <strong>and</strong>modes. Tectonophysics, 342, 435–449.Ó 2007 Blackwell Publishing Ltd156


EXHUMATION IN LARGE HOT OROGEN 297Ring, U. & Br<strong>and</strong>on, M. T., 1999. Ductile deformation <strong>and</strong> massloss in the Franciscan Subduction Complex: implications forexhumation processes in accretionary wedges. In: ExhumationProcesses: Normal Faulting, Ductile Flow <strong>and</strong> Erosion, SpecialPublication 154 (ed. Ring, U., Br<strong>and</strong>on, M.T., Lister, G.S. &Willet, S.D.), pp. 55–86. Geological Society, London.Romanova´ ,M.&Sˇtı´ pska´ , P., 2001. Structural <strong>and</strong> metamorphicevolution <strong>of</strong> the Stronie formation near Javorník. MineralogicalSociety <strong>of</strong> Pol<strong>and</strong>, Special Papers, 19, 147–149.Schneider, D. A., Zahniser, S. J., Clascock, J. M., Gordon, S. M.& Manecki, M., 2006. Thermochronology <strong>of</strong> the West Sudetes(Bohemian Massif): rapid <strong>and</strong> repeated eduction in the easternVariscides, Pol<strong>and</strong> <strong>and</strong> Czech Republic. American Journal <strong>of</strong>Science, 306, 846–873.Schulmann, K., 1990. Fabric <strong>and</strong> kinematic study <strong>of</strong> the Bitesorthogneiss (southwestern Moravia): result <strong>of</strong> large-scalenortheastward shearing parallel to the Moldanubian⁄ Moravian boundary. Tectonophysics, 177, 229–244.Schulmann, K. & Gayer, R., 2000. A model for a continentalaccretionary wedge developed by oblique collision: The NEBohemian Massif. Journal <strong>of</strong> the Geological Society, 157, 401–416.Schulmann, K., Ledru, P., Autran, A. et al., 1991. Evolution <strong>of</strong>nappes in the eastern margin <strong>of</strong> the Bohemian Massif: akinematic interpretation. Geologische Rundschau, 80, 73–92.Schulmann, K., Melka, R., Lobkowicz, M. Z., Ledru, P., Lardeaux,J.-M. & Autran, A., 1994. Contrasting styles <strong>of</strong>deformation during progressive nappe stacking at the southeasternmargin <strong>of</strong> the Bohemian Massif (Thaya Dome).Journal <strong>of</strong> Structural Geology, 16, 355–370.Schulmann, K., Schaltegger, U., Ježek, J., Thompson, A.B. &Edel, J.-B., 2002. Rapid burial <strong>and</strong> exhumation during orogeny:thickening <strong>and</strong> synconvergent exhumation <strong>of</strong> thermallyweakened <strong>and</strong> thinned crust (Variscan orogen in WesternEurope). American Journal <strong>of</strong> Science, 302, p 856–879.Schulmann, K., Thompson, A. B., Lexa, O. & Ježek, J., 2003.Strain distribution <strong>and</strong> fabric development modeled in active<strong>and</strong> ancient transpressive zones. Journal <strong>of</strong> GeophysicalResearch–Solid Earth, 108(B1), 2023, doi: 10.1029/200/JB000632./.Schulmann, K., Kroner, A., Hegner, E. et al., 2005. Chronologicalconstraints on the pre-orogenic history, burial <strong>and</strong>exhumation <strong>of</strong> deep-seated rocks along the eastern margin <strong>of</strong>the Variscan orogen, Bohemian Massif, Czech Republic.American Journal <strong>of</strong> Science, 305, 407–448.Steltenpohl, M. G., Cymerman, Z., Krogh, E. J. & Kunk, M. J.,1993. Exhumation <strong>of</strong> eclogitized continental basement duringVariscan lithospheric delamination <strong>and</strong> gravitational collapse,Sudety Mountains, Pol<strong>and</strong>. Geology, 21, 1111–1114.Sˇtı´ pska´ , P. & Powell, R., 2005a. Constraining the P–T path <strong>of</strong> aMORB-type eclogite using pseudosections, garnet zoning <strong>and</strong>garnet–clinopyroxene thermometry: an example from theBohemian Massif. Journal <strong>of</strong> Metamorphic Geology, 23, 725–743.Sˇtı´ pska´ , P. & Powell, R., 2005b. Does ternary feldspar constrainthe metamorphic conditions <strong>of</strong> high-grade meta-igneousrocks? Evidence from orthopyroxene granulites, BohemianMassif Journal <strong>of</strong> Metamorphic Geology, 23, 627–647.Sˇtı´ pská, P. & Schulmann, K., 1995. Inverted metamorphiczonation in a basement-derived nappe sequence, easternmargin <strong>of</strong> the Bohemian Massif. Geological Journal, 30, 385–413.Sˇtı´ pská, P., Schulmann, K. & Ho¨ ck, V., 2000. Complex metamorphiczonation <strong>of</strong> the Thaya dome: result <strong>of</strong> buckling <strong>and</strong>gravitational collapse <strong>of</strong> an imbricated nappe sequence. In:Forced Folds <strong>and</strong> Fractures, Special Publication, 169 (edsCosgrove, J. W. & Ameen, M. S.), pp. 197–211. GeologicalSociety, London.Sˇtı´ pská, P., Schulmann, K., Thompson, A. B., Ježek, J. & Kro¨ ner,A., 2001. Thermomechanical role <strong>of</strong> a Cambro–Ordovicianpaleorift during the Variscan collision: the NE margin <strong>of</strong> theBohemian Massif. Tectonophysics, 332, 239–253.Sˇtı´ pská, P., Schulmann, K. & Kro¨ ner, A., 2004. Vertical extrusion<strong>and</strong> middle crustal spreading <strong>of</strong> omphacite granulite: amodel <strong>of</strong> syn-convergent exhumation (Bohemian Massif,Czech Republic). Journal <strong>of</strong> Metamorphic Geology, 22, 179–198.Sˇtı´ pská, P., Pitra, P. & Powell, R., 2006. Separate or sharedmetamorphic histories <strong>of</strong> eclogites <strong>and</strong> surrounding rocks? Anexample from the Bohemian Massif Journal <strong>of</strong> MetamorphicGeology, 24, 219–240.Suess, F. E., 1912. Die moravischen Fenster und ihre Beziehungzum Grundgebirge des Hohen Gesenke. Denkschrichte desKoeniglischen Akademie der Wissenscahft, Mathematik.Naturwissenschaft, 12, 541–631.Suess, F. E., 1926. Intrusionstektonik und W<strong>and</strong>ertektonik imvariszichen Grundgebirge. Borntträger, Berlin.Tajčmanova´ , L., Konopa´ sek, J. & Schulmann, K., 2006. Thermalevolution <strong>of</strong> the orogenic lower crust during exhumationwithin a thickened Moldanubian root <strong>of</strong> the Variscan belt <strong>of</strong>Central Europe. Journal <strong>of</strong> Metamorphic Geology, 24, 119–134.Thompson, A. B., Schulmann, K., Jezˇek, J. & Tolar, V., 2001.Thermally s<strong>of</strong>tened continental extensional zones (arcs <strong>and</strong>rifts) as precursors to thickened orogenic belts. Tectonophysics,332, 115–141.Tollmann, A., 1982. Großraumiger variszischer Deckenbau imMoldanubikum und neue Gedanken zum Variszikum Europas.Geotektektonische Forschungen, 64, 1–91.Urban, M., 1992. Kinematics <strong>of</strong> the Variscan thrusting in theEastern Moldanubicum (Bohemian Massif, Czechoslovakia):evidence from the Namest granulite massif. Tectonophysics,201, 371–391.Van Breemen, O., Aftalion, M., Bowes, D. R. et al., 1982.Geochronological studies <strong>of</strong> the Bohemian massif, Czechoslovakia,<strong>and</strong> their significance in the evolution <strong>of</strong> CentralEurope. Transactions <strong>of</strong> the Royal Society <strong>of</strong> Edinburgh: EarthSciences, 73, 89–108.V<strong>and</strong>erhaeghe, O. & Teyssier, C., 2001. Partial melting <strong>and</strong> flow<strong>of</strong> orogens. Tectonophysics, 342, 451–472.Ziegler, P. A., 1986. Geodynamic model for the Palaeozoiccrustal consolidation <strong>of</strong> western <strong>and</strong> central Europe. Tectonophysics,126, 303–328.Received 3 April 2007; revision accepted 27 November 2007.Ó 2007 Blackwell Publishing Ltd157


Author's personal copyC. R. Geoscience 341 (2009) 266–286TectonicsAn Andean type Palaeozoic convergence in the Bohemian MassifKarel Schulmann a, *, Jiří Konopásek b,c , Vojtĕch Janoušek b , Ondrej Lexa c ,Jean-Marc Lardeaux d , Jean-Bernard Edel a , Pavla Štípská a , Stanislav Ulrich ea EOST, UMR 7517, université Louis-Pasteur, 1, rue Blessig, 67084 Strasbourg, Franceb Czech Geological Survey, Klárov 3, 118 21 Praha 1, Czech Republicc Institute <strong>of</strong> Petrology <strong>and</strong> Structural geology, Charles University, Albertov 6, 128 43 Prague, Czech Republicd Géosciences Azur, UMR CNRS 6526, université de Nice Sophia-Antipolis, parc Valrose, 06108 Nice cedex 02, Francee Geophysical Institute, Czech Academy <strong>of</strong> Sciences, Boční II/1401, 141 31 Praha 4, Czech RepublicReceived 24 December 2008; accepted after revision 24 December 2008Available online 25 February 2009Written on invitation <strong>of</strong> the Editorial BoardAbstractThe geological inventory <strong>of</strong> the Variscan Bohemian Massif can be summarized as a result <strong>of</strong> Early Devonian subduction <strong>of</strong> theSaxothuringian ocean <strong>of</strong> unknown size underneath the eastern continental plate represented by the present-day Teplá-Barr<strong>and</strong>ian <strong>and</strong>Moldanubian domains. During mid-Devonian, the Saxothuringian passive margin sequences <strong>and</strong> relics <strong>of</strong> Ordovician oceanic crusthave been obducted over the Saxothuringian basement in conjunction with extrusion <strong>of</strong> the Teplá-Barr<strong>and</strong>ian middle crust along the socalledTeplá suture zone. This event was connected with the development <strong>of</strong> the magmatic arc further east, together with a fore-arc basinon the Teplá-Barr<strong>and</strong>ian crust. The back-arc region – the future Moldanubian zone – was affected by lithospheric thinning whichmarginally affected also the eastern Brunia continental crust. The subduction stage was followed by a collisional event caused by thearrival <strong>of</strong> the Saxothuringian continental crust that was associated with crustal thickening <strong>and</strong> the development <strong>of</strong> the orogenic rootsystem in the magmatic arc <strong>and</strong> back-arc region <strong>of</strong> the orogen. The thickening was associated with depression <strong>of</strong> the Moho <strong>and</strong> the flux<strong>of</strong> the Saxothuringian felsic crust into the root area. Originally subhorizontal anisotropy in the root zone was subsequently folded bycrustal-scale cusp folds in front <strong>of</strong> the Brunia backstop. During the Visean, the Brunia continent indented the thickened crustal root,resulting in the root’s massive shortening causing vertical extrusion <strong>of</strong> the orogenic lower crust, which changed to a horizontal viscouschannel flow <strong>of</strong> extruded lower crustal material in the mid- to supra-crustal levels. Hot orogenic lower crustal rocks were extruded: (1)in a narrow channel parallel to the former Teplá suture surface; (2) in the central part <strong>of</strong> the root zone in the form <strong>of</strong> large scale antiformalstructure; <strong>and</strong> (3) in form <strong>of</strong> hot fold nappe over the Brunia promontory, where it produced Barrovian metamorphism <strong>and</strong> subsequentimbrications <strong>of</strong> its upper part. The extruded deeper parts <strong>of</strong> the orogenic root reached the surface, which soon thereafter resulted in thesedimentation <strong>of</strong> lower-crustal rocks pebbles in the thick forel<strong>and</strong> Culm basin on the stable part <strong>of</strong> the Brunia continent. Finally, duringthe Westfalian, the forel<strong>and</strong> Culm wedge was involved into imbricated nappe stack together with basement <strong>and</strong> orogenic channel flownappes. To cite this article: K. Schulmann et al., C. R. Geoscience 341 (2009).# 2009 Published by Elsevier Masson SAS on behalf <strong>of</strong> Académie des sciences.RésuméConvergence paléozoïque de type Andin dans le Massif de Bohême. Le Massif varisque de Bohême est le résultat de lasubduction, au Dévonien supérieur, de l’océan Saxothuringien sous la plaque continentale représentée à l’est par les zones actuelles* Corresponding author.E-mail address: schulmann.karel@gmail.com (K. Schulmann).1631-0713/$ – see front matter # 2009 Published by Elsevier Masson SAS on behalf <strong>of</strong> Académie des sciences.doi:10.1016/j.crte.2008.12.006159


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 267de Teplá-Barr<strong>and</strong>ien et de Moldanubien. L’ampleur de cet océan demeure inconnue. Pendant le Dévonien moyen, les sériessédimentaires de la marge passive saxothuringienne et les reliques de la croûte océanique ordovicienne ont été obductées sur le soclesaxothuringien alors que, dans le même temps, la croûte moyenne Teplá-barr<strong>and</strong>ienne était extrudée le long de la suture de Teplá.Plus à l’est, cet événement était associé au développement d’un arc magmatique et à la formation d’un bassin avant-arc sur le socleTeplá-Barr<strong>and</strong>ien. Le domaine arrière-arc – la future zone moldanubienne – subissait un amincissement lithosphérique qui affectaitaussi en partie la croûte continentale de Brunia. La subduction s’est achevée avec la collision de la croûte continentalesaxothuringienne et s’est traduite par un épaississement crustal et la formation d’une racine orogénique dans les zones del’arc magmatique et de l’arrière-arc. L’épaississement était associé à une dépression du Moho et au flux de croûte felsiquesaxothuringienne dans la racine orogénique. L’anisotropie sub-horizontale de la racine a donné suite à un plissement serré, d’échellecrustale, au front du butoir continental de Brunia. Pendant le Viséen, le continent Brunia a indenté la racine crustale épaissie enprovoquant un important raccourcissement de la racine, l’extrusion verticale de la croûte inférieure de l’orogène passant à un fluxvisqueux en chenal horizontal, de matériaux de la croûte inférieure à des niveaux médio- et supra-crustaux. Les roches de la croûteinférieure orogénique ont été extrudées (1) dans d’étroits chenaux parallèles à la surface de l’ancienne suture de Teplá, (2) au centrede la zone de racine sous la forme d’un large antiforme et (3) en nappe plissée chaude au-dessus du promontoire de Brunia, tout enproduisant un métamorphisme barrovien, ainsi que l’imbrication des niveaux supérieurs. Les roches extrudées les plus pr<strong>of</strong>ondes dela racine ont atteint la surface, puis ont été reprises, peu de temps après, sous la forme de galets dans la sédimentation de type Culmdans l’épais bassin d’avant-pays qui recouvre la partie stable du continent Brunia. Enfin, pendant le Westphalien, le prismed’accrétion du Culm, ainsi que le socle et les nappes de chenaux crustaux ont été repris dans une suite de nappes imbriquées. Pourciter cet article : K. Schulmann et al., C. R. Geoscience 341 (2009).# 2009 Publié par Elsevier Masson SAS pour l’Académie des sciences.Keywords: Bohemian Massif; Saxothuringian oceanic subduction; Building <strong>of</strong> Variscan orogenic root system; Channel flowMots clés : Massif de Bohême ; Subduction océanique saxothuringienne ; Formation de la racine orogénique varisque ; Flux en chenal1. IntroductionStarting with [16] <strong>and</strong> followed by [33] <strong>and</strong> [93,94],the eastern Variscan belt is interpreted as the result <strong>of</strong>Devonian to Carboniferous continent–continent collision,resembling the (sub-)recent Himalayan-Tibetantype collisional system. However, there is a wealth <strong>of</strong>data (see for example [93,94]) suggesting that thisorogenic belt could have resulted from an Andean typeconvergence, i.e. as a typical upper plate orogen locatedabove a long-lasting Devonian–Carboniferous subductionsystem.The aim <strong>of</strong> this article is to show that all the currentcriteria defining an Andean type <strong>of</strong> convergent marginare present <strong>and</strong> surprisingly well preserved in theVariscan Bohemian Massif. These criteria are inparticular [66,71,122,157]: the development <strong>of</strong> Franciscan type blueschist-faciesmetamorphism in the lower plate; arc type magmatism marked by calc-alkaline topotassium-rich (shoshonitic) series in the distance <strong>of</strong>150–200 km from the trench; back-arc basin developed on continental upper platecrust <strong>and</strong> replaced by thick continental root; deep granulite-facies metamorphism associated withsupposed underplating <strong>of</strong> the crust by mafic magmasat the bottom <strong>of</strong> the root; continental lithosphere thrust underneath the thickenedroot system.Based on these criteria, the architecture <strong>of</strong> theeastern Variscan belt is interpreted as the result <strong>of</strong> alarge-scale <strong>and</strong> long-lasting subduction process associatedwith crustal tectonics, metamorphism, magmatic<strong>and</strong> sedimentary additions that developed over thewidth <strong>of</strong> at least 500 km, in present-day coordinates,<strong>and</strong> time scale <strong>of</strong> 80 Ma.2. Present-day architecture <strong>of</strong> the BohemianMassif <strong>and</strong> location <strong>of</strong> Palaeozoic suturesThe Bohemian Massif occurs at the eastern extremity<strong>of</strong> the European Variscan belt, representing one <strong>of</strong> itslargest exposures (Fig. 1). From west to east, the EasternVariscan belt forming the Bohemian Massif can besubdivided into four major units: the Saxothuringian Neoproterozoic basement with itsPalaeozoic cover corresponding to the continentalcrust <strong>of</strong> the Armorican plate [87,105,147]; the Teplá-Barr<strong>and</strong>ian Unit consisting <strong>of</strong> Neoproterozoicbasement <strong>and</strong> its Early Palaeozoic coverinterpreted as an independent crustal block (theBohemia Terrane <strong>of</strong> South Armorica sensu [34]); the Moldanubian high- to medium-grade metamorphicdomain intruded by numerous Carboniferous160


Author's personal copy268K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286Fig. 1. Simplified geological map <strong>of</strong> the Bohemian Massif (modified after [34]). CBPC: Central Bohemian Plutonic Complex; CMP: CentralMoldanubian Pluton. The lower left insert shows position <strong>of</strong> the Bohemian Massif in the frame <strong>of</strong> the European Variscides (modified [22]).RH: Rhenohercynian zone; ST: Saxothuringian Zone; M: Moldanubian Zone; B: Brunia Continent; L: Lugian domain.Fig. 1. Carte géologique simplifiée du Massif de Bohême (modifiée d’après [34]). CBPC : Complexe plutonique de Bohême centrale ; CMP : Plutoncentral moldanubien. L’encart du bas à gauche montre la position du Massif de Bohême dans le cadre des Variscides européennes (modifié d’après[22]). RH : Zone rhénohercynienne ; ST : Zone saxothuringienne ; M : Zone moldanubienne ; B : Brunia ; L : Zone lugienne.granitic plutons, altogether forming the high-gradecore <strong>of</strong> the orogeny; the eastern Brunia Neo-Proterozoic basement withEarly to Late Palaeozoic cover [35,140].The palaeontological record <strong>of</strong> Lower Palaeozoic(Cambrian <strong>and</strong> Ordovician) sediments <strong>of</strong> the Saxothuringian<strong>and</strong> Teplá-Barr<strong>and</strong>ian domains shows affinitiesto the Gondwana faunas implying that these blockswere derived from the northern Gondwana margin[20,27,28,68]. In addition, there is a range <strong>of</strong> isotopic<strong>and</strong> U–Pb zircon data suggesting a Gondwananprovenance <strong>of</strong> all units composing the BohemianMassif [96,106].2.1. Saxothuringian domainThis domain is represented by Neoproterozoic parautochthonousrocks (580–550 Ma) formed by migmatites<strong>and</strong> paragneisses intruded by Cambro-Ordoviciancalc-alkaline porphyritic granitoids converted to augenorthogneiss during the Variscan orogeny [38,39,76,89].These rocks are unconformably covered by Cambrian<strong>and</strong> Ordovician rift sequences (Fig. 1) overlain by LateOrdovician to Famennian pelagic sediments <strong>and</strong>Famennian to Visean flysh <strong>of</strong> the Thuringian facies[85,125,144]. The par-autochthon is overthrust byallochthonous units containing deep-water equivalents<strong>of</strong> the Ordovician to Devonian rocks <strong>of</strong> the paraautochthon,<strong>and</strong> by proximal flysh sediments (theBavarian facies).The allochthonous units occur in Münchberg,Wildenfels <strong>and</strong> Frankenberg klippens <strong>and</strong> exhibit pile<strong>of</strong> thrust sheets marked by decreasing pressure <strong>and</strong>metamorphic age from the top to the bottom [34]. In thehangingwall occur thrust sheets with Mid-Ocean RidgeBasalt (MORB)-type metabasites <strong>of</strong> Ordovician protolithage eclogitized (Fig. 2; P = 25 kbar, T = 650 8C)161


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 269[34] during the Devonian 395 Ma [102]. Structurallydeeper occur sheets marked by medium pressure (MP)assemblages <strong>and</strong> Late Devonian (365 Ma) zircon <strong>and</strong>hornblende cooling ages. This rock pile represents bothdistal <strong>and</strong> proximal Late Ordovician to Devonianpassive margin rocks tectonically inverted during theDevonian convergence [44,134]. In the Sudetic part(Figs. 1 <strong>and</strong> 2) <strong>of</strong> the Bohemian Massif, the Ordovicianrift sequences are well developed <strong>and</strong> marked by thepresence <strong>of</strong> deep marine sediments <strong>and</strong> MORB-typevolcanics followed by Silurian <strong>and</strong> Devonian sedimentarysequences [95]. The Ordovician oceanic rocks areenhanced by a blueschist-facies metamorphic overprint<strong>of</strong> Late Devonian age [33,34].However, a later Carboniferous underthrusting <strong>of</strong>the Saxothuringian/Armorican continental rocksunderneath the easterly Teplá-Barr<strong>and</strong>ian block wasidentified suggesting a continuous convergence in thisarea, which was responsible for eclogitization <strong>of</strong> boththe oceanic <strong>and</strong> continental crust resulting fromcontinental underthrusting [90] at 340 Ma [71].The continental underthrusting reached even thepressure conditions <strong>of</strong> the diamond stability [150].This event is responsible for the global reworking <strong>of</strong> theSaxothuringian terrane at HP conditions, imbrication<strong>of</strong> subducted continental crust <strong>and</strong> exhumation <strong>of</strong> deeprocks in form <strong>of</strong> crustal scale nappes [92]. Thus,thestructure <strong>of</strong> the Saxothurinigian domain is defined bypar-autochthonous domain (Fig. 2; P =13–15 kbar,T =580–630 8C) [69,70,72] <strong>and</strong> eclogite-bearing‘‘lower’’ crustal nappe (P =20–26 kbar, T =630–700 8C) [70]. The highest tectonic unit, ‘‘the uppernappe’’ reached the high pressure (HP) granulite faciespeak conditions (P =15–20 kbar, T =8008C [69,116])at 340 Ma [79,154]. The exhumation <strong>of</strong> granulite faciesunit occurred at 340 Ma as shown by 40 Ar/ 39 Ar coolingages [80,83].2.2. The Saxothuringian–Teplá-Barr<strong>and</strong>ianboundaryThe boundary between the two crustal domains ischaracterized by the presence <strong>of</strong> units with highproportion <strong>of</strong> ultramafic <strong>and</strong> mafic rocks (Fig. 1;Mariánské Lázně Complex <strong>and</strong> Erbendorf–VohenstraussZone). The former complex is marked by a presence <strong>of</strong>serpentinites at the bottom overlain by a thick sequence<strong>of</strong> amphibolites, eclogites <strong>and</strong> metagabbros. The U–Pbzircon protolith ages discriminate the mafic rocks intotwo main groups with Cambrian (540 Ma [145]) <strong>and</strong>Ordovician (496 Ma [5]) ages. On the other h<strong>and</strong>, themetamorphic <strong>and</strong> cooling ages (Sm–Nd garnet–pyroxene<strong>and</strong> Ar–Ar hornblende) are Devonian, ranging between410 <strong>and</strong> 370 Ma [4,15]). Similarly, the newly determinedU–Pb ages for metamorphic zircon in mafic rocks, as wellas monazite from orthogneiss, <strong>and</strong> titanite in leucosomeall cluster around 380 Ma <strong>and</strong> possibly date theexhumation <strong>of</strong> the whole Mariánské Lázně Complex[163]. The Devonian metamorphic evolution started witheclogite-facies metamorphism (Fig. 2; P =16–18 kbar,T = 640–715 8C) [145] <strong>and</strong> continued at c. 380 Ma bygranulite-facies re-equilibration [99]. The Erbendorf–Vohenstrauss Zone further west shows similar lithological<strong>and</strong> petrological zonation as the Mariánské LázněComplex <strong>and</strong> it is thus commonly interpreted as a part <strong>of</strong>the same tectonic unit [101,145].2.3. Teplá-Barr<strong>and</strong>ian domainStratigraphically, the Teplá-Barr<strong>and</strong>ian Unit (Fig. 1)consists <strong>of</strong> Neoproterozoic basement with the lower arcrelatedvolcano-sedimentary sequence (the Kralupy–Zbraslav Group), followed by siliceous black shales <strong>and</strong>a flyshoid sequence (shales, greywackes <strong>and</strong> conglomerates,[35]). The Neoproterozoic basement is unconformablyoverlain by a thick sequence (1500–2000 m)<strong>of</strong> Lower Cambrian conglomerates, greywackes, <strong>and</strong>s<strong>and</strong>stones [19,82] followed by shales <strong>and</strong> a rift-relatedvolcanic sequence in the Upper Cambrian. Thedevelopment <strong>of</strong> the Lower Palaeozoic Prague Basin[17] is marked by Early Ordovician (Tremadocian)transgression followed by mid-Ordovician riftingassociated with volcanic activity, <strong>and</strong> with sedimentation<strong>of</strong> Silurian graptolite shales. The sedimentationcontinued mainly with carbonates, namely the UpperSilurian to Devonian calcareous flyshoid sequence. TheEarly Silurian was associated with important volcanicactivity, accompanied by basaltic <strong>and</strong> ultramaficintrusions [51]. The sedimentation terminated in mid-Devonian with distal turbidites <strong>of</strong> the Srbsko Formation(Givetian) containing, among others, detrital zircons <strong>of</strong>Devonian (390 Ma) age [108].The whole sequence is folded by steep foldspresumably <strong>of</strong> Late Devonian age as shown by Culmfacies sediments unconformably deposited on foldedOrdovician strata [135]. The deformation affected alsothe underlying Neoproterozoic basement, havingintensity <strong>and</strong> age increasing progressively to the west.In the same direction rises also the metamorphic gradereaching amphibolite-facies conditions close to theTeplá-Barr<strong>and</strong>ian/Saxothuringian boundary [10]. In thisarea is developed a typical Barrovian metamorphiczonation ranging from biotite–garnet zone (Fig. 2a;P =3–4 kbar, T = 450–520 8C) [146,156,160,161] in162


Author's personal copy270K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286Fig. 2. Pressure–temperature (P–T) diagram showing the location <strong>of</strong> the principal metamorphic facies in the P–T space (after Brown, [6]). BS:blueschist facies; AEE: amphibole-epidote eclogite facies; ALE: amphibole-lawsonite eclogite facies; LE: lawsonite eclogite facies; AE: amphiboleeclogite facies; GS: greenschist facies; A: amphibolite facies; E-HPG: medium-temperature eclogite facies – high-pressure granulite metamorphism;G: granulite facies; UHTM: the ultra-high-temperature metamorphic part <strong>of</strong> the granulite facies. a: Devonian HP–LT metamorphism (390–380 Ma) from the Münchberg <strong>and</strong> Mariánské Láznĕ units contrasted to Devonian (380 Ma) MP metamorphism <strong>of</strong> the western part <strong>of</strong> the Teplá-Barr<strong>and</strong>ian; b: Carboniferous (340 Ma) HP–LT metamorphism from the Saxothuringian par-autochthon <strong>and</strong> lower nappe contrasted to the HP–HTmetamorphism <strong>of</strong> the granulite bearing upper nappe (modified after [71]); c: Carboniferous (340 Ma) HP–HT metamorphism <strong>of</strong> the orogenic lowercrustal belt from the Moldanubian domain, HT–MP <strong>and</strong> LP metamorphism related to the exhumation <strong>of</strong> the orogenic lower crust contrasted to theMP–MT peak metamorphism <strong>of</strong> the orogenic middle crust followed by heating during exhumation (modified after [123]; d: Carboniferous HP–MTmetamorphism <strong>of</strong> the Moravian eclogites contrasted to the MP–MT Barrovian metamorphism <strong>of</strong> the Moravian Zone (modified after [77]).Fig. 2. Diagramme pression–temperature (P–T) montrant la localisation des principaux faciès métamorphiques dans l’espace P–T (d’après [6]). BS :faciès schiste bleu ; AEE : faciès éclogitique à amphibole-épitote ; ALE : faciès éclogitique à amphibole-lawsonite ; LE : faciès éclogitique àlawsonite ; E-HPG : métamorphisme à éclogite de moyenne température et à granulite de haute pression ; G : faciès granulite ; UHTM : partiemétamorphique d’ultrahaute température du faciès granulite. a : métamorphisme dévonien HP–LT (390–380 Ma) des unités de Münchberg etMariánské Láznĕ contrastant avec le métamorphisme MP (380 Ma) de la partie occidentale de l’unité de Teplá-Barr<strong>and</strong>ien ; b : métamorphismecarbonifère HP–LT (340 Ma) du parautochtone Saxothuringien et de la nappe inférieure, contrastant avec le métamorphisme HP–HT de la granulite163


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 271the east up to kyanite zone (Figs. 1 <strong>and</strong> 2a; P =5–8 kbar,T = 530–620 8C) [156] in the west marked by a‘‘normal’’ gradient (increase <strong>of</strong> pressure <strong>and</strong> temperatureto the <strong>structural</strong> footwall). The 40 Ar/ 39 Ar muscovite<strong>and</strong> hornblende dating yielded exclusively MiddleDevonian cooling ages [155].2.4. The Teplá-Barr<strong>and</strong>ian–Moldanubian domainsboundary – the Central Bohemian PlutonicComplexThe igneous activity in the area <strong>of</strong> the CentralBohemian Plutonic Complex (Fig. 1) started withintrusions <strong>of</strong> calc-alkaline Devonian (protolith370 Ma) tonalites to granodiorites, latter transformedinto highly sheared orthogneisses [74]. The firstunmetamorphosed plutonic rocks were Late Devonian(354 Ma) calc-alkaline tonalites, granodiorites,trondhjemites, quartz diorites <strong>and</strong> gabbros <strong>of</strong> theSázava suite [62,64,158]. Their Sr–Nd isotopic ratios<strong>and</strong> trace-element signature indicate that the source <strong>of</strong>the basic magmas was a slightly depleted mantle abovea subduction zone. In addition, a significant role formixing with acidic magmas is to be assumed [60,64].Further south/southeast occur voluminous Early Carboniferous(349–346 Ma [18,55,56]) high-K calcalkalineplutonic bodies <strong>of</strong> the Blatná suite (mainlygranodiorites with minor quartz monzonite <strong>and</strong> monzogabbrobodies). The intermediate rock types resultedfrom mixing <strong>of</strong> slightly enriched mantle-derived <strong>and</strong>crustal magmas [61]. Finally, further east occur syndeformationalbodies or post-tectonic elliptical intrusions<strong>of</strong> (ultra-)potassic rocks <strong>of</strong> mid-Carboniferous(343–337 Ma) ages [53,55,63,157]. Both the Sázava<strong>and</strong> Blatná suites contain numerous xenoliths, screens<strong>and</strong> ro<strong>of</strong> pendants <strong>of</strong> the Barr<strong>and</strong>ian-like Palaeozoic <strong>and</strong>Neo-Proterozoic sequences, indicating a major role forstopping as an emplacement mechanism [159].All these features indicate that the Central BohemianPlutonic Complex corresponds to a relatively shallowsection (


Author's personal copy272K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286conditions to the magmatic emplacement <strong>of</strong> granuliteprotoliths at moderate depth <strong>of</strong> c. 7–13 kbar, followed bycooling <strong>and</strong> prograde metamorphic overprint <strong>and</strong>decompression occuring at temperatures lower than850 8C [128,129]. The granulite-facies overprint isprobably Visean in age as shown by a number <strong>of</strong> zirconages [1,84,122,131,153] but recent studies indicatepossible Devonian age <strong>of</strong> 370 Ma [2].Based on existing pressure-temperature (P–T)estimates two NW–SE trending belts <strong>of</strong> HP rocks(granulites, eclogites <strong>and</strong> peridotites) are distinguished,one located close to the Barr<strong>and</strong>ian–Moldanubianboundary (the western belt <strong>of</strong> Finger et al. [29]) <strong>and</strong>the other rimming the eastern margin <strong>of</strong> the BohemianMassif. These belts alternate with MP units, representedby the Varied <strong>and</strong> Monotonous groups, which also formNW–SE trending wide belts.The deformation history in the Moldanubian Zonereveals early vertical NNE–SSW trending fabrics,associated with crystallization <strong>of</strong> HP mineral assemblages[31,113,131,142,151]. These are reworked byflat deformation fabrics that are associated with MP tolow-pressure (LP) <strong>and</strong> high-temperature (HT) mineralassemblages [50,133,142,143,151]. The flat fabricsshow intense NE–SW trending mineral lineation that iscommonly associated with generalized ductile flowtowards northeast <strong>and</strong> this kind <strong>of</strong> deformation is typical<strong>of</strong> the whole eastern margin <strong>of</strong> the Bohemian Massif.The early steep fabrics are dated at 350 to 340 Ma [122],while the ages <strong>of</strong> the flat ones cluster generally around335 Ma [123]. In the southwestern part <strong>of</strong> theMoldanubian domain, younger set <strong>of</strong> steep NW–SEmetamorphic fabrics reworks the flat foliation, havingbeen associated with LP metamorphic conditions ataround 325–315 Ma [29,138].The HP granulites are spatially, <strong>structural</strong>ly <strong>and</strong>temporally associated with (ultra-)potassic melasyenitesto melagranites, which can be divided into twogroups differing in modal mineralogy <strong>and</strong> textures: coarsely porphyritic K-feldspar melasyenites tomelagranites <strong>of</strong> the so-called durbachite group/serieswith a ‘‘wet’’ mineral assemblage Mg-biotite plusactinolitic hornblende (e.g., Čertovo břemeno <strong>and</strong>Třebíč intrusions); even-grained melasyenites–melagranites (Tábor <strong>and</strong>Jihlava intrusions), containing a variously retrogressed,originally almost ‘‘dry’’, assemblage <strong>of</strong>two pyroxenes plus Mg-biotite [29,53,59,149].For the most basic ultra-potassic rocks, the highcontents <strong>of</strong> Cr <strong>and</strong> Ni with high melting point toderivation from an olivine-rich source (i.e., Earth’smantle). On the other h<strong>and</strong>, elevated concentrations <strong>of</strong> U,Th, LREE <strong>and</strong> LILE, pronounced depletion in HFSE aswell as high K/Na <strong>and</strong> Rb/Sr ratios seem to contradict themantle origin. This dual geochemical character <strong>and</strong>crustal-like Sr–Nd isotopic compositions require melting<strong>of</strong> anomalous lithospheric mantle sources, metasomatised<strong>and</strong> contaminated by mature crustal material ([59]<strong>and</strong> references therein), <strong>and</strong> interaction <strong>of</strong> these maficmelts with crustally-derived leucogranitic magmas.The Moldanubian metamorphic units were penetratedby numerous <strong>and</strong> voluminous anatectic plutons looselygrouped into the Moldanubian (or South Bohemian)Plutonic Complex [26,45,46,54]. These are mostlyfelsic–intermediate, two-mica granitic to granodioriticintrusions with either S- or transitional I/S type character[26,88,148]. High-resolution conventional U–Pb zircon<strong>and</strong> monazite dating showed that the bulk <strong>of</strong> theMoldanubian Plutonic Complex (c. 80%) was emplacedat 331–323 Ma. The fine-grained granodiorites associatedwith minor diorites followed in a second, lessimportant event at 319–315 Ma [47]. The post-tectonicintrusions <strong>of</strong> the Moldanubian Plutonic Complexpostdated shortly the thermal peak <strong>of</strong> the regionalmetamorphism, but were significantly younger than theemplacement <strong>of</strong> both the Central Bohemian PlutonicComplex <strong>and</strong> (ultra-)potassic plutonic rocks scatteredthroughout the Moldanubian domain.Granitic rocks in the southwestern sector <strong>of</strong> theBohemian Massif (Bavarian Forest) have largelyindependent position. Here, large-scale crustal anatexiswas connected with a significant reheating (LP–HTregional metamorphism) <strong>and</strong> a tectonic remobilisation <strong>of</strong>the crust (Bavarian Phase sensu [29]). The intrusionsnortheast <strong>of</strong> the Bavarian Pfahl Zone are dated between328 <strong>and</strong> 321 Ma, whereas ages between 324 <strong>and</strong> 321 Maare obtained southwest <strong>of</strong> this zone. It apparentlyrepresents an important terrane boundary <strong>and</strong> the graniticintrusions sampled distinct basement units [124].2.6. The Moldanubian–Brunia continentaltransition zoneThis boundary was defined by Suess [136,137] as azone <strong>of</strong> severe deformation <strong>and</strong> metamorphism <strong>of</strong>continental rocks (the so-called Moravo-Silesian Zone,Fig. 1). In his seminal work, Suess [136] defined the deepthrusting <strong>of</strong> the Moldanubian Zone over more external<strong>and</strong> shallower Moravo-Silesian Zone, the latter emergingthrough Moldanubian nappe in a form <strong>of</strong> several tectonicwindows. The contact between these units is marked by aparticular unit, the Moravian ‘‘micaschist zone’’, which165


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 273is composed <strong>of</strong> kyanite-bearing micaschists that wereinterpreted by Suess [136] as a result <strong>of</strong> deep crustalretrogression (‘‘diaftorism’’) <strong>of</strong> the Moldanubiangneisses. Modern studies by Konopásek et al. [73] <strong>and</strong>Štípská et al. [132] have shown that this zonecontains boudins <strong>of</strong> eclogites (Fig. 2; P = 16 kbar,T = 650 8C), HP granulites <strong>and</strong> peridotites embeddedin the metapelites. This zone, which contains elements <strong>of</strong>both Moravian <strong>and</strong> Moldanubian parentage, is thusregarded as a first order tectonic boundary.The underlying Moravo-Silesian Zone (Fig. 1) ischaracterized by two nappes composed <strong>of</strong> orthogneissesat the bottom <strong>and</strong> a metapelite sequence at the top.These nappes are thrust over the Neoproterozoicbasement which is <strong>of</strong>ten imbricated with its Pragianto Famennian cover [9,43]. A number <strong>of</strong> isotopic studies(e.g. [28]) show that the orthogneisses <strong>of</strong> the MoravianZone are derived from the underlying Brunia continent[20]. This 50 km wide <strong>and</strong> 300 km long zone <strong>of</strong> intensedeformation is marked by an inverted metamorphicsequence ranging from chlorite to kyanite-sillimanitezones <strong>and</strong> P–T conditions ranging from 5–10 kbar <strong>and</strong>550–650 8C (Fig. 2) [77,127]. The prograde metamorphismis interpreted as a result <strong>of</strong> continentalunderthrusting associated with intense top-to-the-NNEoriented shearing, development <strong>of</strong> sheath folds <strong>and</strong>gneissification <strong>of</strong> Brunia-derived granite protoliths[118,121]. The subsequent deformation is connectedwith recumbent folding <strong>and</strong> imbrication <strong>of</strong> Neoproterozoicgneisses with Devonian cover [119,120]. This laterphase is interpreted as late imbrications <strong>of</strong> metamorphosedBrunia crust in a kinematic continuum with earlyunderthrusting event. The cooling after the Barrovianmetamorphism is constrained at 340–325 Ma byhornblende Ar–Ar data [14,40] <strong>and</strong> monazite inclusionsin garnets yielding 340–330 Ma CHIME ages [78].2.7. The Brunia continentThe Brunia continental forel<strong>and</strong> (Fig. 1) originallycalled the Bruno-Vistulicum by Dudek [20] consists <strong>of</strong>Neoproterozoic granitoids, migmatites <strong>and</strong> schists.They reveal the existence <strong>of</strong> a 680-Ma-old crust,intruded by 550-Ma-old granites [28,40]. This basementis unconformably overlain by Cambrian strata [7],Ordovician <strong>and</strong> shallow marine Lower Devonianquartzites <strong>and</strong> conglomerates followed by Givetiancarbonate platform sedimentation [3,68]. Since theEarly Carboniferous (350 Ma), forel<strong>and</strong> sedimentaryenvironment developed being accompanied by extensionalfaulting indicating the flexural subsidence <strong>of</strong> theBrunia margin [49,96]. From 345 until 300 Ma, a7.5 km thick Variscan flysch (Culm facies) wasdeposited onto the Brunia forel<strong>and</strong>. This sedimentationwas coeval with the onset <strong>of</strong> massive exhumation in theneighbouring Moldanubian domain <strong>and</strong> the continuousloading <strong>of</strong> the Brunia continent [49]. Low-grade sourcerocks associated with clastic muscovites dated at 345–330 Ma [117] gradually pass to a high-grade metamorphicsource material marked by pyrope-rich mineralfraction <strong>and</strong> granulite pebbles dated at 340–330 Ma[13,49,81,152]. Since 330 Ma, began also lateralshortening <strong>of</strong> the flysch basin marked by a deformationfront progressively migrating from eastward in conjunctionwith decreasing intensity <strong>of</strong> deformation[30,57]. Deformation <strong>of</strong> this flysch terminated atc. 300 Ma as indicated by Ar–Ar cooling ages <strong>of</strong> themetamorphosed Culm facies in the west [91] <strong>and</strong>deformation <strong>of</strong> the Variscan molasse further east [11].3. Geodynamic evolution <strong>of</strong> the BohemianMassifThe spatial <strong>and</strong> temporal distribution <strong>of</strong> geologicalunits, magmatic fronts <strong>and</strong> metamorphic zones can beinterpreted in an evolutionary scheme very similar to thatcurrently reported from the Andean type orogeny. Thefollowing account provides a succession <strong>of</strong> tectonicevents that can be interpreted in terms <strong>of</strong> south-eastward(in the present-day coordinates) oceanic subductionunderneath an active continental margin, obduction <strong>of</strong>the Saxothuringian oceanic domain, formation <strong>of</strong> afore-arc region, growth <strong>of</strong> magmatic arc <strong>and</strong> development<strong>of</strong> a large-scale back-arc system on the continentallithosphere. The early oceanic subduction event couldhave been followed by a continental underthrusting <strong>of</strong> theArmorican plate leading to increased friction between theupper <strong>and</strong> lower plates, gradual flattening <strong>of</strong> thesubduction zone marked by eastward migration <strong>of</strong> arc<strong>and</strong> subsequent crustal thickening. The latter event couldhave been responsible for the development <strong>of</strong> a thickcontinental root due to thickening <strong>of</strong> the upper platerepresented by the Teplá-Barr<strong>and</strong>ian <strong>and</strong> Moldanubiancrustal material. The final evolution is marked by thecontinental indentation <strong>of</strong> eastern Brunia continent into aweak orogenic root, exhumation <strong>of</strong> the Moldanubianorogenic lower crust, collapse <strong>of</strong> the Teplá-Barr<strong>and</strong>ian lid<strong>and</strong> Moldanubian thrusting over the Brunia platform.3.1. Early Devonian oceanic subductionunderneath the active continental marginThe contact between the Saxothuringian-Armoricanplate <strong>and</strong> the overriding Teplá-Barr<strong>and</strong>ian continent is166


Author's personal copy274K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286167


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 275marked by relics <strong>of</strong> Ordovician MORB eclogites <strong>and</strong>metabasites locally associated with Ordovician sedimentsmetamorphosed under blueschist–eclogite faciesconditions indicating a Mid-Devonian oceanic subduction[34] <strong>and</strong> an existence <strong>of</strong> Lower PalaeozoicSaxothuringian oceanic domain (Fig. 3a). The exactage <strong>and</strong> size <strong>of</strong> the Saxothuringian ocean is not known,because the differences <strong>of</strong> Ordovician <strong>and</strong> LowerDevonian faunas between Saxothuringian <strong>and</strong> Barr<strong>and</strong>iancontinental domains are not confirmed, precludingan existence <strong>of</strong> large oceanic barrier <strong>of</strong> biogeographicalsignificance [10,52,106]. Therefore, the oceanic domainmust have been narrow <strong>and</strong> potentially short lived inorder not to be recorded by palaeobiogeographic faunalevidence. The only existing argument for larger oceanicseparation are from palaeomagnetic data <strong>of</strong> Tait et al.,[141] suggesting that the Saxothuringian <strong>and</strong> Tepla-Barr<strong>and</strong>ian blocks rotated independently during Silurian<strong>and</strong> Devonian times.The hornblende <strong>and</strong> mica cooling ages <strong>of</strong> theVohenstraus–Erbendorf Zone, Mariánské Lázně complex<strong>and</strong> the Sovie Gory granulites cluster around380 Ma, which indicates early collision betweenthe Teplá-Barr<strong>and</strong>ian <strong>and</strong> Saxothuringian domains(Figs. 3middle <strong>and</strong> 4B). This is also confirmed byFamennian flysh sediments <strong>of</strong> the Saxothuringian basinthat contain detrital zircons dated at 380 Ma [114]which is a direct evidence <strong>of</strong> mutual contact betweenthe Teplá-Barr<strong>and</strong>ian <strong>and</strong> Saxothuringian domainsduring Famennian but most likely already during mid-Devonian. The Barrovian metamorphic zonation developedat the western flank <strong>of</strong> the Teplá-Barr<strong>and</strong>ian domain(Figs. 3middle <strong>and</strong> 4B) is potentially related todeformation <strong>of</strong> the overriding Teplá-Barr<strong>and</strong>ian continentalmargin <strong>and</strong> is consistent with a model <strong>of</strong>extrusion <strong>of</strong> lower part <strong>of</strong> the Teplá-Barr<strong>and</strong>ian crustduring Devonian shortening event [160,161]. The folding<strong>of</strong> the eastern very low grade shales <strong>and</strong> volcanics <strong>of</strong> theNeoproterozoic sequences is interpreted as a supracrustalresponse to the same deformation event affecting highgrade rocks in the west [161]. Thus, the exhumation <strong>of</strong>HP rocks in the Mariánské Lázně Complex <strong>and</strong> thedeformation <strong>of</strong> the Teplá-Barr<strong>and</strong>ian basement arecoupled <strong>and</strong> related to convergent processes at theSaxothuringian plate boundary (Figs. 3middle <strong>and</strong> 4B).The strong argument for eastward subduction <strong>of</strong> theSaxothuringian ocean underneath the eastern Teplá-Barr<strong>and</strong>ian continent is a Devonian–Lower Carboniferousmagmatic arc that is firmly founded on thecontinental crust. This arc, the Central BohemianPlutonic Complex, separated the Teplá-Barr<strong>and</strong>i<strong>and</strong>omain from the future Moldanubian domain <strong>and</strong> inthis model the former unit represented a fore-arc regionduring Devonian (Figs. 3middle <strong>and</strong> 4B). Thus, theposition <strong>of</strong> Devonian HP rocks thrust over theSaxothuringian basement, existence <strong>of</strong> the MariánskéLázně Complex – the oceanic fragment at sutureposition, <strong>and</strong> location <strong>of</strong> calc-alkaline magmatic rocksfurther east confirm a polarity <strong>of</strong> the oceanic subductionunderneath the eastern fore-arc <strong>and</strong> magmatic arcsystem during Late Devonian (Figs. 3middle <strong>and</strong> 4B).The distance between the arc <strong>and</strong> the trench arearepresented by the suture indicates that the dip <strong>of</strong>subduction zone was probably moderate (30–408,Fig. 4B). The temporal evolution <strong>of</strong> magma geochemistryfrom calc-alkaline to more potassic/shoshoniticaffinities (from 370 to 336 Ma) is compatible withflattening <strong>of</strong> the subduction zone <strong>and</strong> increased melting<strong>of</strong> continental material during Early Carboniferous. TheBarr<strong>and</strong>ian (Prague) Basin is interpreted as a part <strong>of</strong> thefore-arc basin marked by 380-Ma-old detrital zirconsfound in Mid-Devonian flyshoid sequences [135]suggesting early erosion <strong>of</strong> the eastern magmatic arc.The most problematic question in the Devoniansubduction model for the Bohemian massif is the role <strong>of</strong>the future Moldanubian domain as well as the position<strong>of</strong> the Brunia continent. The amphibolites derived fromSiluro-Devonian tholeiitic basalts associated withcarbonates, widespread in Lower Austria <strong>and</strong> SouthBohemia, are interpreted as volcanic products <strong>of</strong> a relicFig. 3. Time slices maps showing the chronological <strong>and</strong> tectonic significance <strong>of</strong> the various units <strong>of</strong> the Bohemian Massif. Top: Ordovician timeslice showing the location <strong>of</strong> Ordovician MORB <strong>and</strong> deep marine sediments, <strong>and</strong> that <strong>of</strong> the Proterozoic units containing the Ordovician magmatism;middle: Devonian time slice showing the location <strong>of</strong> Devonian high pressure (HP) rocks, the passive margin sequences, the active marginmetamorphism, the fore-arc region, the magmatic arc, the location <strong>of</strong> Devonian infrastructure <strong>and</strong> sediments; bottom: Carboniferous time sliceshows the location <strong>of</strong> orogenic lower <strong>and</strong> middle crust belts in the Moldanubian domain, the metamorphism <strong>of</strong> the Moravian Zone, the MoravianMicaschist Zone, the HP units in Saxothuringian domain <strong>and</strong> the various types <strong>of</strong> magmatic rocks.Fig. 3. Cartes selon des tranches de temps montrant la signification chronologique et tectonique des diverses unités du Massif de Bohême. Enhaut : tranche de temps ordovicienne montrant la localisation du MORB ordovicien et des sédiments marins pr<strong>of</strong>onds et celle des unitésprotérozoïques incluant le magmatisme ordovicien ; au milieu : tranche de temps dévonienne montrant la localisation des roches de haute pression(HP) dévoniennes, les séquences de marge active, le domaine avant-arc, l’arc magmatique et la localisation de l’infrastructure et des sédimentsdévoniens ; en bas : tranche de temps carbonifère montrant la localisation des chaînes orogéniques crustales inférieure et moyenne du Domainemoldanubien, le métamorphisme de la zone moravienne, de la zone des micaschistes moraviens, des unités HP du domaine saxothuriningien et desdifférents types de roches magmatiques.168


Author's personal copy276K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286Fig. 4. Conceptual model <strong>of</strong> the geodynamic evolution <strong>of</strong> the Bohemian Massif shown as schematic cross sections through the orogen. A. Silurian–Early Devonian subduction setting with the beginning <strong>of</strong> arc <strong>and</strong> back-arc formation on upper plate. B. Onset <strong>of</strong> the Mid-Devonian continentalunderthrusting <strong>of</strong> Saxothuringian lithosphere, deformation <strong>of</strong> active margin, formation <strong>of</strong> magmatic arc <strong>and</strong> persistence <strong>of</strong> back-arc region on theupper plate. C. Crustal thickening processes marked by the influx <strong>of</strong> Saxothuringian crust <strong>and</strong> progressive individualization <strong>of</strong> the Brunia activemargin. D. Brunia continent indentation associated with the channel flow process, imbrications <strong>of</strong> underthrust Brunia <strong>and</strong> the formation <strong>of</strong> theMoravian Zone.Fig. 4. Modèle conceptuel de l’évolution géodynamique du Massif de Bohême représenté par des coupes schématiques au travers de l’orogène.A. Subduction Silurien–Dévonien inférieur se mettant en place avec le début de la formation de l’arc et de l’arrière-arc sur la plaque supérieure.B. Au Dévonien moyen, mise en place du sous-charriage continental et de la lithosphère saxothuringienne, déformation de la marge active, formationde l’arc magmatique et persistance de la zone d’arrière-arc sur la plaque supérieure. C. Processus d’épaississement crustal marqué par la venue decroûte saxothuringienne et individualisation progressive de la marge active de Brunia. D. Indentation de Brunia, associée à un processus de flux dechenal, imbrications du sous-charriage de Brunia et formation de la zone moravienne.169


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 277<strong>of</strong> large-scale back-arc system [24,67]. In addition, thefelsic metavolcanics <strong>and</strong> amphibolite layers in theVaried Group are regarded as the continuity <strong>of</strong> back-arcbimodal volcanism till Givetian [37]. However, thesupposed depositional Devonian age <strong>of</strong> the Variedgroup is questioned thanks to low Sr isotopic ratioswhich indicate shallow marine environment duringLate Proterozoic rather than Palaeozoic [32]. Schulmannet al. [122] interpreted the common occurrence <strong>of</strong>Late Silurian–Early Devonian zircons in the high gradeamphibolites <strong>of</strong> the Moldanubian domain as a result <strong>of</strong>important magmatic reworking <strong>of</strong> the lower part <strong>of</strong> thecontinental crust associated with mafic magmaticadditions during lithospheric thinning <strong>of</strong> this domain(Fig. 3middle). In their model, the Monotonous Grouprepresents relic <strong>of</strong> Proterozoic middle crust that is lessaffected by thermal <strong>and</strong> magmatic reworking while theupper crust recorded sedimentary <strong>and</strong> volcanic evolutionrelated to the Silurian–Devonian extensionalevent. A back-arc environment is further supportedby bimodal volcanic activity in narrow Devonian basinsdeveloped on the north-eastern margin <strong>of</strong> the Bruniacontinent [107] suggesting only minor thinning <strong>of</strong>continental crust at the easternmost termination <strong>of</strong> theback-arc system (Figs. 3middle <strong>and</strong> 4B). In this conceptthe rest <strong>of</strong> the Brunia platform represents a stablecontinental domain not affected by the back-arcspreading.The Devonian Saxothuringian oceanic subuduction<strong>and</strong> onset <strong>of</strong> continental underthrusting <strong>of</strong> the Saxothuringian–Armoricanlower plate underneath theupper plate is thus recorded in all units forming thepresent-day Bohemian Massif (Fig. 5). The followingcontemporaneous processes were identified: eclogitization <strong>of</strong> the Saxothuringian crust <strong>and</strong> itsexhumation along the Teplá suture; sedimentation <strong>of</strong> Mid-Devonian distal turbidites <strong>of</strong>the Srbsko Formation followed by inversion <strong>of</strong> thefore-arc basin in the Teplá-Barr<strong>and</strong>ian domain, origin<strong>of</strong> the magmatic arc in the area <strong>of</strong> the CentralBohemian Plutonic Complex <strong>and</strong> exhumation <strong>of</strong> thewestern metamorphosed margin <strong>of</strong> the Teplá unitsuggesting convergent processes in the upper plate; formation <strong>of</strong> the back-arc system on the continentallithosphere as shown by isotopic data in theMoldanubian domain <strong>and</strong> marginally in the Devonianbasins affecting western margin <strong>of</strong> the Bruniacontinent (here, the formation <strong>of</strong> small oceanic basinis not excluded). All existing data point out to thesubduction process that operated at least 45 millionyears from 400 to 355 Ma (Fig. 5).3.2. The Early Carboniferous crustal thickening <strong>of</strong>the upper plateThis event is recognized in all units except the Teplá-Barr<strong>and</strong>ian supracrustal unit. The western margin <strong>of</strong> theBohemian Massif is characterized by the arrival <strong>of</strong> theSaxothuringian continental crust <strong>and</strong> its subductionunderneath the eastern Teplá-Barr<strong>and</strong>ian–Moldanubi<strong>and</strong>omain (Fig. 4B). The main thrust boundary migratedfurther west, so that the continental crust was thrustunderneath the fossil Devonian suture <strong>and</strong> former forearcregion [71]. At the same time the deformationregime changed in the far field back-arc region (futureMoldanubian domain), which recorded the progressivethickening <strong>of</strong> the whole previously thinned <strong>and</strong>thermally s<strong>of</strong>tened domain as indicated by severalrecent petrological studies (e.g. [77,128]). Recent<strong>structural</strong> studies have shown that the earliest preservedfabrics have been sub-horizontal [31,112,123,131],which may indicate that the lower crustal materialwas originally flowing horizontally from the area <strong>of</strong> thecontinental subduction channel towards the region <strong>of</strong>eastern backstop in a manner proposed by [104] for theclosure <strong>of</strong> the Rhenohercynian Basin <strong>and</strong> formation <strong>of</strong>the Mid-German Crystalline Rise.Indeed, the influx <strong>of</strong> lower crustal material transportedby south-east dipping Saxothuringian continentalsubduction zone underneath the fore arc (the Teplá-Barr<strong>and</strong>ian domain) <strong>and</strong> further below the former backarcdomain is regarded to be at the origin <strong>of</strong> the future‘‘Gföhl Unit’’. This hypothesis is in line with the wholerockgeochemical <strong>and</strong> Sr–Nd isotopic composition aswell as the zircon inheritance patterns in the MoldanubianHP–HT granulites [59,65]. Importantly, thecrustal material involved in the subduction <strong>and</strong> extrudedover the sub-arc <strong>and</strong> sub-back-arc mantle lithospheremay have been turned into voluminous HP granulitesknown from many regions <strong>of</strong> the Bohemian Massif[59,100]. Such a processing <strong>of</strong> the lower platecontinental lithosphere in a subduction zone <strong>and</strong> itsextrusion above mantle wedge was successfullymodelled by Gerya <strong>and</strong> Stockert [48]. Alternatively,the back-arc domain with high thermal budget inheritedfrom Devonian stretching may have been thickened <strong>and</strong>the partially molten lower crust may have beentransported downwards <strong>and</strong> transformed into the HPgranulites as suggested by Štípská <strong>and</strong> Powell [129] orSchulmann et al.[122]. However, this model fails toexplain the occurrence <strong>of</strong> the Gföhl gneiss <strong>and</strong> felsicgranulites in the lower crustal position.The onset <strong>of</strong> thickening <strong>of</strong> the root is not recordedin the Teplá domain (Fig. 4C), which behaved as a170


Author's personal copy278K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286Fig. 5. Summary <strong>of</strong> geochronology data showing the time scales <strong>of</strong> the three major events forming the Bohemain Massif: oceanic subduction <strong>and</strong>continental underthrusting <strong>of</strong> Saxothuringian domain, building <strong>of</strong> thick orogenic root system <strong>and</strong> collapse <strong>of</strong> orogen due to indentation <strong>of</strong> Brunia.Fig. 5. Résumé des données chronologiques montrant les échelles de temps des trois évènements majeurs de la formation du Massif de Bohême :subduction océanique et sous-charriage du domaine saxothuringien, édification de l’épais système d’enracinement orogénique et « collapse » del’orogène, en raison de l’indentation de Brunia.supra-<strong>structural</strong> unit at this time, but it is shown bydeformation <strong>of</strong> the Lower Palaeozoic rocks <strong>of</strong> thePrague basin <strong>and</strong> adjacent Late Proterozoic rocks. Here,the steep fabric is well dated by syntectonic calcalkalineplutons at about 355–345 Ma [64,122,158]. Incontrast, the eastern sector <strong>of</strong> the orogen records onset<strong>of</strong> loading <strong>of</strong> the Brunia platform (Fig. 4C) duringTournaisian manifested by destruction <strong>of</strong> the Givetiancarbonate platform <strong>and</strong> sedimentation <strong>of</strong> coarse basalclastics [49].The timing <strong>of</strong> crustal thickening is relatively poorlyconstrained compared to the subduction <strong>and</strong> laterexhumation processes (Fig. 5). However, the Tournaisian<strong>and</strong> Early Visean massive clastic sedimentation onboth Saxothuringian <strong>and</strong> Brunia plates suggests load <strong>of</strong>both continental lithospheric plates <strong>and</strong> high topographyin between them. This corroborates the peakmetamorphic ages in the granulites <strong>and</strong> the Moldanubianeclogites as well as the Visean age <strong>of</strong> compression<strong>of</strong> the magmatic arc. All geochronological <strong>and</strong> othergeological information point to a crustal thickeningperiod that was very short <strong>and</strong> did not last more than20 million years, from 355 to 335 Ma, with a peakaround 340 Ma (Fig. 5).3.3. Late Visean exhumation <strong>of</strong> orogenic lowercrust <strong>of</strong> the upper plateThe exhumation <strong>of</strong> the Variscan lower crust duringEarly Carboniferous is exemplified by the three NE–SW171


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 279trending belts <strong>of</strong> granulites, eclogites <strong>and</strong> peridotites(Fig. 3c) intimately associated with the (ultra-)potassicmagmatites [29,59]. The first granulite belt is representedby narrow strip <strong>of</strong> felsic granulites that occur at theSaxothuringian–Teplá-Barr<strong>and</strong>ian boundary <strong>and</strong> it isinterpreted as an extrusion <strong>of</strong> the orogenic lower crustalong a large scale crustal shear zone at 340 Ma[71,163]. The coeval 340 Ma age <strong>of</strong> the Saxothuringianeclogites <strong>and</strong> <strong>of</strong> the felsic granulites [80,150] ledKonopásek <strong>and</strong> Schulmann [71] to propose a model <strong>of</strong>simultaneous exhumation <strong>of</strong> nappes derived from theSaxothuringian crust <strong>and</strong> viscous extrusion <strong>of</strong> orogeniclower crust from underneath the Teplá-Barr<strong>and</strong>ian supracrustalunit (Fig. 4D) [36]. The second belt, recognizedeast <strong>of</strong> the Central Bohemian Plutonic Complex, i.e.,‘‘the magmatic arc’’, was exhumed along huge westdipping detachment zone [111,157,158], which was alsoresponsible for collapse <strong>of</strong> the upper part <strong>of</strong> the magmaticarc system <strong>and</strong> the downthrow <strong>of</strong> the whole Barr<strong>and</strong>iansection [115,157]. Such a huge vertical material transfer(Fig. 4D) could have been responsible for verticalexchange <strong>of</strong> the lower crustal <strong>and</strong> upper crustal materialin a range <strong>of</strong> 50 km with final throw <strong>of</strong> 15 km [110,162].The cooling ages from the lower crustal domain show thatthe granulites have crossed the 300 8C isotherm duringCarboniferous (330–310 Ma) [75,138] suggesting thetime at which the lower crustal bulge reached shallowposition in the upper plate.The third lower crustal belt rims the eastern margin<strong>of</strong> the Bohemian Massif, i.e. the boundary with theBrunia continent (Fig. 4D). Here the granulite fabricis also vertical <strong>and</strong> interpreted in terms <strong>of</strong> massivevertical exchanges with orogenic middle crust [123].The zone <strong>of</strong> the lower crustal bulge is interpreted as anenormous zone <strong>of</strong> vertical extrusion surrounded by amiddle crust coevally transported downwards in form<strong>of</strong> a crustal scale synform. The vertical materialtransfer along the Moldanubian lower crustal beltswas a matter <strong>of</strong> research for several Czech authorsduring the last five years [31,112,123,131,142]. Themodel <strong>of</strong> vertical extrusion is based on the concept <strong>of</strong>buckling <strong>of</strong> the lower <strong>and</strong> mid-crustal interfacefollowed by growth <strong>of</strong> crustal scale antiforms. Thisprocess is thought to be triggered by rheological <strong>and</strong>thermal instabilities in the arc region, while to the eastit is forced by rigid backstop, preserved only locally[130].However, the most important feature <strong>of</strong> the easternVariscan front is the development <strong>of</strong> horizontal fabricsin the Moldanubian root zone parallel to the Bruniacontinental margin. The intense deformation <strong>of</strong> theBrunia continent leading to the formation <strong>of</strong> theMoravo-Silesian imbricated nappe system was associatedwith the development <strong>of</strong> tectonically invertedBarrovian metamorphism (Fig. 4D) <strong>and</strong> formation <strong>of</strong>crustal mélange in its upper part (the MoravianMicaschist Zone). These phenomena, as well as mixing<strong>of</strong> HP rocks <strong>and</strong> migmatites in the overlyingMoldanubian nappe have been recently interpreted interms <strong>of</strong> indentation <strong>of</strong> the Brunia continent into the hot<strong>and</strong> thick continental root [123]. This lower crustalindentation <strong>and</strong> flow <strong>of</strong> hot lower crustal rocks insupracrustal levels are consistent with a model <strong>of</strong>continental channel flow driven by the arrival <strong>of</strong> acrustal plunger, a model which is advocated for twodecades for the deformation <strong>of</strong> the Eastern Cordillera inthe Andes (e.g. [87]). Finally, the continuous load <strong>of</strong> theBrunia platform related to deep indentation process ledto the development <strong>and</strong> eastward propagation <strong>of</strong> theforel<strong>and</strong> basin. In our model, as the hot Moldanubianrocks advance over the Brunia platform, an imbricatedfootwall nappe system is generated <strong>and</strong> thrust over theprogressively buried forel<strong>and</strong> basin rocks.The time scale <strong>of</strong> exhumation <strong>of</strong> the orogenic lowercrust is well constrained due to a set <strong>of</strong> well dateddiachronous processes that start with the exhumation <strong>of</strong>the West-Bohemian granulites, growth <strong>of</strong> western <strong>and</strong>eastern orogenic lower crustal antiforms <strong>and</strong> a shallowchannel flow <strong>of</strong> partially molten lower crust associatedwith the inversion <strong>of</strong> the eastern forel<strong>and</strong> basin. All thatis linked with a major thermal event in the mantle asshown by the ages <strong>of</strong> a syn-extrusion high potassicmagmatism, the exhumation <strong>of</strong> large number <strong>of</strong> mantlefragments <strong>and</strong> the overall melting <strong>of</strong> the Moldanubiancrust. The time scale <strong>of</strong> all these processes wassurprisingly short, i.e., about <strong>of</strong> 20 Ma, <strong>and</strong> ranges from335 to 315 Ma (Fig. 5).4. Palaeographic constraints for Andean typeorogeny in the Bohemian MassifFinger <strong>and</strong> Steyrer [24] attempted to explain thetectonic processes at the Moldanubian–Moravianboundary with a plate tectonic model involving thesubduction <strong>of</strong> a Silurian–Devonian oceanic domainwestward beneath the Moldanubian zone. This conceptsupposes large rotation <strong>of</strong> the Brunia platform forming apart <strong>of</strong> an Old Red continent (Avalonia) around the core<strong>of</strong> the Bohemian Massif during Visean times [140]. Themodel <strong>of</strong> westward subduction is also based on thepresence <strong>of</strong> Silurian MORB-type amphibolites [25]within the Gföhl Unit, which are interpreted as relics <strong>of</strong>oceanic crust <strong>of</strong> the Rheic Ocean. In addition, theoccurrences <strong>of</strong> eclogites located along the eastern172


Author's personal copy280K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286Variscan front [73,132] can serve as an additionalargument for this subduction model.However, the palaeomagnetic investigations carriedout on granitoids <strong>of</strong> the Central Bohemian Pluton <strong>and</strong> itsextension to the east (the Nasavrky Plutonic Complex[58]) demonstrate that the Teplá-Barr<strong>and</strong>ian, theMoldanubian as well as Brunia domains had a commongeodynamic behaviour since the Late Visean (335–330 Ma according to Edel et al. [22]). Therefore, theinterpretation <strong>of</strong> palaeomagnetic directions from Devoniansediments <strong>of</strong> the Brunovistulian platform[86,98,139] in terms <strong>of</strong> an oroclinal bending <strong>of</strong> theRhenohercynian domain along the Moldanubian corecannot be valid for several reasons: the consistency <strong>of</strong> these ‘‘Devonian’’ palaeomagneticdirections with Middle-Late Carboniferous directions(Cp directions at 330–325 Ma <strong>and</strong> B directions at320–315 Ma, Fig. 6) in the Moldanubian zonesuggests that the Devonian sediments overlying theBrunia continent were re-magnetized in Late Variscantime. Consequently, no relative rotation between theBrunia continent <strong>and</strong> the Moldanubian core can beconsidered prior to 330 Ma; magnetic overprinting <strong>of</strong> the Devonian sequences onthe Brunia continent is supported by importantdeformation <strong>and</strong> burial <strong>of</strong> Culm <strong>and</strong> Devoniansediments during Late Carboniferous deformation <strong>of</strong>the Brunia margin [30,120,126]. Hence, it is unlikelythat primary magnetizations could have survived sucha severe <strong>structural</strong> <strong>and</strong> thermal reworking <strong>of</strong> theDevonian sediments. We suggest that magnetic overprintingwas the result <strong>of</strong> the Carboniferous tectonothermalprocesses associated with the underthrusting <strong>of</strong>the Brunia platform together with Culm accretionarywedge [11] underneath the hot Moldanubian root zone<strong>and</strong> with later uplift <strong>and</strong> erosion <strong>of</strong> the Moldanubian–Moravian nappe sequence.Fig. 6. a: paleomagnetic north directions obtained in Devonian (meta)sediments from the Teplá-Barr<strong>and</strong>ian <strong>and</strong> Brunia zones (K: [86];N:[98];T:[139]; E:[21]) <strong>and</strong> from Early Carboniferous granitoids from Central Bohemian Pluton <strong>and</strong> Nasavrky Pluton [22]. Cn <strong>and</strong> Co directions representmagnetizations acquired at the end <strong>of</strong> the NW–SE shortening <strong>and</strong> the uplift <strong>of</strong> the Moldanubian root; Cp <strong>and</strong> B directions correspond to overprintsacquired during NNE–SSW compression <strong>and</strong> clockwise rotation <strong>of</strong> the Variscides. Note the similarity <strong>of</strong> ‘‘Devonian’’ directions with Carboniferousdirections; b: published mean virtual geomagnetic poles (VGPs) <strong>and</strong> associated apparent polar w<strong>and</strong>er curve (grey dashed line) from Early Paleozoicrocks <strong>of</strong> the Bohemian Massif <strong>and</strong> mean poles from western Europe Variscides (stars with ages <strong>of</strong> magnetization) <strong>and</strong> associated apparent polarw<strong>and</strong>ering curve (dark grey line) [22].Fig. 6. a : directions du nord paléomagmatique obtenues dans les (méta)sédiments dévoniens des zones Teplá-barr<strong>and</strong>ienne et Brunia (K :[86] ;N :[98] ; T :[139] ; E :[21]) et dans les granitoïdes du Carbonifère inférieur et du Pluton Nasavrky [22]. Les directions Cn et Co représentent lesmagnétisations acquises à la fin du raccourcissement NW–SE et du soulèvement de la racine moldanubienne ; les directions Cp et B correspondent àdes réaimantations acquises durant la compression NNE–SSW et la rotation dans le sens des aiguilles d’une montre des Variscides. À noter, lasimilarité des directions « dévoniennes » et carbonifères ; b : pôles géomagnétiques virtuels (VGP) publiés, associés à la courbe de dérive apparentedes pôles (ligne en tiretés) pour les roches du Paléozoïque inférieur du Massif de Bohême et pôles moyens pour les Variscides d’Europe occidentale(étoiles avec âges de la magnétisation), associés à la courbe de dérive des pôles (ligne continue gris foncé) [22].173


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 281The Siluro-Devonian basin in the area <strong>of</strong> theMoldanubian zone existed as a back-arc basin abovethe Saxothuringian subduction zone [122] but thequestion <strong>of</strong> development <strong>of</strong> oceanic crust in this arearemains a matter <strong>of</strong> discussions. Therefore, in contrastto the generally accepted rotation <strong>of</strong> a Brunia continentindependent <strong>of</strong> a stationary Moldanubian domainassociated with a closure <strong>of</strong> a large (Rheic) oceanicdomain we propose a model <strong>of</strong> common geodynamichistory <strong>of</strong> the Bohemian Massif during Devonian <strong>and</strong>Early Carboniferous (Fig. 5). In our model thepalaeomagnetic, crustal scale geophysical <strong>and</strong> <strong>structural</strong>data are in favour <strong>of</strong> an early counterclockwiserotation <strong>of</strong> composite blocks (Saxothuringian, Barr<strong>and</strong>ian,Moldanubian <strong>and</strong> Brunia altogether) accommodatedby a large scale NW–SE trending dextral wrenchzones (such as the Elbe, Pfahl, Franconian <strong>and</strong> Pays deBray faults) during Early Visean [22]. These movementscould have resulted from northwest drift <strong>of</strong>Gondwana continental masses <strong>and</strong> continued after330 Ma by clockwise rotation <strong>of</strong> the whole Variscan beltaround the Euler pole that was continuously translatedwestward parallel to the Teysseire–Tornquist zone(southern margin <strong>of</strong> the Baltica).AcknowledgementsThe French National Science Foundation project‘‘ANR LFO in orogens’’, internal research funds <strong>of</strong>CNRS UMR 7615 <strong>and</strong> grant MSM0021620855 <strong>of</strong> theMinistry <strong>of</strong> Education <strong>of</strong> the Czech Republic areacknowledged for financial support <strong>and</strong> salary <strong>of</strong> OndrejLexa. Jiří Konopásek appreciates the financial support<strong>of</strong> the Grant Agency <strong>of</strong> the Charles University (projectNo. B-GEO-270/2006), as well as the support by theMinistry <strong>of</strong> Education, Youth <strong>and</strong> Sports <strong>of</strong> the CzechRepublic through the Scientific Centre ‘‘AdvancedRemedial Technologies <strong>and</strong> Processes’’ (identificationcode 1M0554).References[1] M. Aftalion, D. Bowes, S. Vrána, Early Carboniferous U-Pbzircon age for garnetiferous, perpotassic granulites, Blanský lesmassif, Czechoslovakia, Neues Jarhb. Miner. Monat. 4 (1989)145–152.[2] R. Anczkiewicz, J. Szczepański, S. Mazur, C. Storey, Q.Crowley, I. Villa, M. Thirlwall, T. Jeffries, Lu–Hf geochronology<strong>and</strong> trace element distribution in garnet: Implications foruplift <strong>and</strong> exhumation <strong>of</strong> ultra-high pressure granulites in theSudetes, SW Pol<strong>and</strong>, Lithos 95 (2007) 363–380.[3] O. Bábek, Č. Tomek, R. Melichar, J. Kalvoda, J. Otava,Structure <strong>of</strong> unmetamorphosed Variscan tectonic units <strong>of</strong> thesouthern Moravo-Silesian zone, Bohemian Massif: a review,Neues Jahrb. Geol. Palaontol. Abh. 239 (2006) 37–75.[4] B.L. Beard, L.G. Medaris, C.M. Johnson, E. Jelínek, J. Tonika,L.R. Riciputi, Geochronology <strong>and</strong> geochemistry <strong>of</strong> eclogitesfrom the Mariánské Lázně Complex, Czech Republic: implicationsfor Variscan orogenesis, Geol. Rundsch. 84 (1995) 552–567.[5] D.R. Bowes, M. Aftalion, U-Pb zircon isotopic evidence forEarly Ordovician <strong>and</strong> Late Proterozoic units in the MariánskéLázné complex, Central European Hercynides, Neues Jahrb.Miner. Monat. (1991) 315–326.[6] M. Brown, Metamorphic conditions in Orogenic Belts: a record<strong>of</strong> secular change, Int. Geol. Rev. 49 (2007) 193–234.[7] Z. Bula, M. Jachowicz, J. Zaba, Principal characteristics <strong>of</strong> theUpper Silesian Block <strong>and</strong> Malopolska Block border zone(southern Pol<strong>and</strong>), Geol. Mag. 134 (1997) 669–677.[8] D. Carswel, P. O’Brien, Thermobarometry <strong>and</strong> geotectonicsignificance in the Moldanubian Zone <strong>of</strong> the Bohemian Massifin Lower Austria, J. Petrol. 34 (1993) 427–459.[9] I. Chlupáč, Fossil communities in the metamorphic LowerDevonian <strong>of</strong> the Hrubý Jeseník Mts., Czechoslovakia, NeuesJahrb. Geol. Palaontol. Abh. 177 (1989) 367–392.[10] I. Chlupáč, Facies <strong>and</strong> biogeographic relationships in Devonian<strong>of</strong> the Bohemian Massif, Courier Forschungsinstitut Senckenberg169 (1994) 299–317.[11] P. Čížek, Č. Tomek, Large-scale thin-skinned tectonics in theeastern boundary <strong>of</strong> the Bohemian Massif, Tectonics 10 (1991)273–286.[12] R. Cooke, P. O’Brien, Resolving the relationship between highP–T rocks <strong>and</strong> gneisses in collisional terranes: an example fromthe Gföhl gneiss–granulite association in the MoldanubianZone, Austria, Lithos 58 (2001) 33–54.[13] R. Čopjaková, P. Sulovský, B. Paterson, Major <strong>and</strong> traceelements in pyrope–alm<strong>and</strong>ine garnets as sediment provenanceindicators <strong>of</strong> the Lower Carboniferous Culm sediments, DrahanyUpl<strong>and</strong>s, Bohemian Massif, Lithos 82 (2005) 51–70.[14] R. Dallmeyer, F. Neubauer, V. Höck, Chronology <strong>of</strong> LatePaleozoic tectonothermal activity in the southeastern BohemianMassif, Austria (Moldanubian <strong>and</strong> Moravo-Silesianzones): 40 Ar/ 39 Ar mineral age controls, Tectonophysics 210(1992) 135–153.[15] R. Dallmeyer, M. Urban, Variscan vs Cadomian tectonothermalactivity in northwestern sectors <strong>of</strong> the Teplá-Barr<strong>and</strong>ian zone,Czech Republic: constraints from 40 Ar/ 39 Ar ages, Geol.Rundsch. 87 (1998) 94–106.[16] J.F. Dewey, K.C.A. Burke, Tibetan, Variscan <strong>and</strong> Precambrianbasement reactivation: products <strong>of</strong> continental collision, J.Geol. 81 (1973) 683–692.[17] W. Dörr, G. Zulauf, J. Fiala, W. Franke, Z. Vejnar, Neoproterozoicto Early Cambrian history <strong>of</strong> an active plate margin inthe Teplá–Barr<strong>and</strong>ian unit—a correlation <strong>of</strong> U–Pb isotopicdilution-TIMSages (Bohemia, Czech Republic), Tectonophysics352 (2002) 65–85.[18] W. Dörr, J. Fiala, W. Franke, U. Haack, S. Philippe, J. Schastok,D. Scheuvens, Z. Vejnar, G. Zulauf, Cambrian vs. Variscantectonothermal evolution within the Teplá-Barr<strong>and</strong>ian: evidencefrom U-Pb zircon ages <strong>of</strong> syn-tectonic plutons (BohemianMassif, Czech Republic), Acta Univ. Carol. Geol. 42(1998) 229–230.[19] K. Drost, U. Linnemann, N. McNaughton, O. Fatka, P. Kraft,M. Gehmlich, C. Tonk, J. Marek, New data on the Neoproterozoic–Cambriangeotectonic setting <strong>of</strong> the Teplá-Barr<strong>and</strong>ian174


Author's personal copy282K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286volcano-sedimentary successions: geochemistry, U-Pb zirconages, <strong>and</strong> provenance (Bohemian Massif, Czech Republic), Int.J. Earth Sci. 93 (2004) 742–757.[20] A. Dudek, The crystalline basement block <strong>of</strong> the outer Carpathiansin Moravia – Brunovistulicum, Rozpravy ČeskoslovenskéAkademie věd, Rada matematika prirodni ved 90(1980) 85.[21] J. Edel, T. Aïfa, M. Jelenska, M. Kodzialko-H<strong>of</strong>mokl, A.Zelazniewicz, Réaimantations des formations paléozoïquesdes Sudètes polonaises et courbe de dérive des pôles géomagnétiquesd’Europe du Carbonifère moyen au Jurassique moyen,C. R. Acad. Sci. Ser. IIa 325 (1997) 479–486.[22] J.B. Edel, K. Schulmann, F.V. Holub, Anticlockwise <strong>and</strong>clockwise rotations <strong>of</strong> the eastern Variscides accommodatedby dextral lithospheric wrenching: palaeomagnetic <strong>and</strong> <strong>structural</strong>evidence, J. Geol. Soc. 160 (2003) 209–218.[23] S. Faryad, M. Perraki, S. Vrána, P–T evolution <strong>and</strong> reactiontextures in retrogressed eclogites from Světlík, the MoldanubianZone (Czech Republic), Mineral. Petrol. 88 (2006) 297–319.[24] F. Finger, H.P. Steyrer, A tectonic model for the eastern.Variscides: indications from a chemical study <strong>of</strong> amphibolitesin. the southeastern Bohemian Massif, Geol. Carpath. 46 (1995)137–150.[25] F. Finger, A. von Quadt, U/Pb ages <strong>of</strong> zircons from a plagiogranite-gneissin the south-eastern Bohemian Massif, Austriafurtherevidence for an important Early Paleozoic rifting episodein the eastern Variscides, Schweiz, Mineral. Petrogr. Mitt.75 (1995) 265–270.[26] F. Finger, M. Roberts, B. Haunschmid, A. Schermaier, H.Steyrer, Variscan granitoids <strong>of</strong> central Europe: their typology,potential sources <strong>and</strong> tectonothermal relations, Mineral. Petrol.61 (1997) 67–96.[27] F. Finger, M. Tichomirowa, C. Pin, P. Hanžl, Relics <strong>of</strong> an Early-Panafrican metabasite-metarhyolite formation in the BrnoMassif, Moravia, Czech Republic, Int. J. Earth Sci. 89(2000) 328–335.[28] F. Finger, P. Hanžl, C. Pin, A. von Quadt, H.P. Steyrer, TheBrunovistulian: Avalonian Precambrian sequence at the easternend <strong>of</strong> the central European Variscides? Geol. Soc. LondonSpec. Publ. 179 (2000) 103–112.[29] F. Finger, A. Gerde, V. Janoušek, M. Ren, G. Riegier, Resolvingthe Variscan evolution <strong>of</strong> the Moldanubian sector <strong>of</strong> the BohemianMassif: the significance <strong>of</strong> the Bavarian <strong>and</strong> the Moravo–Moldanubian tectonometamorphic phases, J. Geosci. 52 (2007)9–28.[30] E. Franců, J. Franců, J. Kalvoda, H. Poelchau, J. Otava, Burial<strong>and</strong> uplift history <strong>of</strong> the Palaeozoic Flysch in the Variscanforel<strong>and</strong> basin (SE Bohemian Massif, Czech Republic), in:Continental collision <strong>and</strong> the tectonosedimentary evolution <strong>of</strong>forel<strong>and</strong>s, 2002, pp. 167–179.[31] J. Franěk, K. Schulmann, O. Lexa, Kinematic <strong>and</strong> rheologicalmodel <strong>of</strong> exhumation <strong>of</strong> high pressure granulites in the Variscanorogenic root: example <strong>of</strong> the Blanský les granulite,Bohemian Massif, Czech Republic, Mineral. Petrol. 86(2006) 253–276.[32] W. Frank, S. Hammer, F. Popp, S. Scharbert, M. Thöni,Isotopengeologische Neuergebnisse zur Entwicklungsgeschichteder Böhmischen Masse. Proterozoische Gesteinsserienund variscische Hauptorogenese, Osterr. Beitr. Meteorol. Geophys.3 (1990) 185–228.[33] W. Franke, Variscan plate tectonics in central Europe—currentideas <strong>and</strong> open questions, Tectonophysics 169 (1989) 221–228.[34] W. Franke, The mid-European segment <strong>of</strong> the Variscides:tectonostratigraphic units, terrane boundaries <strong>and</strong> plate tectonicevolution, Geol. Soc. London Spec. Publ. 179 (2000) 35–61.[35] W. Franke, The Variscan orogen in central Europe: construction<strong>and</strong> collapse, Geol. Soc. London Mem. 32 (2006) 333–343.[36] W. Franke, E. Stein, Exhumation <strong>of</strong> high-grade rocks in theSaxo-Thuringian Belt: geological constraints <strong>and</strong> geodynamicconcepts, Geol. Soc. London Spec. Publ. 179 (2000) 337–354.[37] G. Friedl, A. von Quadt, A. Oshner, F. Finger, Timing <strong>of</strong> theVariscan orogeny in the southern Bohemian Massif (NE Austria)deduced fro new U-Pb zircon <strong>and</strong> monazite dating, TerraNova 5 (1993) 235–236.[38] G. Friedl, F. Finger, N.J. McNaughton, I.R. Fletcher, Deducingthe ancestry <strong>of</strong> terranes: SHRIMP evidence for South AmericaderivedGondwana fragments in central Europe, Geology 28(2000) 1035–1038.[39] G. Friedl, F. Finger, J.-L. Paquette, A. von Quadt, N.J.McNaughton, I.R. Fletcher, Pre-Variscan geological eventsin the Austrian part <strong>of</strong> the Bohemian Massif deduced fromU–Pb zircon ages, Int. J. Earth Sci. 93 (2004) 802–823.[40] H. Fritz, R.D. Dallmeyer, F. Neubauer, Thick-skinned versusthin-skinned thrusting: Rheology controlled thrust propagation inthe Variscan collisional belt (The southeastern Bohemian Massif,Czech Republic - Austria), Tectonics 15 (1996) 1389–1413.[41] G. Fuchs, Zur Entwicklung der Böhmischen Masse, Jahrb.Geol. Bundelsansalt 129 (1976) 41–49.[42] G. Fuchs, Zur Diskussion um den Deckenbau der BöhmischenMasse, Jahrb. Geol. Bundelsansalt 129 (1986) 41–49.[43] A. Galle, J. Hladil, P. Isaacson, Middle Devonian biogeography<strong>of</strong> closing South Laurussia; North Gondwana Variscides; examplesfrom the Bohemian Massif (Czech Republic), with emphasison Horní Benešov, Palaios 10 (1995) 221–239.[44] D. Gebauer, M. Grünenfelder, U–Pb zircon <strong>and</strong> Rb–Sr mineraldating <strong>of</strong> eclogites <strong>and</strong> their country rocks. Example:Münchberg Gneiss Massif, Northeast Bavaria, Earth Planet.Sci. Lett. 42 (1979) 35–44.[45] A. Gerdes, Geochemische und thermische Modelle zur Frageder spätorogenen Granitgenese am Beispiel des SüdböhmischenBatholiths: Basaltisches Underplating oder Krustenstapelung,Unpublished Ph.D. thesis, University <strong>of</strong> Göttingen(1997) 113.[46] A. Gerdes, G. Wörner, A. Henk, Post-collisional granite generation<strong>and</strong> HT-LP metamorphism by radiogenic heating: theVariscan South Bohemian Batholith, J. Geol. Soc. 157 (2000)577–587.[47] A. Gerdes, G. Friedl, R. Parrish, F. Finger, High-resolutiongeochronology <strong>of</strong> Variscan granite emplacement–the SouthBohemian Batholith, J. Czech. Geol. Soc. 48 (2003) 53.[48] T. Gerya, B. Stockhert, Two-dimensional <strong>numerical</strong> modeling<strong>of</strong> tectonic <strong>and</strong> metamorphic histories at active continentalmargins, Int. J. Earth Sci. 95 (2006) 250–274.[49] A.J. Hartley, J. Otava, Sediment provenance <strong>and</strong> dispersal in adeep marine forel<strong>and</strong> basin: the Lower Carboniferous CulmBasin, Czech Republic, J. Geol. Soc. 158 (2001) 137–150.[50] P. Hasalová, P. Štípská, R. Powell, K. Schulmann, V. Janoušek,O. Lexa, Transforming mylonitic metagranite by open-systeminteractions during melt flow, J. Metamorph. Geol. 26 (2008)55–80.[51] V. Havlíček, Development <strong>of</strong> a linear sedimentary depressionexemplified by the Prague Basin (Ordovician–Middle Devonian;Barr<strong>and</strong>ian area–central Bohemia), Sb. Geologick. VedGeol. 35 (1981) 7–48.175


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 283[52] V. Havlíček, J. Vaněk, O. Fatka, Perunica microcontinent in theOrdovician (its position within the Mediterranean Province,series division, benthic <strong>and</strong> pelagic associations), Sb. Geologick.Ved Paleontol. 46 (1994) 23–56.[53] F. Holub, Ultrapotassic plutonic rocks <strong>of</strong> the durbachite seriesin the Bohemian Massif: petrology, geochemistry <strong>and</strong> petrogeneticinterpretation, J. Geol. Sci. Econ. Geol. Min. 31 (1997)5–26.[54] F. Holub, M. Klečka, D. Matějka, Moldanubian Zone VII. C. 3.Igneous activity, in : R. Dallmeyer, W. Franke, K. Weber (Eds.),Pre-Permian Geology <strong>of</strong> central <strong>and</strong> eastern Europe, Springer,Berlin, 1995, pp. 444–452.[55] F. Holub, A. Cocherie, P. Rossi, Radiometric dating <strong>of</strong> graniticrocks from the Central Bohemian Plutonic Complex (CzechRepublic): constraints on the chronology <strong>of</strong> thermal <strong>and</strong> tectonicevents along the Moldanubian-Barr<strong>and</strong>ian boundary, C. R.Acad. Sci. Ser. IIa 325 (1997) 19–26.[56] F. Holub, J. Machart, M. Manová, The Central BohemianPlutonic Complex: geology, chemical composition <strong>and</strong> geneticinterpretation, J. Geol. Sci. Econ. Geol. Min. 31 (1997) 27–50.[57] F. Hrouda, O. Krejčí, J. Otava, Magnetic fabric in folds <strong>of</strong> theeasternmost Rheno-Hercynian Zone, Phys. Chem. Earth Part A.25 (2000) 505–510.[58] F. Hrouda, Š. Taborská, K. Schulmann, J. Ježek, D. Dolejš,Magnetic fabric <strong>and</strong> rheology <strong>of</strong> co-mingled magmas in theNasavrky Plutonic Complex (E Bohemia); implications forintrusive strain regime <strong>and</strong> emplacement mechanism, Tectonophysics30 (1999) 93–111.[59] V. Janoušek, F. Holub, The causal link between HP/HT metamorphism<strong>and</strong> ultrapotassic magmatism in collisional orogens:case study from the Moldanubian Zone <strong>of</strong> the BohemianMassif, Proc. Geol. Assoc. 118 (2007) 75–86.[60] V. Janoušek, G. Rogers, D. Bowes, Sr-Nd isotopic constraintson the petrogenesis <strong>of</strong> the Central Bohemian Pluton, CzechRepublic, Int. J. Earth Sci. 84 (1995) 520–534.[61] V. Janoušek, D.R. Bowes, C.J.R. Braithwaite, G. Rogers,Micro<strong>structural</strong> <strong>and</strong> mineralogical evidence for limited involvement<strong>of</strong> magma mixing in the petrogenesis <strong>of</strong> a Hercynianhigh-K calc-alkaline intrusion: the Kozárovice granodiorite,Central Bohemian Pluton, Czech Republic, Trans. R. Soc.Edinburgh Earth Sci. 91 (2000) 15–26.[62] V. Janoušek, D.R. Bowes, G. Rogers, C.M. Farrow, E. Jelínek,Modelling diverse processes in the petrogenesis <strong>of</strong> a compositebatholith: the Central Bohemian Pluton, Central EuropeanHercynides, J. Petrol. 41 (2000) 511–543.[63] V. Janoušek, F. Holub, A. Gerdes, K-rich magmatism in theMoldanubian Unit, Bohemian Massif – a complex story featuringvariably enriched lithospheric mantle melts <strong>and</strong> theirinteraction with the crust, Geolines 16 (2003) 48–49.[64] V. Janoušek, C. Braithwaite, D. Bowes, A. Gerdes, Magmamixingin the genesis <strong>of</strong> Hercynian calc-alkaline granitoids: anintegrated petrographic <strong>and</strong> geochemical study <strong>of</strong> the Sázavaintrusion, Central Bohemian Pluton, Czech Republic, Lithos 78(2004) 67–99.[65] V. Janoušek, S. Vrána, V. Erban, K. Vokurka, M. Drábek,Metabasic rocks in the Varied Group <strong>of</strong> the Moldanubian Zone,southern Bohemia – their petrology, geochemical character <strong>and</strong>possible petrogenesis, J. Geosci. 53 (2008) 31–64.[66] V. Janoušek, F. Finger, M. Roberts, J. Frýda, C. Pin, D. Dolejš,Deciphering the petrogenesis <strong>of</strong> deeply buried granites: wholerockgeochemical constraints on the origin <strong>of</strong> largely undepletedfelsic granulites from the Moldanubian Zone <strong>of</strong> theBohemian Massif, Trans. R. Soc. Edinburgh: Earth Sci. 95(2004) 141–159.[67] V. Janoušek, A. Gerdes, S. Vrána, F. Finger, V. Erban, G. Friedl,C.J.R. Braithwaite, Low-pressure granulites <strong>of</strong> the Lišov Massif,southern Bohemia: Visean metamorphism <strong>of</strong> Late Devonianplutonic arc rocks, J. Petrol. 47 (2006) 705–744.[68] J. Kalvoda, O. Bábek, O. Fatka, J. Leichmann, R. Melichar, S.Nehyba, P. Špaček, Brunovistulian terrane (Bohemian Massif,central Europe) from Late Proterozoic to late Paleozoic: areview, Int. J. Earth Sci. 97 (2008) 497–518.[69] H. Klápová, J. Konopásek, K. Schulmann, Eclogites from theCzech part <strong>of</strong> the Erzgebirge multi-stage metamorphic <strong>and</strong><strong>structural</strong> evolution, J. Geol. Soc. 155 (1998) 567–583.[70] J. Konopásek, K. Schulmann, Variscan transpressional deformation<strong>and</strong> crustal folding in the Krkonose Mountains (northernmargin <strong>of</strong> the Bohemian Massif), in : K. Schulmann (Ed.),Palaeozoic orogenesis <strong>and</strong> crustal evolution <strong>of</strong> European lithosphere,Univerzita Karlova, Prague, Czech Republic, 1998, pp.279–280.[71] J. Konopásek, K. Schulmann, Contrasting Early Carboniferousfield geotherms: evidence for accretion <strong>of</strong> a thickened orogenicroot <strong>and</strong> subducted Saxothuringian crust (central EuropeanVariscides), J. Geol. Soc. 162 (2005) 463–470.[72] J. Konopásek, K. Schulmann, O. Lexa, Structural evolution <strong>of</strong>the central part <strong>of</strong> the Krusne hory (Erzgebirge) Mountains inthe Czech Republic - evidence for changing stress regimeduring Variscan compression, J. Struct. Geol. 23 (2001) 1373–1392.[73] J. Konopásek, K. Schulmann, V. Johan, Eclogite-facies metamorphismat the eastern margin <strong>of</strong> the Bohemian Massif –subduction prior to continental underthrusting? Eur. J. Mineral.14 (2002) 701–713.[74] J. Košler, M. Aftalion, D. Bowes, Mid–Late Devonian plutonicactivity in the Bohemian Massif: U–Pb zircon isotopic evidencefrom the Staré Sedlo <strong>and</strong> Mirotice gneiss complexes, CzechRepublic, Neues Jahrb. Miner. Monat. (1993) 417–431.[75] J. Košler, S. Kelley, D. Vance, M. Svojtka, Independent dating<strong>of</strong> cooling <strong>and</strong> decompression <strong>of</strong> high grade rocks in theSouthern Bohemian Massif with Ar-Ar, Sm-Nd <strong>and</strong> U-PbTechniques, J. Conf. Abs. 4 (1999) 39.[76] J. Košler, D.R. Bowes, J. Konopásek, J. Míková, Laser ablationICPMS dating <strong>of</strong> zircons in Erzgebirge orthogneisses: evidencefor Early Cambrian <strong>and</strong> Early Ordovician granitic plutonism inthe western Bohemian Massif, Eur. J. Mineral. 16 (2004) 15–22.[77] M. Košuličová, P. Špaček, Variations in the transient progradegeothermal gradient from chloritoid-staurolite equilibria: a casestudy from the Barrovian <strong>and</strong> Buchan-type domains in theBohemian Massif, J. Metamorph. Geol. 25 (2007) 19–36.[78] M. Košuličová, R. Montigny, K. Schulmann, P. Špaček, LatePaleozoic thermal overprints exemplified by Th-U-Pb monazite<strong>and</strong> K-Ar mica dating at the eastern margin <strong>of</strong> the BohemianMassif (West Sudets, Czech Republic), Bull. Soc. geol. France(2008), submitted for publication.[79] J. Kotková, Tectonometamorphic history <strong>of</strong> lower crust in theBohemian Massif: example <strong>of</strong> north Bohemian granulites,Czech Geol. Surv. Spec. Paper 2 (1993) 1–42.[80] J. Kotková, A. Kröner, W. Todt, J. Fiala, Zircon dating <strong>of</strong> NorthBohemian granulites, Czech Republic: further evidence for theLower Carboniferous high-pressure event in the BohemianMassif, Geol. Rundsch. 85 (1996) 154–161.[81] J. Kotková, A. Gerdes, R. Parrish, M. Novák, Clasts <strong>of</strong> Variscanhigh-grade rocks within Upper Visean conglomerates-cons-176


Author's personal copy284K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286traints on exhumation history from petrology <strong>and</strong> U-Pb chronology,J. Metamorph. Geol. 25 (2007) 781–801.[82] B. Kříbek, Z. Pouba, V. Skoček, J. Waldhausrová, Neoproterozoic<strong>of</strong> the Teplá-Barr<strong>and</strong>ian Unit as a part <strong>of</strong> the Cadomianorogenic belt: a review <strong>and</strong> correlation aspects, Bull. CzechGeol. Soc. 75 (2000) 175–196.[83] A. Kröner, A. Willner, Time <strong>of</strong> formation <strong>and</strong> peak <strong>of</strong> VariscanHP-HT metamorphism <strong>of</strong> quartz-feldspar rocks in the centralErzgebirge, Saxony, Germany, Contrib. Mineral. Petrol. 132(1998) 1–20.[84] A. Kröner, P. O’Brien, A. Nemchin, R. Pidgeon, Zircon agesfor high pressure granulites from South Bohemia, CzechRepublic, <strong>and</strong> their connection to Carboniferous high temperatureprocesses, Contrib. Mineral. Petrol. 138 (2000) 127–142.[85] A. Kröner, A. Willner, E. Hegner, A. Frischbutter, J. H<strong>of</strong>mann,R. Bergner, Latest Precambrian (Cadomian) zircon ages, Ndisotopic systematics <strong>and</strong> PT evolution <strong>of</strong> granitoid orthogneisses<strong>of</strong> the Erzgebirge, Saxony <strong>and</strong> Czech Republic, Int. J. EarthSci. 84 (1995) 437–456.[86] M. Krs, P. Pruner, Paleomagnetism, Palaeogeography <strong>of</strong> theVariscan Formations <strong>of</strong> the Bohemian Massif, J. Czech. Geol.Soc. 40 (1995) 3–45.[87] S. Lamb, L. Hoke, L. Kennan, J. Dewey, Cenozoic evolution <strong>of</strong>the central Andes in Bolivia <strong>and</strong> northern Chile, Geol. Soc.London Spec. Publ. 121 (1997) 237–264.[88] T. Liew, F. Finger, V. Höck, The Moldanubian granitoid plutons<strong>of</strong> Austria: chemical <strong>and</strong> isotopic studies bearing on theirenvironmental setting, Chem. Geol. 76 (1989) 41–55.[89] U. Linnemann, N.J. McNaughton, R.L. Romer, M. Gehmlich,K. Drost, C. Tonk, West African provenance for Saxo-Thuringia(Bohemian Massif): did Armorica ever leave pre-PangeanGondwana? U/Pb-SHRIMP zircon evidence <strong>and</strong> the Nd-isotopicrecord, Int. J. Earth Sci. 93 (2004) 683–705.[90] H. Maluski, F. Patočka, Geochemistry <strong>and</strong> 40 Ar/ 39 Ar geochronology<strong>of</strong> the mafic metavolcanic rocks from the RýchoryMountains complex (West Sudetes, Bohemian Massif): palaeotectonicsignificance, Geol. Mag. 134 (1997) 703–716.[91] H. Maluski, P. Rajlich, J. Souček, Pre-Variscan, Variscan <strong>and</strong>early Alpine thermo-tectonic history <strong>of</strong> the north-eastern BohemianMassif; an (super 40) Ar/ (super 39) Ar study, Geol.Rundsch. 84 (1995) 345–358.[92] H.J. Massonne, Early metamorphic evolution <strong>and</strong> exhumation<strong>of</strong> felsic high-pressure granulites from the north-western BohemianMassif, Springer Wien, 2006, pp. 177–202.[93] P. Matte, The Variscan collage <strong>and</strong> orogeny (480–290 Ma) <strong>and</strong>the tectonic definition <strong>of</strong> the Armorica microplate: a review,Terra Nova 13 (2001) 122–128.[94] P. Matte, H. Maluski, P. Rajlich, W. Franke, Terrane boundariesin the Bohemian Massif – result <strong>of</strong> large-scale Variscan shearing,Tectonophysics 177 (1990) 151.[95] S. Mazur, P. Aleks<strong>and</strong>rowski, R. Kryza, T. Oberc-Dziedzic, TheVariscan Orogen in Pol<strong>and</strong>, Geol. Q. 50 (2006) 89–118.[96] W.S. McKerrow, C. Mac Niocaill, P.E. Ahlberg, G. Clayton,C.J. Cleal, R.M.C. Eagar, The Late Palaeozoic relations betweenGondwana <strong>and</strong> Laurussia, Geol. Soc. London Spec. Publ.179 (2000) 9–20.[97] G. Medaris, E. Jelínek, Z. Mísař, Czech eclogites. Terranesettings, interpretation for Variscan tectonic evolution <strong>of</strong> theBohemian Massif, Eur. J. Mineral. 7 (1995) 7–28.[98] J. Nawrocki, The Devonian–Carboniferous platform paleomagneticdirections from the Silesian–Krakow area <strong>and</strong> theirimportance for Variscan paleotectonic reconstructions, Geol.Q. 37 (1993) 397–430.[99] P. O’Brien, Garnet zoning <strong>and</strong> reaction textures in overprintedeclogites, Bohemian Massif, European Variscides: a record <strong>of</strong>their thermal history during exhumation, Lithos 41 (1997) 119–133.[100] P. O’Brien, The fundamental Variscan problem: high-temperaturemetamorphism at different depths <strong>and</strong> high-pressuremetamorphism at different temperatures, in: W. Franke,V. Haak, O. Oncken, D. Tanner (Eds.), Orogenic Processes:Quantification <strong>and</strong> Modelling in the Variscan Belt, GeologicalSociety, London, 2000, pp. 369.[101] P. O’Brien, C. Rohr, M. Okrush, M. Patzak, Eclogite faciesrelics <strong>and</strong> a multistage breakdown in metabasites <strong>of</strong> the KTBpilot hole, NE Bavaria: implications for the Variscan tectonometamorphicevolution <strong>of</strong> the NW Bohemian Massif, Contrib.Mineral. Petrol. 112 (1992) 261–278.[102] P.J. O’Brien, D.A. Carswell, Tectonometamorphic evolution <strong>of</strong>the Bohemian Massif - evidence from high-pressure metamorphicrocks, Geol. Rundsch. 82 (1993) 531–555.[103] P.J. O’Brien, S. Vrána, Eclogites with a short-lived granulitefaciesoverprint in the Moldanubian Zone, Czech Republic –petrology, geochemistry <strong>and</strong> diffusion modeling <strong>of</strong> garnetzoning, Geol. Rundsch. 84 (1995) 473–488.[104] O. Oncken, C. von Winterfeld, U. Dittmar, Accretion <strong>of</strong> a riftedpassive margin: the Late Paleozoic Rhenohercynian fold <strong>and</strong>thrust belt (Middle European Variscides), Tectonics 18 (1999)75–91.[105] O. Oncken, D. Hindle, J. Kley, K. Elger, P. Victor, K. Schemmann,Deformation <strong>of</strong> the central Andean Upper Plate System—-facts, fiction, <strong>and</strong> constraints for plateau models, in : O. Oncken,et al. (Eds.), The Andes, Springer Verlag, 2006, pp. 3–27.[106] F. Paris, Early Palaezoic palaeography <strong>of</strong> northern Gondwanaregions, Acta Univ. Carol. Geol. 42 (1998) 473–483.[107] F. Patočka, J. Valenta, Geochemistry <strong>of</strong> the Late Devonianintermediate to acid metavolcanic rocks from the southern part<strong>of</strong> the Vrbno Group, the Jeseniky Mts. (Moravo-Silesian Belt,Bohemian Massif, Czech Republic): paleotectonic implications,Geolines 4 (1996) 42–54.[108] F. Patočka, P. Vlašímský, K. Blechová, Geochemistry <strong>of</strong> EarlyPaleozoic volcanics <strong>of</strong> the Barr<strong>and</strong>ian Basin (Bohemian Massif,Czech Republic): implications for Paleotectonic reconstructions,Jahrb. Geol. Bundelsansalt 136 (1993) 871–894.[109] K. Petrakakis, Evolution <strong>of</strong> Moldanubian rocks in Austria:review <strong>and</strong> synthesis, J. Metamorph. Geol. 15 (1997) 203–222.[110] P. Pitra, J.P. Burg, M. Giraud, Late-Variscan strike-slip tectonicsbetween the Teplá-Barr<strong>and</strong>ian <strong>and</strong> Moldanubian terranes(Czech Bohemian Massif): petro<strong>structural</strong> evidence, J. Geol.Soc. 156 (1999) 1003–1020.[111] P. Pitra, J.P. Burg, K. Schulmann, P. Ledru, Late Orogenicextension in the Bohemian Massif – petro<strong>structural</strong> evidence inthe Hlinsko Region, Geodin. Acta 7 (1994) 15–30.[112] M. Racek, P. Štípská, P. Pitra, K. Schulmann, O. Lexa, Metamorphicrecord <strong>of</strong> burial <strong>and</strong> exhumation <strong>of</strong> orogenic lower <strong>and</strong>middle crust: a new tectonothermal model for the Drosendorfwindow (Bohemian Massif, Austria), Mineral. Petrol. 86(2006) 221–251.[113] P. Rajlich, J. Synek, M. Sarbach, K. Schulmann, Hercynianthrustrelated shear zones <strong>and</strong> deformation <strong>of</strong> the Varied Groupon the contact <strong>of</strong> granulites/southern Moldanubian, BohemianMassif, Int. J. Earth Sci. 75 (1986) 665–683.177


Author's personal copyK. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286 285[114] J. Schäfer, W. Dörr, Ergebnisse zur Liefergebietsanbindung dessaxothuringischen Flyschs durch St<strong>of</strong>fbest<strong>and</strong>suntersuchungenund U–Pb-Datierungen detrischer Zirkone, Terra Nostra 984(1994) 95–96.[115] D. Scheuvens, G. Zulauf, Exhumation, strain localization, <strong>and</strong>emplacement <strong>of</strong> granitoids along the western part <strong>of</strong> the centralBohemian shear zone (Bohemian Massif), Int. J. Earth Sci. 89(2000) 617–630.[116] E. Schmädicke, M. Okrusch, W. Schmidt, Eclogite-facies rocksin the Saxonian Erzgebirge Germany: high pressure metamorphismunder contrasting P-T conditions, Contrib. Mineral.Petrol. 1992 (1992) 226–241.[117] D. Schneider, S. Zahniser, J. Glascock, S. Gordon, M. Manecki,Thermochronology <strong>of</strong> the West Sudetes (Bohemian Massif):Rapid <strong>and</strong> repeated eduction in the eastern Variscides, Pol<strong>and</strong><strong>and</strong> Czech Republic, Am. J. Sci. 306 (2006) 846.[118] K. Schulmann, Fabric <strong>and</strong> kinematic study <strong>of</strong> the Bites Orthogneiss(Southwestern Moravia) – result <strong>of</strong> large-scale northeastwardshearing parallel to the Moldanubian MoravianBoundary, Tectonophysics 177 (1990) 229–244.[119] K. Schulmann, R. Gayer, A model for a continental accretionarywedge developed by oblique collision; the NE BohemianMassif, J. Geol. Soc. 157 Part 2 (2000) 401–416.[120] K. Schulmann, P. Ledru, A. Autran, R. Melka, J. Lardeaux, M.Urban, M. Lobkowicz, Evolution <strong>of</strong> nappes in the easternmargin <strong>of</strong> the Bohemian Massif: a kinematic interpretation,Int. J. Earth Sci. 80 (1991) 73–92.[121] K. Schulmann, R. Melka, M. Lobkowicz, P. Ledru, J. Lardeaux,A. Autran, Contrasting styles <strong>of</strong> deformation during progressivenappe stacking at the southeastern margin <strong>of</strong> the BohemianMassif (Thaya Dome), J. Struct. Geol. 16 (1994) 355–370.[122] K. Schulmann, A. Kroner, E. Hegner, I. Wendt, J. Konopásek,O. Lexa, P. Štípská, Chronological constraints on the preorogenichistory, burial <strong>and</strong> exhumation <strong>of</strong> deep-seatedrocks along the eastern margin <strong>of</strong> the Variscan Orogen, BohemianMassif, Czech Republic, Am. J. Sci. 305 (2005) 407–448.[123] K. Schulmann, O. Lexa, P. Štípská, M. Racek, L. Tajčmanová,J. Konopásek, J.B. Edel, A. Peschler, J. Lehmann, Verticalextrusion <strong>and</strong> horizontal channel flow <strong>of</strong> orogenic lower crust:key exhumation mechanisms in large hot orogens? J. Metamorph.Geol. 26 (2008) 273–297.[124] W. Siebel, C. Shang, E. Reitter, J. Rohrmuller, K. Breiter, Twodistinctive granite suites in the SW Bohemian Massif <strong>and</strong> theirrecord <strong>of</strong> emplacement: Constraints from geochemistry <strong>and</strong>zircon 207 Pb/ 206 Pb chronology, J. Petrol. 49 (2008) 1853–1872.[125] W. Siebel, H. Raschka, W. Irber, H. Kreuzer, K. Lenz, I. Wendt,Early Palaeozoic acid magmatism in the Saxothuringian Belt:new insights from a geochemical <strong>and</strong> isotopic study <strong>of</strong> orthogneisses<strong>and</strong> metavolcanic rocks from the Fichtelgebirge, SEGermany, J. Petrol. 38 (1997) 203–229.[126] P. Špaček, J. Kalvoda, E. Franců, R. Melichar, Variation <strong>of</strong>deformation mechanisms within the progressive-retrogressivemylonitization cycle <strong>of</strong> limestones: Brunovistulian sedimentarycover (the Variscan orogeny <strong>of</strong> the southeastern BohemianMassif), Geol. Carpath. 52 (2001) 263–275.[127] P. Štípská, K. Schulmann, Inverted metamorphic zonation in abasement-derived nappe sequence; eastern margin <strong>of</strong> the BohemianMassif, Geol. J. 30 (1995) 385–413.[128] P. Štípská, R. Powell, Constraining the PT path <strong>of</strong> a MORBtypeeclogite using pseudosections, garnet zoning <strong>and</strong> garnetclinopyroxenethermometry: an example from the BohemianMassif, J. Metamorph. Geol. 23 (2005) 725–743.[129] P. Štípská, R. Powell, Does ternary feldspar constrain themetamorphic conditions <strong>of</strong> high-grade meta-igneous rocks?Evidence from orthopyroxene granulites, Bohemian Massif,J. Metamorph. Geol. 23 (2005) 627–647.[130] P. Štípská, K. Schulmann, A.B. Thompson, J. Ježek, A. Kröner,Thermo-mechanical role <strong>of</strong> a Cambro-Ordovician paleoriftduring the Variscan collision: the NE margin <strong>of</strong> the BohemianMassif, Tectonophysics 332 (2001) 239–253.[131] P. Štípská, K. Schulmann, A. Kröner, Vertical extrusion <strong>and</strong>middle crustal spreading <strong>of</strong> omphacite granulite: a model <strong>of</strong>syn-convergent exhumation (Bohemian Massif, Czech Republic),J. Metamorph. Geol. 22 (2004) 179–198.[132] P. Štípská, P. Pitra, R. Powell, Separate or shared metamorphichistories <strong>of</strong> eclogites <strong>and</strong> surrounding rocks? An example fromthe Bohemian Massif, J. Metamorph. Geol. 24 (2006) 219–240.[133] P. Štípská, K. Schulmann, R. Powell, Contrasting metamorphichistories <strong>of</strong> lenses <strong>of</strong> high-pressure rocks <strong>and</strong> host migmatiteswith a flat orogenic fabric (Bohemian Massif, Czech Republic):a result <strong>of</strong> tectonic mixing within horizontal crustal flow? J.Metamorph. Geol. 26 (2008) 623–646.[134] H. Stosch, G. Lugmair, Geochemistry <strong>and</strong> evolution <strong>of</strong> MORBtypeeclogites from the Münchberg Massif, southern Germany,Earth Planet. Sci. Lett. 99 (1990) 230–249.[135] L. Strnad, M. Mihaljevič, Sedimentary provenance <strong>of</strong> Mid-Devonian clastic sediments in the Teplá-Barr<strong>and</strong>ian Unit(Bohemian Massif): U–Pb <strong>and</strong> Pb–Pb geochronology <strong>of</strong> detritalzircons by laser ablation ICP-MS, Mineral. Petrol. 84 (2005)47–68.[136] F.E. Suess, Die moravischen Fenster und ihre Beziehung zumGrundgebirge des Hohen Gesenkes, Denkschr. K. K. Akad.Wiss. 83 (1912) 541–631.[137] F.E. Suess, Intrusionstektonik und W<strong>and</strong>ertektonik im variszischenGrundgebirge, Verlag von Gebrüder Borntraeger, Berlin,1926p. 268.[138] M. Svojtka, J. Košler, Z. Venera, Dating granulite-facies structures<strong>and</strong> the exhumation <strong>of</strong> lower crust in the MoldanubianZone <strong>of</strong> the Bohemian Massif, Int. J. Earth Sci. 91 (2002) 373–385.[139] J. Tait, V. Bachtadse, H. S<strong>of</strong>fel, Eastern Variscan fold belt;paleomagnetic evidence for oroclinal bending, Geology 24(1996) 871–874.[140] J.A. Tait, V. Bachtadse, W. Franke, H.C. S<strong>of</strong>fel, Geodynamicevolution <strong>of</strong> the European Variscan fold belt: palaeomagnetic<strong>and</strong> geological constraints, Geol. Rundsch. 86 (1997) 585–598.[141] J. Tait, V. Bachtadse, J. Dinares-Turell, Paleomagnetism <strong>of</strong>Siluro-Devonian sequences, NE Spain, J. Geophys. Res. 105(2000) 23,595–23,603.[142] L. Tajčmanová, J. Konopásek, K. Schulman, Thermal evolution<strong>of</strong> the orogenic lower crust during exhumation within a thickenedMoldanubian root <strong>of</strong> the Variscan belt <strong>of</strong> central Europe,J. Metamorph. Geol. 24 (2006) 119–134.[143] L. Tajčmanová, J. Konopásek, J. Connolly, Diffusion-controlleddevelopment <strong>of</strong> silica-undersaturated domains in felsicgranulites <strong>of</strong> the Bohemian Massif (Variscan belt <strong>of</strong> centralEurope), Contrib. Mineral. Petrol. 153 (2007) 237–250.[144] M. Tichomirowa, Die Gneise des Erzgebirges-hochmetamorpheÄquivalente von Neoproterozoisch-frühpaläozoischenGrauwacken und Granitoiden der Cadomiden, TechnischeUniversität Bergakademie Freiberg, 2003.[145] H. Timmermann, V. Štědrá, A. Gerdes, S.R. Noble, R.R.Parrish, W. Dörr, The problem <strong>of</strong> dating high-pressure metamorphism:a U-Pb isotope <strong>and</strong> geochemical study on eclogites178


Author's personal copy286K. Schulmann et al. / C. R. Geoscience 341 (2009) 266–286<strong>and</strong> related rocks <strong>of</strong> the Marianské Lázně Complex, CzechRepublic, J. Petrol. 45 (2004) 1311–1338.[146] H. Timmermann, W. Dörr, E. Krenn, F. Finger, G. Zulauf,Conventional <strong>and</strong> in situ geochronology <strong>of</strong> the Teplá crystallineunit, Bohemian Massif: implications for the processes involvingmonazite formation, Int. J. Earth Sci. 95 (2006) 629–647.[147] R.B. Trumbull, U. Riller, O. Oncken, E. Scheuber, K. Munier, F.Hongn, The time-space distribution <strong>of</strong> Cenozoic volcanism inthe South-Central Andes: a new data compilation <strong>and</strong> sometectonic implications, in : O. Oncken, et al. (Eds.), The Andes,Springer Verlag, 2006, pp. 29–43.[148] C. Vellmer, K. Wedepohl, Geochemical characterization <strong>and</strong>origin <strong>of</strong> granitoids from the South Bohemian Batholith inLower Austria, Contrib. Mineral. Petrol. 118 (1994) 13–32.[149] K. Verner, J. Žák, R. Nahodilová, F. Holub, Magmatic fabrics <strong>and</strong>emplacement <strong>of</strong> the cone-sheet-bearing Knížecí Stolec durbachiticpluton (Moldanubian Unit, Bohemian Massif): implicationsfor mid-crustal reworking <strong>of</strong> granulitic lower crust in thecentral European Variscides, Int. J. Earth Sci. 97 (2008) 19–33.[150] A. von Quadt, D. Gebauer, Evolution <strong>of</strong> eclogitic rocks inthe Erzgebirge: a conventional <strong>and</strong> SHRIMP U-Pb zircon<strong>and</strong> Sm-Nd study, Acta Univ. Carol. Geol. 42 (1998) 324–1324.[151] S. Vrána, Polyphase shear folding <strong>and</strong> thrusting in the Moldanubicum<strong>of</strong> southern Bohemia, Vestn. Ustr. Ust. Geol. 51(1979) 75–86.[152] S. Vrána, M. Novák, Petrology <strong>and</strong> geochemistry <strong>of</strong> granuliteclasts in the Visean Luleč conglomerate, Kulm in centralMoravia, Czech Republic, Bull. Czech Geol. Soc. 75 (2000)405–413.[153] J.I. Wendt, A. Kröner, J. Fiala, W. Todt, U-Pb zircon <strong>and</strong> Sm-Nddating <strong>of</strong> Moldanubian HP/HT granulites from South Bohemia,Czech Republic, J. Geol. Soc. 151 (1994) 83–90.[154] A. Willner, K. Roetzler, W. Maresch, Pressure-temperature <strong>and</strong>fluid evolution <strong>of</strong> quartzo-feldspathic metamorphic rocks with arelic high-pressure, granulite-facies history from the centralErzgebirge (Saxony, Germany), J. Petrol. 38 (1997) 307–336.[155] V. Žáček, Garnets <strong>and</strong> metamorphic evolution <strong>of</strong> the Teplácrystalline complex, western Bohemia, Zentralbl. Geologie.Palaontol. 1 (1994) 847–856.[156] V. Žáček, J. Cháb, Metamorphism in the Tepla upl<strong>and</strong>, Bohemianmassif, Czech Republic (preliminary report), Bull. CzechGeol. Soc. 68 (1993) 33.[157] J. Žák, F. Holub, K. Verner, Tectonic evolution <strong>of</strong> a continentalmagmatic arc from transpression in the upper crust to exhumation<strong>of</strong> mid-crustal orogenic root recorded by episodicallyemplaced plutons: the Central Bohemian Plutonic Complex(Bohemian Massif), Int. J. Earth Sci. 94 (2005) 385–400.[158] J. Žák, K. Schulmann, F. Hrouda, Multiple magmatic fabrics inthe Sázava pluton (Bohemian Massif, Czech Republic): a result<strong>of</strong> superposition <strong>of</strong> wrench-dominated regional transpressionon final emplacement, J. Struct. Geol. 27 (2005) 805–822.[159] J. Žák, F. Holub, V. Kachlík, Magmatic stoping as an importantemplacement mechanism <strong>of</strong> Variscan plutons: evidence fromro<strong>of</strong> pendants in the Central Bohemian Plutonic Complex(Bohemian Massif), Int. J. Earth Sci. 95 (2006) 771–789.[160] G. Zulauf, Von der Anchizone bis zur Eklogitfazies: AngekippteKrustenpr<strong>of</strong>ile als Folge der cadomischen und variscischenOrogenese im Teplá-Barr<strong>and</strong>ium (Böhmische Masse),Geotekton. Forsch. 89 (1997) 1–302.[161] G. Zulauf, Structural style, deformation mechanisms <strong>and</strong> paleodifferentialstress along an exposed crustal section: constraintson the rheology <strong>of</strong> quartz<strong>of</strong>eldspathic rocks at supra- <strong>and</strong>infra<strong>structural</strong> levels (Bohemian Massif), Tectonophysics 332(2001) 211–237.[162] G. Zulauf, C. Bues, W. Dörr, Z. Vejnar, 10 km Minimum throwalong the West Bohemian shear zone: evidence for dramaticcrustal thickening <strong>and</strong> high topography in the Bohemian Massif(European Variscides), Int. J. Earth Sci. 91 (2002) 850–864.[163] G. Zulauf, W. Dörr, J. Fiala, J. Kotková, H. Maluski, P. Valverde-Vaquero, Evidence for high-temperature diffusional creep preservedby rapid cooling <strong>of</strong> lower crust (North Bohemian shearzone, Czech Republic), Terra Nova 14 (2002) 343–354.179


J. metamorphic Geol., 2011, 29, 79–102 doi:10.1111/j.1525-1314.2010.00906.xHeat sources <strong>and</strong> trigger mechanisms <strong>of</strong> exhumation <strong>of</strong> HPgranulites in Variscan orogenic rootO. LEXA, 1 K. SCHULMANN, 2 V. JANOUŠEK, 1,3 P. ŠTÍPSKÁ, 2 A. GUY 2,4 AND M. RACEK 1,31 Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Faculty <strong>of</strong> Science, Charles University in Prague, Albertov 6, 128 43 Prague 2,Czech Republic (lexa@natur.cuni.cz)2 Institute de Physique de Globe, UMR 7516, École et Observatoire de Science de la Terre, Université de Strasbourg, 1 RueBlessig, Strasbourg 67084, France3 Czech Geological Survey, Klárov 3, 118 21 Prague 1, Czech Republic4 Department <strong>of</strong> Geophysics, Faculty <strong>of</strong> Mathematics <strong>and</strong> Physics, Charles University in Prague, V Holešovičkách 3, 180 00Prague 8, Czech RepublicABSTRACTThe structure <strong>of</strong> the Moldanubian domain is marked by felsic granulites <strong>of</strong> Ordovician protolith ageforming the cores <strong>of</strong> domes that are separated from mid-crustal Neoproterozoic <strong>and</strong> Palaeozoicmetasedimentary rocks that occur in synclines by a late Ordovician to Silurian metabasic unit. Reflection<strong>and</strong> refraction seismic sections combined with gravity inversion <strong>modelling</strong> suggest the presence <strong>of</strong> a lowdensity layer at the bottom <strong>of</strong> the crust (interpreted as felsic granulite) overlain by a denser layer(interpreted as amphibolite) with layers <strong>of</strong> intermediate density at the top (interpreted as metasedimentaryrocks). It is proposed that the granulite domes surrounded by middle crustal rocks reflect transposedhorizontal layering originally similar to that preserved in the deep crust <strong>and</strong> imaged by the geophysicalsurveys. This geological <strong>and</strong> geophysical structure is considered to be a result <strong>of</strong> Vise´an gravityredistribution initiated by radioactive heating <strong>of</strong> felsic crust tectonically emplaced at the bottom <strong>of</strong> aPalaeozoic orogenic root. The radioactive layer with heat production <strong>of</strong> 4 lW m )3 correspondsgeochemically <strong>and</strong> isotopically to Ordovician felsic metaigneous rocks <strong>of</strong> the Saxothuringian domain thathave been emplaced at Moho depth under thickened crust during late Devonian–early Carboniferouscontinental subduction. Part <strong>of</strong> the continental crust continued to be subducted <strong>and</strong> produced fluids ⁄low-volume melts which directly contaminated <strong>and</strong> enriched the local lithospheric mantle by lithophileelements, most notably Cs, Rb, Li, Pb, U, Th <strong>and</strong> K. Thermal incubation <strong>of</strong> 10–15 Myr was sufficient toheat <strong>and</strong> convert the underplated felsic layer into granulites via dehydration melting <strong>and</strong> meltsegregation. The process <strong>of</strong> melt loss was responsible for the removal <strong>of</strong> radioactive elements <strong>and</strong> forswitching <strong>of</strong>f the heat at the beginning <strong>of</strong> the exhumation process. At the same time, the metasomatizedunderlying mantle was heated producing characteristic ultrapotassic magmas. Gravitational instabilitywas then induced by the density contrast between the light granulites <strong>and</strong> the overlaying denser maficlower crustal layer <strong>and</strong> a viscosity drop related to thermal weakening <strong>and</strong> partial melting <strong>of</strong> the latter.Key words: crustal structure; exhumation <strong>of</strong> lower crust; heat sources; radioactive heating.Mineral abbreviations: bi, biotite; mu, muscovite; g, garnet; ky, kyanite; sill, sillimanite; cd, cordierite; pl,plagioclase; ksp, K-feldspar; liq, granitic liquid; o, omphacitic clinopyroxene; ilm, ilmenite; ru, rutile.INTRODUCTIONThe Variscan Bohemian Massif, Czech Republic, ischaracterized by the occurrence <strong>of</strong> large high-pressure(HP) granulite bodies within the central high-gradepart <strong>of</strong> the orogen. These granulite bodies are mainlycomposed <strong>of</strong> alkali feldspar-garnet-kyanite granulites,eclogites <strong>and</strong> fragments <strong>of</strong> mantle peridotites surroundedby mid-crustal rocks. The exhumation <strong>and</strong>emplacement <strong>of</strong> the whole assemblage is commonlyinterpreted in terms <strong>of</strong> two contrasting models. In thefirst model, granulite massifs are inferred to be allochthonous,representing klippen <strong>of</strong> far travelled nappesrooted in the central part <strong>of</strong> the Bohemian Massif (e.g.Franke, 2000), whereas in the second model, granulitescorrespond to eroded windows <strong>of</strong> the orogenic infrastructurethat have been vertically extruded to midcrustallevels from lower crustal depths (e.g. Sˇtı´pska´et al., 2004; Schulmann et al., 2005). The latter modelis analogous to that proposed for the Saxonian granuliteMassif by Behr (1978) <strong>and</strong> Weber (1984).The problem <strong>of</strong> exhumation <strong>of</strong> granulite lower crusthas been discussed by a number <strong>of</strong> authors in relationto Archean <strong>and</strong> Mesoproterozoic orogenic belts (e.g.Perchuk, 1989). In these terranes, the emplacement <strong>of</strong>hot granulite lower crust has been interpreted in terms<strong>of</strong> gravity overturn driven by an inverted densitypr<strong>of</strong>ile (e.g. Roering et al., 1992), whereby heavy maficÓ 2010 Blackwell Publishing Ltd 79181


80 O. LEXA ET AL.rocks (greenstone belts) rest upon light intermediate t<strong>of</strong>elsic TTG rocks [Tonalite-Trondhjemite-Granodioritegneisses]. Perchuk (1989) proposed that such crustalgravity redistributions are triggered by heat fluxresulting from large-scale plume tectonics.Gerya et al. (2001) proposed a conceptually similarmodel <strong>of</strong> crustal-scale gravity redistribution to explainthe emplacement <strong>of</strong> felsic orogenic crust into supracrustallevels <strong>of</strong> Palaeozoic orogens. In contrast toArchean <strong>and</strong> Mesoproterozoic orogenic belts, the heatsource is inferred to be bulk radioactive heat productionwithin a granodioritic lower crustal layer locatedat the bottom <strong>of</strong> the crustal pile (Gerya et al., 2002,2004). These authors suggested that the gravitationalinstability <strong>of</strong> doubly stacked lithologically heterogeneouscrust is related to an initial density contrast <strong>of</strong>dissimilar intercalated layers enhanced by hightemperaturephase transformations.In this article, it is argued that the granulite bodiesabundant in the central parts <strong>of</strong> the Bohemian Massifcould have originated through a similar kind <strong>of</strong>gravity redistribution enhanced by lateral forcing.Using gravity inversion <strong>modelling</strong>, we demonstratethat relics <strong>of</strong> an original stratification with densityinversion are still preserved in the Variscan crust <strong>of</strong>the Bohemian Massif, with a high density mafic middlecrustal layer resting upon 10 km <strong>of</strong> felsic crust(Guy et al., 2010). It is also shown that the internalstructure within each <strong>of</strong> the granulite massifs forms apattern consistent with deceleration <strong>of</strong> verticallyemplaced diapiric-like bodies <strong>and</strong> horizontal spreading<strong>of</strong> low-viscosity, partially molten rocks at supracrustallevels. Geochemical <strong>and</strong> geochronological data areused to argue that the felsic lower crust (FLC) is anallochthonous body emplaced underneath the preexistingmafic lower crust during late Devonian–earlyCarboniferous continental subduction. The time-scales<strong>of</strong> thermal processes driving the gravity redistributionare estimated for two end-member scenarios: (i) amodel <strong>of</strong> radioactive heat production solely in theFLC <strong>and</strong> ⁄ or mantle; <strong>and</strong> (ii) a large-scale thermalanomaly in the mantle caused by deep mantle processessuch as tectospheric root delamination (Deweyet al., 1993) or slab break <strong>of</strong>f (Chemenda et al., 2000)as proposed for the Bohemian Massif by Janousˇek &Holub (2007). Finally, we <strong>of</strong>fer a dynamic <strong>numerical</strong>model which is consistent with the available geological<strong>and</strong> geophysical data <strong>and</strong> allows assessment <strong>of</strong> exhumationmetamorphic paths <strong>of</strong> the granulite rocks inthe core <strong>of</strong> the Bohemian Massif.GEOLOGICAL SETTINGTraditionally the Bohemian Massif has been subdividedinto four main tectonic domains, which are, fromthe west to the east (Fig. 1): the Saxothuringian, theTepla´-Barr<strong>and</strong>ian, the Moldanubian <strong>and</strong> the Bruniadomains. The Saxothuringian domain is regarded as aNeoproterozoic continental block accreted in theDevonian to, <strong>and</strong> partly subducted under, the moreeasterly Tepla´-Barr<strong>and</strong>ian (suprastructure or orogeniclid) <strong>and</strong> Moldanubian (infrastructure or deep orogenicroot) continental domains <strong>of</strong> the Variscan orogen(Schulmann et al., 2009). The oceanic Tepla´ suture, leftover after the closure <strong>of</strong> Saxothuringian ocean, is representedby Devonian to Carboniferous HP <strong>and</strong> ultrahigh-pressure rocks, gabbros <strong>and</strong> mantle fragmentslocated between the Saxothuringian <strong>and</strong> the Tepla´-Barr<strong>and</strong>ian domains (Mlcˇoch & Konopa´sek, 2010 <strong>and</strong>references therein). The whole system is bounded fromthe east by a Neoproterozoic Brunia promontory,which indented the Moldanubian orogenic root duringthe early Carboniferous (Schulmann et al., 2008).Moldanubian domainThe Moldanubian domain is composed mostly <strong>of</strong> theorogenic middle crustal unit subdivided into two lithostratigraphicgroups: the Monotonous Group <strong>of</strong> probableNeoproterozoic age (Friedl et al., 2004), <strong>and</strong> theVaried Group, at least partly <strong>of</strong> early Palaeozoic affinity(e.g. Janousˇek et al., 2008). Their mutual contact isdefined by uniformly deformed bodies <strong>of</strong> granitic gneiss<strong>of</strong> Neoproterozoic to Mesoproterozoic age (Wendtet al., 1993; Friedl et al., 2004). The Varied Group iscomposed <strong>of</strong> metasedimentary rocks intercalated withamphibolites, quartzites, marbles <strong>and</strong> calc silicates. TheMonotonous Group consists <strong>of</strong> paragneisses intercalatedwith orthogneiss bodies. The lower part <strong>of</strong> thelatter group is characterized by a thick sequence <strong>of</strong>amphibolites <strong>and</strong> metagabbros (Racek et al., 2006)locally containing eclogites (OÕBrien & Vra´na, 1995).The orogenic lower crust is represented by the Gfo¨ hlUnit (Fuchs, 1976), which comprises felsic <strong>and</strong> intermediateHP granulites accompanied by A type eclogites,garnet pyroxenites <strong>and</strong> peridotites (Medaris et al.,1995), amphibolites accompanied by Mid-Ocean RidgeBasalt-type eclogites (Sˇtı´pska´ & Powell, 2005a) <strong>and</strong>anatectic Gfo¨ hl orthogneisses.Today the structure <strong>of</strong> the Moldanubian domain ischaracterized by alternation <strong>of</strong> three, NE–SW trendingorogenic lower crustal <strong>and</strong> middle crustal belts (Fig. 1;Behr, 1978; Finger et al., 2007). The first occurs mostlyin south Bohemia, the second belt follows the easternboundary between the Moldanubian domain <strong>and</strong>Brunia continent <strong>and</strong> the third is a less extensive belt <strong>of</strong>granulite rocks tracing the contact between the Tepla´-Barr<strong>and</strong>ian <strong>and</strong> the Saxothuringian domains The lastlarge occurrence <strong>of</strong> granulites is represented by theNE–SW trending domal structure <strong>of</strong> the Saxoniangranulite Massif (e.g. Franke, 2000).Recent studies <strong>of</strong> the principal granulite bodiesreveal a systematic lithotectonic pattern marked bygranulite lenses rimmed by amphibolite–gabbro belts(AGB) <strong>of</strong> the local ÔBegleitÕ (= accompanying) series<strong>and</strong> Gfo¨ hl orthogneiss (Fig. 2). The boundary betweenthe AGB <strong>and</strong> the Monotonous Group is not clearlydefined, <strong>and</strong> hence these basic rocks are sometimesÓ 2010 Blackwell Publishing Ltd182


HEAT SOURCES AND EXHUMATION MECHANISMS 81Fig. 1. Simplified geological map <strong>of</strong> the Bohemian Massif (modified after Franke, 2000). CBPC: Central Bohemian Plutonic Complex;CMP: Central Moldanubian Pluton. The lower left insert shows the position <strong>of</strong> the Bohemian Massif in the European Variscides(modified after Edel et al., 2003). RH: Rhenohercynian zone; ST: Saxothuringian Zone; M: Moldanubian Zone; B: Brunia Continent;L: Lugian domain.attributed to the lowermost part <strong>of</strong> the MonotonousGroup (Racek et al., 2006) <strong>and</strong> sometimes to theGfo¨ hl Unit (Fuchs, 1976). In addition, the generalpattern is complicated by the presence <strong>of</strong> large accumulations<strong>of</strong> amphibolites within the Varied <strong>and</strong>Monotonous groups without clear tectonostratigraphicaffinity. The amphibolites <strong>and</strong> gabbros inLower Austria reach several kilometres in thickness,but locally narrow to several hundred metres (e.g.Tajcˇmanova´ et al., 2010) <strong>and</strong> in some places they arecut out completely (Franeˇk et al., 2006). Because <strong>of</strong>the regional significance <strong>of</strong> these rocks, they are interpretedby some authors as a relict <strong>of</strong> the Silurian (c.430 Ma) oceanic crust (Finger & von Quadt, 1995) oras the result <strong>of</strong> late Ordovician igneous activity relatedto thinning <strong>of</strong> continental crust (Schulmann et al.,2009). In contrast, the Gfo¨ hl orthogneiss yields dominantlyNeoproterozoic protolith U–Pb zircon ages(e.g. 550 ± 1 Ma; Schulmann et al., 2005) accompaniedby early Ordovician ages (e.g. 488 ± 6 Ma;Friedl et al., 2004). The Moldanubian felsic–intermediategranulites reveal a more complex spectrum <strong>of</strong>protolith U–Pb zircon ages clustering at c. 360, c. 400<strong>and</strong> 470–450 Ma (Kro¨ ner et al., 2000; Friedl et al.,2004; Janousˇek et al., 2004b).CURRENT DEEP STRUCTURE OF THE FORMERVARISCAN ROOTA new model <strong>of</strong> the structure <strong>and</strong> composition <strong>of</strong>Variscan crust in the Bohemian Massif was recentlyproposed by Guy et al. (2010) based on 3D gravity<strong>modelling</strong>, geological data <strong>and</strong> seismic refraction <strong>and</strong>reflection sections (Tomek et al., 1997; Hrubcova´et al., 2005; Ru˚žek et al., 2007). All results suggest thatthe deep structure <strong>of</strong> the Bohemian Massif crust, whichwas consolidated during the Variscan orogeny, reflectstectonic processes related to Palaeozoic subduction<strong>and</strong> collision at the subcrustal lithosphere level as well(Babusˇka et al., 2010).According to this model, the crust is characterizedby a succession <strong>of</strong> positive <strong>and</strong> negative anomalies <strong>of</strong>60–80 km wavelength for a nearly constant Mohodepths. The central part <strong>of</strong> the Bohemian Massif displaysa large negative Bouguer anomaly correspondingto the Palaeozoic crustal root represented by theMoldanubian domain (Fig. 3). The adjacent NeoproterozoicBrunia microcontinent displays an importantgravity high caused by mafic <strong>and</strong> intermediate medium-grademetamorphic <strong>and</strong> magmatic rocks. However,the strong gradient marking the deep crustalÓ 2010 Blackwell Publishing Ltd183


82 O. LEXA ET AL.(a)(a´)(b)(b´)(a)(b´)(b)(a´)Fig. 2. Geological maps <strong>of</strong> Blansky´ les <strong>and</strong> St. Leonhard granulite areas with schematic <strong>structural</strong> pr<strong>of</strong>iles. Two P–T space insetsshow the metamorphic evolution <strong>of</strong> lower crustal <strong>and</strong> middle crustal rocks (Petrakakis, 1997; Pitra et al., 1999; Scheuvens, 2002;Sˇtı´pska´ & Powell, 2005b; Racek et al., 2006, 2008). Protolith ages (Kro¨ ner et al., 2000; Friedl et al., 2004; Verner et al., 2008) <strong>of</strong>major rock types are shown on the map.boundary between the root domain <strong>and</strong> the Bruniamicrocontinent is shifted 50–70 km westwards relativeto their contact on the surface suggesting that the highdensity basement rocks are covered by a thin sheet <strong>of</strong>light granulites <strong>and</strong> migmatites in this area (Schulmannet al., 2008). North-west <strong>of</strong> the Moldanubian domainthere is an important gravity high corresponding to theNeoproterozoic basement <strong>of</strong> the Tepla´-Barr<strong>and</strong>ianUnit. This is limited to the north by southeast dippingreflectors <strong>of</strong> the Tepla´ suture, which is characterized byhigh density eclogites <strong>and</strong> ultramafic rocks. The footwall<strong>of</strong> the suture corresponds to low density felsiccrust <strong>of</strong> the Saxothuringian basement.The seismic reflection <strong>and</strong> refraction sections <strong>and</strong>gravity <strong>modelling</strong> suggest a complex lithologicalstructure <strong>of</strong> the Moldanubian domain marked by a lowdensity, 5–10 km thick lower crustal layer locatedabove the Moho, a 5–10 km thick dense mafic layer, a10-km thick mid-crustal layer <strong>of</strong> intermediate density<strong>and</strong> a locally developed 2–5 km thick low density layerat the top (Fig. 3). The low density lower crust correlateswell with low-P velocities in the range 6.0–6.4 km s )1 in the CEL09 section (Ru˚žek et al., 2007).Guy et al. (2010) proposed that the low density layerlocated above the Moho corresponds to felsic rocks,which are interpreted as deep crustal equivalents <strong>of</strong>surface outcrops <strong>of</strong> the Gfo¨ hl Unit. These authorsinterpreted the high density thick layer located abovelight granulites as an equivalent <strong>of</strong> the AGB (Fig. 3).The intermediate density layer forming recent uppercrust <strong>of</strong> the Moldanubian domain is interpreted asMonotonous <strong>and</strong> Varied group rocks, whereas the lowdensity rocks on the surface are directly correlated withexposures <strong>of</strong> the Gfo¨ hl Unit. It is suggested that thislayered structure <strong>of</strong> the Variscan crust reflects that <strong>of</strong>the original thick root, which was thinned by lateVariscan <strong>and</strong> Permian extensional processes (Burget al., 1994).Exhumation model connecting deep crustal geophysics <strong>and</strong>surface geologyGuy et al. (2010) <strong>and</strong> Franěk et al. (2011a) proposed amodel connecting the deep structure <strong>of</strong> the Variscancrust with surface distribution <strong>of</strong> lower <strong>and</strong> mid-crustalrocks. The layered structure <strong>of</strong> the orogenic rootreported by geophysics thus represents a relict <strong>of</strong>Carboniferous distribution <strong>of</strong> horizontally layeredcrust prior to exhumation. In addition, it is supposedthat the observed vertically layered distribution <strong>of</strong>orogenic lower crust surrounded by middle crustalunits reflects steepening <strong>of</strong> the deep crustal horizontallayering. The model connecting deep crustal layeringwith surface geology is based on several recent studiessuggesting that the granulites were exhumed alongsteep channels from lower crustal depth by a ductileextrusion mechanism (Sˇtı´pska´ et al., 2004; Schulmannet al., 2005; Franeˇk et al., 2006, 2011a; Tajcˇmanova´et al., 2006).In this view, the granulites represented, beforeexhumation, the <strong>structural</strong>ly deepest orogenic lowercrust located at a depth <strong>of</strong> 60–70 km (18–20 kbar,Ó 2010 Blackwell Publishing Ltd184


HEAT SOURCES AND EXHUMATION MECHANISMS 83ABABFig. 3. Bouguer anomaly map <strong>of</strong> the Bohemian Massif (combined data provided by the Czech Geological Survey <strong>and</strong> Guy et al.(2010)) <strong>and</strong> gravimetric model for section A–B (pr<strong>of</strong>ile no. 4 in Guy et al., 2010). The main lithological unit boundaries are representedby superimposed thick black lines. The section A–B emphasizes the present density structure <strong>of</strong> the Moldanubian Domain with relicts<strong>of</strong> felsic rocks in the lower crust.800–900 °C; Sˇtı´pska´ & Powell, 2005b; Tajcˇmanova´et al., 2006). Structurally above were mafic rocks <strong>of</strong> theAGB, which formed during early Palaeozoic magmaticunderplating <strong>of</strong> the Proterozoic crust <strong>of</strong> the MonotonousGroup (Schulmann et al., 2005, figs 16 & 17), <strong>and</strong>the <strong>structural</strong>ly highest Varied Group corresponding tosupracrustal early Palaeozoic sequences (Schulmannet al., 2009). According to this hypothesis, the extrusionÓ 2010 Blackwell Publishing Ltd185


84 O. LEXA ET AL.mechanism lead to inversion <strong>of</strong> this crustal pr<strong>of</strong>ile sothat the granulites presently form cores <strong>of</strong> crustalanticlines, surrounded by AGB, variably thinned duringextrusion, <strong>and</strong> by the Gfo¨ hl gneiss followed bymetamorphic sequences <strong>of</strong> the Monotonous <strong>and</strong> Variedgroups. In many places, the whole structure is eveninverted because <strong>of</strong> the lateral spreading <strong>of</strong> verticallyextruded lower crustal material (Fig. 2). This model iscorroborated by detailed <strong>structural</strong> <strong>and</strong> petrologicaldata, which show that the granulites <strong>and</strong> surroundingrocks exhibit common vertical fabrics which are associatedwith HP–high temperature (HT) mineralassemblages. The flat fabric, which locally may have abowl shape, is superimposed on vertical extrusionfabrics <strong>and</strong> developed during subsurface spreading <strong>of</strong>the extruded partially molten lower crust (7–4 kbar,700–650 °C; Racek et al., 2006; Tajcˇmanova´ et al.,2006; Hasalova´ et al., 2008b).CHRONOLOGY OF THE VARISCANCALC-ALKALINE TO POTASSIC MAGMATISM –CONSTRAINTS ON TIME-SCALES OF THEOCEANIC AND CONTINENTAL SUBDUCTIONThe contact <strong>of</strong> the Tepla´-Barr<strong>and</strong>ian <strong>and</strong> Moldanubianunits is marked by voluminous granitic plutons <strong>of</strong>Variscan age with (normal or potassic) calc-alkalinechemistry <strong>and</strong> large ion lithophile element (LILE) ⁄ -high field strength element (HFSE) enrichmentresembling magmatic associations <strong>of</strong> active continentalmargins. The oldest vestige <strong>of</strong> this igneous activityinitiated by a late Devonian–early CarboniferousAndean-type subduction is preserved in the Cˇista´ <strong>and</strong>Sˇteˇnovice plutons (Venera et al., 2000; Zˇa´k et al.,2010) as well as orthogneisses in the ro<strong>of</strong> <strong>of</strong> the largeCentral Bohemian Plutonic Complex, CBPC (protolithc. 370–360 Ma; Kosˇler et al., 1993). A newlyidentified member <strong>of</strong> the subduction-related associationis the mafic Lisˇov low-pressure granulite unit insouthern Bohemia, the protolith <strong>of</strong> which was emplacedat c. 360 Ma into middle crustal levels (15–20 km) <strong>of</strong> the same igneous arc (Janousˇek et al.,2006).After a significant time gap, subduction-relatedmembers <strong>of</strong> the CBPC were emplaced, including thenormal calc-alkaline gabbros, quartz diorites <strong>and</strong> tonalites<strong>of</strong> the Sa´zava suite (354.1 ± 3.5 Ma; Janousˇeket al., 2004a) <strong>and</strong> the high-K calc-alkaline, mainlygranodioritic Blatna´ suite (347 + 4 ⁄ )3 Ma; Do¨ rr &Zulauf, 2010; 346.4 ± 1.1 Ma; Janousˇek et al., 2010)with associated monzonitic bodies. Finally, furthereast, commonly in association with HP granulite massifs<strong>and</strong> high-grade Gfo¨ hl orthogneisses, syn-deformationalor post-tectonic intrusions <strong>of</strong> earlyCarboniferous (343–336 Ma) (ultra-) potassic rockswere emplaced (Holub, 1997; Holub et al., 1997;Janousˇek et al., 2003; Verner et al., 2008; Kotkova´et al., 2010; Kusiak et al., 2010). The time lag betweenthe end <strong>of</strong> normal calc-alkaline magmatism (c. 354 Ma)<strong>and</strong> the onset <strong>of</strong> intrusion <strong>of</strong> K-rich magmas (c.346 Ma) can be related to the transition from oceanplate subduction to continental underthrusting.PETROLOGY AND GEOCHEMISTRY OFMOLDANUBIAN GRANULITES AND POTASSICMAGMATIC ROCKSA peculiar feature <strong>of</strong> the Moldanubian domain in theBohemian Massif is an intimate spatial <strong>and</strong> temporalassociation between felsic HP, kyanite–garnet granulites<strong>and</strong> large ultrapotassic plutonic bodies (Janousˇek& Holub, 2007). Whereas the granulites tend to bedepleted in the radioactive elements U, Th <strong>and</strong> K, the(ultra-) potassic rocks are characterized by strongenrichment in these elements that shows clearly in theradiometric map (Fig. 4). Thus, underst<strong>and</strong>ing <strong>of</strong> theK, U <strong>and</strong> Th depletion ⁄ enrichment in individual rocktypes at various stages <strong>of</strong> the Viséan HP metamorphism<strong>and</strong> igneous activity seems to be a key to decipheringthe thermal history <strong>of</strong> the Moldanubianorogenic crust.Moldanubian HP granulitesHigh-pressure granulites represent a voluminous <strong>and</strong>ubiquitous component <strong>of</strong> the Gfo¨ hl Assemblage inboth Austria <strong>and</strong> the Czech Republic. The most typicalare felsic types consisting essentially <strong>of</strong> garnet, quartz<strong>and</strong> hypersolvus feldspar, <strong>and</strong> commonly containingkyanite. Rutile, zircon, apatite, ilmenite ± monaziteare the common accessories (OÕBrien & Ro¨ tzler, 2003).The felsic Moldanubian granulites were considered asformer rhyolites ⁄ granites that acquired their highgrademetamorphic character during the Variscancollision <strong>and</strong> which suffered only limited HP melting(Fiala et al., 1987a; Vellmer, 1992; Janousˇek et al.,2004b). Indeed, some granulites in rare domains thatescaped later mylonitization ⁄ recrystallization aremigmatitic in appearance (Franeˇk et al., 2006). On theother h<strong>and</strong>, other authors (Vra´na & Jakesˇ, 1982; Jakesˇ,1997; Kotkova´ & Harley, 1999) suggested that thefelsic granulites represent HP granitic liquids thatformed <strong>and</strong> separated from their source during theVariscan metamorphic cycle. There seem to be severalarguments against such a model (Janousˇek et al.,2004b): (i) concentrations <strong>of</strong> Zr are far too lowcompared to calculated saturation levels at ‡900 °C,coupled with significant, mainly Ordovician zirconinheritance; (ii) consistently low (750 °C) zircon <strong>and</strong>monazite saturation temperatures, whereas pre-Variscaninheritance is by no means rare, documenting thatsaturation was reached (Janousˇek, 2006); (iii) preservation<strong>of</strong> Ordovician–Silurian whole rock <strong>and</strong> thin slabRb–Sr ages corresponding to the principal inheritedcomponent in granulite zircon; <strong>and</strong> (iv) high heavy rareearth element + Y contents, ruling out the presence <strong>of</strong>large amounts <strong>of</strong> garnet in the residue <strong>and</strong> thus indicatinga rather low-P melting.Ó 2010 Blackwell Publishing Ltd186


HEAT SOURCES AND EXHUMATION MECHANISMS 85Fig. 4. Radiometric map <strong>of</strong> the south-eastern Bohemian Massif with isolines <strong>of</strong> natural air absorbed dose rate (nGy h )1 ). Themain bodies <strong>of</strong> ultrapotassic rocks (durbachite series <strong>and</strong> melasyenitoids sensu Holub, 1997) that show very high radioactivity areidentified by name. The Variscan granites (crosses) <strong>and</strong> HP granulites (dots) are outlined. Source: Czech Geological Survey MapServer, http://www.geology.cz.The most peculiar feature <strong>of</strong> the felsic Moldanubiangranulites is the lack <strong>of</strong> LILE depletion, except for Cs,U <strong>and</strong> Th (Fiala et al., 1987a,b; Janousˇek et al.,2004b). Thus, unlike in many other granulite terranesworldwide (Rudnick & Presper, 1990; Rudnick & Gao,2003), prograde metamorphism appears to have beenlargely isochemical. As shown by Janousˇek et al.(2004b), the composition <strong>of</strong> felsic Moldanubian granulitesmatches well with felsic Ordovician–Silurianmetaigneous rocks from the Saxothuringian domain,for instance orthogneisses <strong>and</strong> meta-rhyolites from theFichtelgebirge. The similarities include whole-rockgeochemistry (excluding the most mobile elements Cs,Rb, Th, U, Pb <strong>and</strong> Li – Fig. 5a), Sr–Nd isotopiccompositions <strong>and</strong> protolith ages <strong>of</strong> c. 480–455 Ma(Siebel et al., 1997; Wieg<strong>and</strong>, 1997), forming animportant maximum within the spectrum <strong>of</strong> inheritedages in the granulites.A good c<strong>and</strong>idate for complementary small-volume,HP–HT (>900 °C) melt that managed to separatefrom the calc-alkaline granulites are the rare hyperpotassicgranulites from Plesˇovice, Blansky´ les Massif(Vra´na, 1989; Janousˇek et al., 2007). These rocks,which are composed essentially <strong>of</strong> K-feldspar, garnet,zircon <strong>and</strong> apatite, show rather extreme geochemicalcompositions (e.g. strong enrichments in Cs, Ba, Rb,U, Th, K, P <strong>and</strong> Zr; Fig. 5b). They yielded U–Pb zirconages <strong>of</strong> 338 ± 1 Ma (Aftalion et al., 1989) <strong>and</strong>337.13 ± 0.37 Ma (Sla´ma et al., 2008), which is closeto the established best estimate <strong>of</strong> the HP metamorphicclimax in the granulite massifs (c. 340 Ma – see above).Moreover, rare coeval glimmerite veins in peridotitefragments enclosed by granulite bodies show highconcentrations <strong>of</strong> LILE, U <strong>and</strong> Th, accompanied bylow Rb ⁄ Cs <strong>and</strong> K ⁄ Rb ratios as well as low HFSEcontents (Fig. 5c); the Sr <strong>and</strong> Nd isotopic compositionsoverlap with the granulites (Becker et al., 1999).The glimmerites were interpreted as having crystallizedfrom an ultrapotassic, F-rich aqueous-carbonic fluid,bearing a direct witness for the HP–HT devolatilization<strong>of</strong> granulite massifs.Viséan (ultra-) potassic magmatism in the Moldanubi<strong>and</strong>omainIn the Moldanubian domain <strong>of</strong> the Bohemian Massif,relatively large volumes <strong>of</strong> (ultra-) potassic plutonicrocks constitute several plutons <strong>and</strong> stocks spatiallyassociated to granitoids <strong>of</strong> the Central Bohemian PlutonicComplex, the Moldanubian Plutonic Complex,Ó 2010 Blackwell Publishing Ltd187


86 O. LEXA ET AL.(a)(b)(c)<strong>and</strong> high-grade metamorphic rocks <strong>of</strong> the Gfo¨ hlAssemblage, most notably felsic HP granulites. Thisprominent group represents the late-syntectonic durbachiteseries (e.g. Cˇertovo brˇemeno, Trˇebı´cˇ <strong>and</strong> Knı´žecı´stolec intrusions – Zˇa´k et al., 2005; Verner et al., 2008)(Fig. 4) <strong>of</strong> quartz melasyenites to melagranites withhydrous ferromagnesian minerals, Mg-rich biotite <strong>and</strong>actinolitic amphibole. These are mainly coarsely porphyriticrocks that contain abundant K-feldspar phenocrysts<strong>and</strong> ultrapotassic mafic microgranularenclaves. The spatially associated, less deformed, oreven post-tectonic, biotite–two-pyroxene melasyenitesto melagranites (Ta´bor <strong>and</strong> Jihlava intrusions) arecharacterized by a ÔdrierÕ ferromagnesian mineralassemblage <strong>of</strong> orthopyroxene, clinopyroxene <strong>and</strong>Mg-biotite <strong>and</strong> lack the porphyritic texture (Ža´k et al.,2005).The petrogenesis <strong>of</strong> the Moldanubian (ultra-)potassic igneous rocks has been a matter <strong>of</strong> debate aseven the basic, Mg <strong>and</strong> K-rich members or primitivelamprophyric dykes have a mixed geochemical character.While their high contents <strong>of</strong> Cr <strong>and</strong> Ni with highmg# point to derivation from an olivine-rich source(mantle peridotite), the elevated concentrations <strong>of</strong> U,Th, light rare earth element (LREE) <strong>and</strong> LILE,depletion in Ti, Nb <strong>and</strong> Ta (Fig. 5d) <strong>and</strong> highK 2 O ⁄ Na 2 O <strong>and</strong> Rb ⁄ Sr ratios apparently contradict amantle origin (Holub, 1997; Janousˇek & Holub, 2007).These features led several authors to invoke partialmelting <strong>of</strong> anomalous (LILE- <strong>and</strong> LREE-enriched)lithospheric mantle domains (e.g. Janousˇek et al.,1995; Holub, 1997; Wenzel et al., 1997, 2000; Janousˇek& Holub, 2007), followed by mixing with lower crustalleucogranitic melts (Holub, 1997; Gerdes et al., 2000).In any case, the crustal-like Sr–Nd isotopic signaturescannot be reconciled solely by crustal assimilation⁄ contamination during the ascent <strong>of</strong> any primitive,mantle-derived magmas <strong>and</strong> require contamination bythe subducted continental crust directly in the source(d)Fig. 5. Multi-element variation diagrams. (a) Box <strong>and</strong> percentileplots for felsic (SiO 2 >70 wt%) Moldanubian granulites,normalized to an average <strong>of</strong> felsic metaigneous rocks from theFichtelgebirge (Siebel et al., 1997; Wieg<strong>and</strong>, 1997). See Janousˇeket al. (2004b <strong>and</strong> references therein) for the data set. The distribution<strong>of</strong> each <strong>of</strong> the normalized trace-element contents isplotted as irregular polygons, the width <strong>of</strong> which at any givenheight is proportional to the empirical cumulative distribution.As in box plots, the median, 25th <strong>and</strong> 75th percentiles aremarked with horizontal lines across the box. Compositions <strong>of</strong>the hyperpotassic Plesˇovice granulites (data from Janousˇeket al., 2007) (b) <strong>and</strong> Lower Austrian glimmerite veins (Beckeret al., 1999) (c) normalized by an average <strong>of</strong> felsic (SiO 2>70 wt%) HP Moldanubian granulites, except Lisˇov (takenfrom Janousˇek et al., 2004b). (d) Multi-element variationdiagram for selected ultrapotassic rocks (durbachite series <strong>and</strong>syenitoids) from the Moldanubian Zone <strong>of</strong> the BohemianMassif (Janousˇek & Holub, 2007) normalized by N-MORB <strong>and</strong>then adjusted to Yb N = 1 to minimize the effects <strong>of</strong> fractionalcrystallization (Pearce & Stern, 2006).Ó 2010 Blackwell Publishing Ltd188


HEAT SOURCES AND EXHUMATION MECHANISMS 87(Janousˇek & Holub, 2007). As noted by the sameauthors, multi-element plots for the Moldanubian felsicgranulites (Fig. 5a) <strong>and</strong> ultrapotassic rocks(Fig. 5d) are largely mutually complementary. Themost striking are the cases <strong>of</strong> Cs, Rb, Th, U, Pb <strong>and</strong> Li,which are impoverished in the felsic granulites butstrongly enriched in the ultrapotassic magmatic rocks.NUMERICAL MODELLING CONSTRAINTS ONHEAT SOURCES AND EMPLACEMENTMECHANISMS OF THE OROGENIC LOWERCRUSTTo constrain the heat source driving the tectonic processes<strong>of</strong> vertical extrusion three contrasting, but notmutually exclusive, hypotheses will be tested. Heatcould be generated by: (i) in situ decay <strong>of</strong> radioactiveelements contained in the FLC (U, Th <strong>and</strong> K); (ii)radioactive elements present in the metasomatized orcrustally contaminated mantle; <strong>and</strong> (iii) a large-scalethermal anomaly generated in the mantle as a result <strong>of</strong>slab break <strong>of</strong>f or mantle delamination.Fig. 6. Model geometry, initial lithology distribution, boundaryconditions <strong>and</strong> location <strong>of</strong> tracked samples used for <strong>numerical</strong>simulations.The annual thermal productions were calculatedfollowing the method <strong>of</strong> Kramers et al. (2001), usingdecay constants <strong>and</strong> specific heat production datasummarized by van Schmus (1995; table 8). The pastannual heat production (lW kg )1 ) can be obtainedfrom the elemental concentrations <strong>of</strong> K, U <strong>and</strong> Thusing the equation:HðlW kg 1 623:45 10 2:638 10Þ¼Ke 0:554t þ The0:0495t324:03777 10 9:396852 10þ Ue 0:985t þe 0:1551t ;ð1Þwhere t represents age in Ga <strong>and</strong> K, U, Th are concentrationsin ppm.Model setupThe <strong>numerical</strong> model studies outlined below describe thetransient thermal evolution <strong>of</strong> a thickened orogenicdomain (Fig. 6) characterized by the presence <strong>of</strong> tectonicallyaccreted felsic rocks, including granulites,within orogenic lower crust. This FLC directly underliesa mafic layer which was added to Neoproterozoic crustduring early Palaeozoic crustal stretching <strong>and</strong> magmaticunder-plating. The presence <strong>of</strong> low density FLC below adense mafic layer introduces significant gravitationalinstability within the lower crust (Gerya et al., 2001)which could trigger crustal diapirism (Ramberg, 1981;Perchuk, 1989) <strong>and</strong> perturbate the thermal field. We usethermal <strong>and</strong> dynamic <strong>numerical</strong> models to examine therole <strong>of</strong> high radioactive heat production located in theFLC as a main c<strong>and</strong>idate triggering the gravitationalinstability as a result <strong>of</strong> pronounced progressive generation<strong>of</strong> heat <strong>and</strong> subsequent change in density because<strong>of</strong> thermal expansion.The model is set up to allow the definition <strong>of</strong> differentmaterial domains with different thermal <strong>and</strong> mechanicalproperties (Table 2) on high-resolution Lagrangianmarkers initially arranged in a rectangular grid (Gerya& Yuen, 2003). Properties are mapped to a Eulerianstaggered grid where governing equations are solved fortemperature change (DT) <strong>and</strong> velocity. In each time,step markers are advected according to the updatedvelocity field <strong>and</strong> all temperature-dependent variablesTable 1. Calculated radioactive heat production values for Fichtelgebirge metaigneous rocks, felsic Moldanubian granulites <strong>and</strong>Moldanubian peridotite (present <strong>and</strong> at 340 Ma).Rock type Age (Ma) Density (kg m )3 ) Concentrations (ppm) A (lW m )3 )K Th U40 K232 Th235 U238 U Present PastFichtelgebirge 340 2700 38642.550 13.00 9.00 7.3421 13.00 0.064 8.935 3.670 3.920Moldanubian granulites 340 2750 38601.045 2.10 1.00 7.3342 2.10 0.007 0.993 0.788 0.885Hornı´ Bory peridotite 340 3200 940.537 0.11 0.57 0.1787 0.11 0.004 0.568 0.176 0.214Data sources for averaged whole-rock compositions: Fichtebeirge metaigneous rocks: Siebel et al. (1997), Wieg<strong>and</strong> (1997).Felsic Moldanubian granulites (SiO 2 > 70 wt%): Janousˇek et al. (2004b <strong>and</strong> references therein).Hornı´ Bory peridotite: Ackerman et al. (2009).Ó 2010 Blackwell Publishing Ltd189


88 O. LEXA ET AL.Fig. 7. Results <strong>of</strong> 1D static thermal models to compare temperature evolution controlled by different radioactive heat production <strong>of</strong>felsic lower crust (FLC; curves labelled 1–6 according to heat production) <strong>and</strong> radioactive heat production within the lithosphericmantle (dotted line) with thermal evolution caused by lithospheric delamination at 90 km depth (dashed line).such as thermal diffusivity, specific heat, viscosity <strong>and</strong>density are recalculated according to temperaturedependence for each marker (see Appendix).The geometry <strong>of</strong> the model, the distribution <strong>of</strong>material layers <strong>and</strong> the boundary conditions <strong>of</strong> the<strong>numerical</strong> simulations are depicted in Fig. 6. The initiallayered geometry introduces an artificial perturbationin the interface between felsic <strong>and</strong> mafic lowercrust to allow immediate relaxation <strong>of</strong> gravitationalinstability in the central part <strong>of</strong> the computationaldomain. The initial temperature distribution is calculatedas the steady-state solution <strong>of</strong> Eqn A.3, withheat sources located only in the upper crust. Thelower crustal radioactive heat production is accountedonly for transient development. One <strong>of</strong> theimportant features <strong>of</strong> the model is the ability toprogressively eliminate sources <strong>of</strong> radioactive heatproduction according to the temperature achieved. Asargued above using the available geochemical data, ata certain stage <strong>of</strong> the evolution <strong>of</strong> the lower crust, most<strong>of</strong> the radioactive elements, in particular U <strong>and</strong> Th,were stripped from the felsic granulites into partialmelt or ÔfluidÕ <strong>and</strong> transported, together with K-richmagmas, into the middle–upper crust. In the models,two values are used to cut-<strong>of</strong>f heat production tosimulate radioactive element evacuation via fluid or viamelt.Heat sourcesThe effect <strong>of</strong> such behaviour on thermal evolution wasfirst examined in terms <strong>of</strong> a 1D static thermal model(Fig. 7). A sharp change in heating rate in Fig. 7ccorresponds to the time when heat production isswitched <strong>of</strong>f in most <strong>of</strong> the lower crust. To compare thescale <strong>and</strong> magnitude <strong>of</strong> temperature change as a result<strong>of</strong> processes like delamination or heat productionwithin lithospheric mantle, the plot is overlaid withresults <strong>of</strong> two additional <strong>numerical</strong> simulations. Thedotted line represents the results <strong>of</strong> a 1D model tosimulate production <strong>of</strong> heat within a >60 km thicklithospheric mantle with a radioactive heat production<strong>of</strong> 0.2 lW m )3 calculated on the basis <strong>of</strong> the wholerockgeochemical data <strong>of</strong> Ackerman et al. (2009) forthe Hornı´ Bory garnet peridotite (Table 1). It is evidentthat even such an enriched mantle cannot providesufficient heat to be responsible for the significantincrease <strong>of</strong> temperature within the lower crust (Fig. 7c).Similarly, we argue that the process <strong>of</strong> lithospheredelamination (simulated by instantaneous replacement<strong>of</strong> lithospheric mantle below 90 km by asthenospherein the model) cannot provide the necessary heat input.Results related to the mantle delamination process areshown by the dashed line in Fig. 7.A series <strong>of</strong> <strong>numerical</strong> experiments was set up to studythe influence <strong>of</strong> radioactive heat production locatedwithin the FLC (Fig. 7). Our calculations show thatthe temperature required for partial melting <strong>of</strong> micabearingfelsic crust located at a depth <strong>of</strong> 70 km (850–900 °C) are reached after 20 Myr for radioactive heatproduction <strong>of</strong> 2 lW m )3 <strong>and</strong> in 7 Myr for radioactiveheat production <strong>of</strong> 4 lW m )3 . At 60 km depth, whichis the assumed upper limit <strong>of</strong> the felsic layer, themelting temperature is reached in >50 Myr <strong>and</strong> inÓ 2010 Blackwell Publishing Ltd190


HEAT SOURCES AND EXHUMATION MECHANISMS 89Table 2. Mechanical properties (density, thermal expansivity <strong>and</strong> coefficients <strong>of</strong> temperature-dependent viscosity) <strong>of</strong> individuallithologies used for <strong>numerical</strong> simulations.Material Description Reference density Thermal expansion coefficient Temperature-dependent viscosity range <strong>and</strong>coefficientsRadioactive heat productionq0 (kg m )3 ) a Effect. viscosity (Pa s )1 ) C1 C2 Hr (lW m )3 )UC Upper crust 2700 0 10 22 Viscosity const. 2 · 10 )6MCs Middle crust (schists) 2800 2 · 10 )5 1.5 · 10 20 to 2.5 · 10 19 10 16 6000 0MCg Middle crust (gneisses) 2800 2 · 10 )5 2.5 · 10 20 to 3.5 · 10 19 10 17 6000 0MLC Mafic lower crust 2950 0 2 · 10 21 to 5 · 10 20 10 18 7000 0FLC Felsic lower crust 2750 3 · 10 )5 1.5 · 10 19 to 6 · 10 18 10 17 5000 2 · 10 )6 to 8 · 10 )6M Lithospheric mantle 3300 0 5 · 10 21 Viscosity const. 0 (2 · 10 )7 )c. 20 Myr for the two radioactive heat productionvalues, respectively. Keeping in mind the time-scales<strong>of</strong> magmatic <strong>and</strong> metamorphic events related toPalaeotethys subduction discussed above, the time <strong>of</strong>5–15 Myr available between the arrival <strong>of</strong> continentalcrust into the subduction zone (354–346 Ma) <strong>and</strong> themetamorphic climax (c. 340 Ma) corresponds to thethermal incubation time estimated for radiogenic heatproduction <strong>of</strong> 4 lW m )3 .Results <strong>of</strong> 2D <strong>modelling</strong>The distribution <strong>of</strong> densities <strong>and</strong> viscosities for allmodels generally resulted in the development <strong>of</strong> adiapiric structure located at the introduced perturbation.The P–T <strong>of</strong> selected samples (samples 1–3 arewithin lower crust, sample 4 in mafic layer <strong>and</strong> sample5 in middle crust; for locations, see Fig. 6), trackedduring the model evolution are plotted on Figs 8–10.Similar to the static 1D models, the simulations show aclear relation between the initial temperature increase(20–150 °C) <strong>and</strong> radioactive heat production withinlower crust.Two types <strong>of</strong> evolution have been calculated: (i) adiapiric structure formed because <strong>of</strong> mantle heatsource (heat production 0.2 lW m )3 ); <strong>and</strong> (ii) a diapiricstructure caused by radioactive heat productionwithin the FLC for radioactive heat productivitiesranging from 1 to 6 lW m )3 . The results <strong>of</strong> the<strong>numerical</strong> simulations for the case <strong>of</strong> mantle heatproduction are shown in Fig. 8. For the case <strong>of</strong>radioactive heat production within the FLC, it wasterminated at 900 <strong>and</strong> 1000 °C. These results are presentedtogether with a simulation in which the radioactiveheat production was not switched <strong>of</strong>f (Fig. 9)allowing a direct comparison <strong>of</strong> differences in P–Tevolution (Fig. 10).Several major conclusions can be drawn from theseresults. The diapir reflecting the mantle heat source(Fig. 8) exhibits a typical bell shape during the first30 Myr <strong>and</strong> most importantly shows contrasting P–Tevolution for samples located in different units <strong>and</strong>initial depths. Whereas the upper part <strong>of</strong> the mantle(Fig. 7d) is almost isobarically heated, the sampleslocated in the felsic crust, mafic lower crustal layers<strong>and</strong> the middle crust reveal relatively slow exhumation<strong>and</strong> moderate heating associated with the diapirgrowth.In contrast, experiments assuming high radiogenicheat production show typical bollard type diapirs.After 10–20 Myr imposed gravitational instability <strong>and</strong>variable radioactive heat production within lowercrust, the viscosity <strong>of</strong> the overlying mafic layer is significantlyreduced allowing relatively fast exchange <strong>of</strong>material <strong>and</strong> development <strong>of</strong> the central diapir. Thereare differences in rates <strong>of</strong> vertical exchange betweenindividual simulations as increased heat productioncause increase <strong>of</strong> buoyancy forces <strong>and</strong> decrease <strong>of</strong>viscosity <strong>of</strong> the diapir surroundings (samples 4 <strong>and</strong> 5).The growth <strong>of</strong> bollard type diapirs is associated witheither isothermal decompression or important coolinglinked to diapir growth for samples located in deepfelsic lower crustal layer. These are indeed the P–Tevolutions retrieved from for Bohemian Massif granulites(Sˇtı´pska´ et al., 2004; Racek et al., 2006;Tajcˇmanova´ et al., 2006). The samples located originallyhigher in the column <strong>and</strong> at the middle crustallevels reveal important heating associated with exhumation,which is also in accord with recent petrologicalstudies (Racek et al., 2006; Sˇtı´pska´ et al., 2008). Theother important consequence is that rocks fromany original position show convergence <strong>of</strong> P–T conditionsafter exhumation to mid-crustal depths. Thetime-scales <strong>of</strong> heating (10–20 Myr) <strong>and</strong> exhumation(5–10 Myr) calculated in this model are also in agreementwith petrological <strong>and</strong> geochronological datareported by Sˇtı´pska´ et al. (2004) <strong>and</strong> Tajcˇmanova´ et al.(2006, 2010). It should be noted that time-scales <strong>of</strong> ourmodels are directly controlled by the rheology <strong>of</strong> thematerials, which in our simulations is significantlysimplified (Eqn A.3, Table 2).Closer inspection <strong>of</strong> Fig. 10 confirms that the best fit<strong>of</strong> modern petrological <strong>and</strong> geochronological data withcalculated P–T paths is with an initial radioactive heatproduction <strong>of</strong> 4 lW m )3 . Here, the maximum temperaturesattained at the bottom <strong>of</strong> thickened crust are950 °C, while samples located in the central part <strong>of</strong>the felsic crustal column reach a maximum 800 °C atthe thermal peak. These values may reconcile modernTHERMOCALC<strong>modelling</strong> data <strong>of</strong> Sˇtı´pska´ & Powell(2005b), Tajcˇmanova´ et al. (2006) <strong>and</strong> Racek et al.(2008) who reported peak temperatures 800 °C withÓ 2010 Blackwell Publishing Ltd191


90 O. LEXA ET AL.Fig. 8. Results <strong>of</strong> <strong>numerical</strong> simulation <strong>of</strong> radioactive heat production within the lithospheric mantle showing distribution <strong>of</strong>lithologies <strong>and</strong> temperature field developed during 50 Myr. Light shade <strong>of</strong> the mantle colour marks the asthenosphere (adiabaticgeotherm <strong>and</strong> no heat production).those <strong>of</strong> OÕBrien (2000) <strong>and</strong> Cooke & OÕBrien (2001)who reached significantly higher peak temperatureestimates. Higher heat production would produce significantlyhigher peak temperatures <strong>of</strong> nearly 1000 °Cmaintained even at pressures as low as 12–13 kbar,which is in contradiction with modern petrologicalstudies (see Schulmann et al., 2008 for review).Other important information comes from the temperaturedistribution in the core <strong>of</strong> the diapiric structure.The diagrams calculated for elevated radioactiveheat production show areas where the temperaturecondition <strong>of</strong> radioactive heat production switch <strong>of</strong>f(i.e. partial melting associated with release <strong>of</strong> radioactiveelements) was attained during the evolution (thepinkish colour inside the diapirs in Fig. 9).As shown above, the metaigneous rocks fromFichtelgebirge are thought to be the best c<strong>and</strong>idates forprecursors <strong>of</strong> the Moldanubian felsic granulites.Therefore, these rocks are interpreted as material thatcould have formed the felsic lower crustal layer, whichwas subsequently heated, partially melted <strong>and</strong>transformed to typical felsic Moldanubian granulites.Ó 2010 Blackwell Publishing Ltd192


HEAT SOURCES AND EXHUMATION MECHANISMS 91Fig. 9. Results <strong>of</strong> <strong>numerical</strong> simulations showing distribution <strong>of</strong> lithologies <strong>and</strong> temperature field developed after 20 Myr. Light shade<strong>of</strong> the lower crustal colour marks places where removal <strong>of</strong> radioactive heat production occurred while light shade <strong>of</strong> the mantle colourmarks asthenosphere (adiabatic geotherm <strong>and</strong> no heat production). First column shows results for low radioactive heat production <strong>and</strong>lithology <strong>and</strong> temperature fields are identical for all switch-<strong>of</strong>f conditions as they are not reached during 20 Myr. The second columngives results for no switch-<strong>of</strong>f condition, whereas the third <strong>and</strong> fourth columns shows results <strong>of</strong> simulations with 900 <strong>and</strong> 1000 °Cswitch <strong>of</strong>f for the lower crustal layer.The metaigneous rocks from Fichtelgebirge yield anaverage radioactive heat production <strong>of</strong> 3.9 lW m )3obtained from the average elemental concentrations <strong>of</strong>K, U <strong>and</strong> Th (Siebel et al., 1997; Wieg<strong>and</strong>, 1997) usingEqn 1 (Table 1). However, the radioactive heat productionfor the average felsic granulites (SiO 2 > 70,Table 1) is extremely low (0.9 lW m )3 ), suggestingthat the radioactive elements were mostly lost duringÓ 2010 Blackwell Publishing Ltd193


92 O. LEXA ET AL.Fig. 10. P–T evolution <strong>of</strong> six tracked samples for models with different radioactive heat production values for the felsic lowercrust (FLC). Samples 1, 2 <strong>and</strong> 3 were located in felsic lower crust, sample 4 within the mafic layer (MLC) <strong>and</strong> samples 5 <strong>and</strong> 6 arelocated in the middle crust. (a) No radioactive elements removed, (b) 900 °C <strong>and</strong> (c) 1000 °C threshold for radioactive elementremoval from the lower crust. Circles mark 10 Myr time step.the partial melting connected with the granulite faciesmetamorphism.In conclusion, the thermal structure which fits bestthe petrological <strong>and</strong> geochronological data obtained s<strong>of</strong>ar from the Moldanubian granulite massifs is thatcalculated with an initial radioactive heat production<strong>of</strong> 4 lW m )3 located in the felsic layer. The P–T–tevolutions <strong>of</strong> the AGB (mafic lower crust) <strong>and</strong> <strong>of</strong> theoverlying mid-crustal rocks are also in good agreementwith the model results, despite the fact that the simulatedevolution represents only the initial part <strong>of</strong> acomplex polyphase exhumation.DISCUSSIONIn this article, we discuss a particular <strong>structural</strong> patternin the Variscan orogenic root in which orogenic lowercrust composed <strong>of</strong> felsic granulites <strong>of</strong> OrdovicianÓ 2010 Blackwell Publishing Ltd194


HEAT SOURCES AND EXHUMATION MECHANISMS 93protolith age forming cores <strong>of</strong> domes that are separatedfrom mid-crustal Neoproterozoic <strong>and</strong> Palaeozoicmetasedimentary rocks in synclines by a late Ordovician–Silurianmetabasic layer. We argue that the origin<strong>of</strong> these structures was related to diapiric materialexchange within the orogenic lower crust. The keyelement in deciphering the Viséan development is theoccurrence <strong>of</strong> FLC underneath dense mafic crust asdepicted by geology <strong>and</strong> geophysics. To underst<strong>and</strong>the formation <strong>of</strong> this peculiar lithological s<strong>and</strong>wich thepossible emplacement models for the FLC at the bottom<strong>of</strong> the root need to be discussed. Subsequently, weshall address the particular role <strong>of</strong> this felsic layer forthe origin <strong>of</strong> the felsic granulite–(ultra-) potassicmagma association. Finally, a model <strong>of</strong> gravity inversionis discussed together with thermal consequencesfor P–T–t paths <strong>of</strong> rocks in different positions withrespect to the diapiric structure.Model <strong>of</strong> relamination <strong>of</strong> felsic crust to early Palaeozoicmafic lower crustThe relamination <strong>of</strong> FLC has been proposed as analternative to the model <strong>of</strong> Chemenda et al. (2000) toexplain the structure <strong>of</strong> the Tibetan Plateau. TheTibetan Plateau is characterized by an exceptionallylarge gravity low indicating dominantly a felsic rootunderplated by Indian felsic crust, the density <strong>of</strong> whichcorresponds to felsic granulite at a pressure <strong>of</strong> 20 kbar(Hetényi et al., 2007). Indeed, Le Pichon et al. (1997)argued that the high topography <strong>of</strong> the Tibetan Plateauis due to presence <strong>of</strong> low density granulites atdepth. A similar gravity low is typical <strong>of</strong> the AltiplanoPlateau in central Andes, having been interpreted as aresult <strong>of</strong> underthrusting <strong>of</strong> the Brazilian crust underneaththe Andean root (Oncken et al., 2006). Thegravity anomaly associated with the Moldanubi<strong>and</strong>omain resembles remarkably the Tibetan <strong>and</strong> Altiplanoplateaux density structure, which in this case isreduced by subsequent isostatic reequilibration (Burget al., 1994). The common denominator <strong>of</strong> all thesegeophysical observations is the occurrence <strong>of</strong> felsiccrust at lower crustal depths. This is explained either ascontinental crust underthrusting thin lithosphericmantle (Chemenda et al., 2000) or directly by influx <strong>of</strong>felsic crust into the orogenic root at Moho depth liftingthe original lower crust <strong>and</strong> depressing the mantlelithosphere (Behr, 1978; Plesch & Oncken, 1999;Avouac, 2008).In the western Bohemian Massif, the influx <strong>of</strong> ductilelower crust at granulite ⁄ eclogite facies conditions wasproposed by OÕBrien (2000). In Fig. 11a,b, the influx <strong>of</strong>Saxothuringian crust into the root domain is visualizedin the form <strong>of</strong> a 10-km wide channel splitting the earlyPalaeozoic mafic lower crust from lithospheric mantle,whereas the other part <strong>of</strong> the continental crust iscontinuously subducted, contaminating the localmantle <strong>and</strong> sampling the mantle lithosphere. The termrelamination (Hacker et al., 2007) is accepted as beingsuitable for the addition <strong>of</strong> low density crust underneaththe dense root, in contrast to the term delaminationto describe loss <strong>of</strong> heavy root material <strong>and</strong> itsreplacement by the asthenosphere.As discussed before, oceanic subduction had to haveceased by 354–346 Ma to be replaced by continentalunderthrusting. Incidentally, a Sm–Nd age <strong>of</strong>354 ± 6 Ma was determined by dating <strong>of</strong> calcium-richcores <strong>of</strong> garnet from the South Bohemian granulitesindicating the onset <strong>of</strong> eclogitization <strong>of</strong> continentalcrust at this time (Prince et al., 2000). If true, such ascenario allows a 5–15 Myr period to c. 340 Ma, thegranulite facies metamorphic climax, for the subductedcontinental crust to thermally incubate <strong>and</strong> elevate theorogenic geotherm (Engl<strong>and</strong> & Thompson, 1984).Based on P–T estimates, crustal thickening had toproduce a 70-km thick crust at this time (Fig. 11c).Melting <strong>of</strong> this continental crust started at c. 345 Ma,as indicated by high-K calc-alkaline magmatism <strong>of</strong> theBlatna´ suite. The maximum melt production at boththe base <strong>of</strong> this crust <strong>and</strong> in the underlying mantlelithosphere occurred during, or soon after, the HPmetamorphic climax at c. 340 Ma, as indicated byintrusions <strong>of</strong> ultrapotassic syenites at mid- to highcrustallevels.The origin <strong>of</strong> the felsic granulite–ultrapotassic plutonic rockassociation <strong>and</strong> the heat sourceRoberts & Finger (1997) proposed that heating <strong>of</strong>relatively refractory felsic metaigneous rocks, the likelysource <strong>of</strong> the felsic granulites, to temperatures as highas 1000 °C would result in production <strong>of</strong> 5–15 vol.%<strong>of</strong> partial melt. This notion was confirmed by thermodynamic<strong>modelling</strong> <strong>of</strong> Janousˇek et al. (2004b), whosuggested that the initial melting, limited by micaavailability, would not exceed 10 vol.% for any <strong>of</strong> theprospective granulite decompression paths <strong>and</strong> rapidincrease <strong>of</strong> the melt fraction would occur only at1100 °C. However, it is important to correctly assessthe melt production for the most typical felsic granulites.The most likely protolith for the felsic granulitesis considered to be granite or orthogneiss <strong>of</strong> theÔFichtelgebirge chemistryÕ that must have lost somemelt during metamorphic evolution in order to preservethe high-pressure assemblage. Therefore, toestimate the degree <strong>of</strong> melt loss, pseudosections werecalculated using THERMOCALC for a granulite with theoldest known preserved fabric from the Blansky´ lesMassif (sample H296-S1 from Franeˇk et al., 2011b; fordetails, see Appendix).The sample modelled contains relict porphyroclasts<strong>of</strong> perthitic feldspar with inclusions <strong>of</strong> kyanite, garnetwith 24 mol.% <strong>of</strong> grossular, rutile <strong>and</strong> biotite, whichconstrains the peak P–T conditions to 16–18 kbar<strong>and</strong> 860 °C (Fig. 12a; Franeˇk et al., 2011b). It isnecessary to add 7 mol.% <strong>of</strong> melt into the rockcomposition to obtain assemblages containing biotite<strong>and</strong> muscovite without garnet or alumosilicate atÓ 2010 Blackwell Publishing Ltd195


94 O. LEXA ET AL.(a)(b)(c)(d)Fig. 11. Model proposed for the tectonic evolution <strong>of</strong> the orogenic root domain in the Bohemian Massif. (a) Model <strong>of</strong> continentalunderthrusting with development <strong>of</strong> the Barr<strong>and</strong>ian forearc region, CBPC magmatic arc <strong>and</strong> backarc region represents the futureMoldanubian domain. The position <strong>of</strong> the Palaeotethys suture is indicated as the future Maria´nské La´zneˇ Complex. (b) Relaminationmodel with part <strong>of</strong> the allochthonous Saxothuringian crust injected between the Moho <strong>and</strong> the continental lithosphere. The other partis subducted thereby producing metasomatism <strong>of</strong> the overlying mantle. (c) Crustal thickening <strong>of</strong> the former backarc domain(Schulmann et al., 2005, 2009). (d) Vertical extrusion <strong>and</strong> gravity redistribution <strong>of</strong> relaminated Saxothuringian crust at the end <strong>of</strong>Variscan orogeny.Ó 2010 Blackwell Publishing Ltd196


HEAT SOURCES AND EXHUMATION MECHANISMS 95(a) (b)Fig. 12. (a) Pseudosection for granulite sample H296-S1 (Blanský les Massif) to show the P–T path from the peak assemblage <strong>of</strong> g–ky–pl–ksp–q–liq–ru at 16 kbar <strong>and</strong> 860 °C bydecompression to 8 kbar <strong>and</strong> 820 °C (modified from Franeˇk et al., 2011b). Molar percentage <strong>of</strong> melt is very low <strong>and</strong> allows the preservation <strong>of</strong> the peak assemblage ondecompression <strong>and</strong> cooling. (b) Pseudosection for sample H296-S1 with 7 mol.% <strong>of</strong> melt added allows the stability <strong>of</strong> the mu–bi–pl–ksp–q assemblage at subsolidus conditions.Such a reconstructed granite whole-rock <strong>and</strong> mineral composition predicts 8 mol.% <strong>of</strong> melt at peak P–T conditions <strong>and</strong> 12 mol.% <strong>of</strong> melt at 8 kbar <strong>and</strong> 860 °C. Some melt mustbe lost to preserve the peak g–ky–pl–ksp–q–liq–ru assemblage.Ó 2010 Blackwell Publishing Ltd197


96 O. LEXA ET AL.subsolidus conditions. Such a reconstructed granitemineral composition calculated at 650 °C <strong>and</strong> 9 kbar is8 mol.% muscovite, 6.5 mol.% biotite, 39 mol.%quartz, 30 mol.% plagioclase <strong>and</strong> 16 mol.% K-feldspar(Fig. 12b). On heating, the melt proportion forthe reconstructed granite at 16 kbar <strong>and</strong> 860 °C is8 mol.%, <strong>and</strong> on isothermal decompression to 8 kbar<strong>and</strong> 860 °C it increases to 13 mol.%. The calculationspredict 14 mol.% <strong>of</strong> melt at 16 kbar <strong>and</strong> 1000 °C thatincreases to 37 mol.% at 8 kbar <strong>and</strong> 1000 °C, confirmingthe earlier estimates by Roberts & Finger(1997) <strong>and</strong> Janousˇek et al.(2004b).The <strong>modelling</strong> demonstrates that heating <strong>of</strong> graniticcrust with a restricted amount <strong>of</strong> muscovite <strong>and</strong> biotitecan generate a garnet-bearing residue <strong>and</strong> up to7 mol.% <strong>of</strong> melt at temperatures 900 °C. This temperaturewas used as a threshold for melt loss relatedremoval <strong>of</strong> radioactive elements <strong>and</strong> the switching <strong>of</strong>f<strong>of</strong> the heat production in lower crustal rocks in the<strong>numerical</strong> experiments.The detailed micro<strong>structural</strong> studies <strong>of</strong> Franěk et al.(2006, 2011b) showed that the origin <strong>of</strong> the granulitemicrostructure is related to significant deformation,dynamic recrystallization <strong>and</strong> viscous flow during thevertical extrusion stage. Before the vertical extrusionstage, the granulite facies rocks acquired a metamorphicfabric in a rather static environment. The deformationoccurred via diffusion-assisted grain boundarysliding, which is an efficient mechanism to extract meltfrom a continuously deforming source by a combination<strong>of</strong> dynamic dilation <strong>and</strong> compaction (Za´vadaet al., 2007). The mechanism <strong>of</strong> dynamically developedporosity <strong>and</strong> melt extraction was discussed by Hasalova´et al. (2008a), who showed that a small amount <strong>of</strong>melt may be efficiently extracted from the source farbelow the rheologically critical melt percentagethreshold (>20–25 vol.% melt; Vigneresse et al.,1996). Therefore, we consider 900 °C to be a reasonabletemperature when melt can be extracted from theparent rock via the grain boundary sliding mechanismleaving behind a garnet-bearing mylonitic rock – theMoldanubian granulite. However, this processrequired significant deformation related to the verticalextrusion event to extract the melt. It is the pervasiveearly vertical fabric which is related to the efficient loss<strong>of</strong> melt (even at very small melt fraction), resulting insignificant depletion <strong>of</strong> the parental rocks in U, Th, Cs,Li ± Rb, but leaving the rest <strong>of</strong> the geochemical signatureunaffected.A consequence <strong>of</strong> melting <strong>and</strong> melt extraction is theremoval <strong>of</strong> a significant part <strong>of</strong> the radioactive elementbudget from the system. This is particularly true for U<strong>and</strong> Th hosted by accessories such as zircon or monazite,which would resist fluid loss in course <strong>of</strong> theprogressive heating. On the other h<strong>and</strong>, they shoulddissolve readily even at low degrees <strong>of</strong> melting, as thepartial melt was likely to be rather corrosive, being hot<strong>and</strong> rich in alkalis with fluorine (Finger & Cooke,2004; Janousˇek et al., 2007). In granitic rocks, therelevant accessories are mostly enclosed in biotite (Bea,1996) that was undergoing melting. Moreover, as arguedby Watson et al. (1989) on theoretical grounds,the larger accessories are likely to be progressivelyconcentrated at grain boundaries in the course <strong>of</strong> highgrademetamorphism. Finally, there is a directevidence for the presence <strong>of</strong> low-degree, high-T, traceelement-richmelt in the felsic HP granulites, as thenewly grown metamorphic zircon <strong>and</strong> rutile are rich inU, Th <strong>and</strong> LREE or Zr <strong>and</strong> Nb, respectively (Finger &Cooke, 2004). Thus, the accessories hosting U <strong>and</strong> Thin the protolith to the HP Moldanubian granulitesseem to have been largely accessible to, <strong>and</strong> dissolvedin, the low-degree HP melt.The 1D thermal <strong>modelling</strong> has demonstrated that,for radioactive heat production <strong>of</strong> 4 lW m )3 , thetemperature threshold <strong>of</strong> 900 °C would be reached indeeply buried crust after 7 Myr (at 70 km) to 15 Myr(at 60 km). A radioactive heat production <strong>of</strong>4 lW m )3 is similar to the average radioactive heatproduction at c. 340 Ma calculated for the Fichtelgebirgemetaigneous crust, which is considered tocorrespond to the most appropriate protolith <strong>of</strong> felsicgranulites (Janousˇek & Holub, 2007). Moreover, the<strong>modelling</strong> shows that heat necessary for crustal meltingindeed could have been produced internally, within thetime frame allowed by the available geochronologicaldata (5–15 Myr).The heat source located at lower crustal depthswould also lead eventually to partial melting <strong>of</strong> theunderlying metasomatized <strong>and</strong> hydrated subcrustalmantle lithosphere. Melting <strong>of</strong> such an anomalous <strong>and</strong>fertile mantle source could have produced, soon afterthe HP metamorphic peak, ultrapotassic rocks withmixed crustal–mantle signatures. The effective removal<strong>of</strong> U <strong>and</strong> Th from the partially molten felsic metaigneousrocks (future felsic granulites) by melt extractionwould have grave consequences for the thermal evolution<strong>of</strong> the whole system. First, the depletion <strong>of</strong>radioactive elements from the granulite means that themelting process must have rapidly switched <strong>of</strong>f. Second,the extracted melts could have partly mixed withenriched mantle-derived ultrapotassic magmas, whichinvaded the overlying partially molten crust, contributingto their further enrichment by U <strong>and</strong> Th.However, this mixing was probably volumetricallyrather insignificant <strong>and</strong> essentially different fromanother, large-scale hybridization event with S-typeleucogranitic magmas assumed for the durbachiteseries in Bohemia (Holub, 1997) <strong>and</strong> the Rastenbergsuite in Lower Austria (Gerdes et al., 2000). Thehybrid magmas finally intruded syn-tectonically orpost-tectonically at mid- to high-crustal levels in closespatial <strong>and</strong> temporal association with the HP granulite<strong>and</strong> orthogneiss complexes so typical <strong>of</strong> the Variscanorogenic crust in the Moldanubian domain <strong>of</strong> theBohemian Massif (Schulmann et al., 2005; Zˇa´k et al.,2005; Tajcˇmanova´ et al., 2006; Janousˇek & Holub,2007).Ó 2010 Blackwell Publishing Ltd198


HEAT SOURCES AND EXHUMATION MECHANISMS 97Gravity overturns <strong>and</strong> 340 Ma crustal redistribution in theBohemian MassifGerya et al. (2001, 2002) in their pioneering work proposeda model <strong>of</strong> gravity redistribution <strong>of</strong> granodioriticrocks <strong>and</strong> overlying mafic crust using radioactive heatproduction localized in granodioritic crust. This progressiveheating <strong>of</strong> the FLC leads to its viscosity drop<strong>and</strong> density decrease triggering diapiric process. However,the time-scales required for gravity overturnsproposed by these authors are not compatible with thosereported for the Variscan orogeny (discussed before; seealso Schulmann et al., 2009). Therefore, the temperatureincrease must have been significantly more radical<strong>and</strong> density contrasts higher than those predicted byGerya et al. (2002, 2004). This is consistent with thegravity inversion <strong>modelling</strong> results <strong>of</strong> Guy et al. (2010)who have shown that the density contrast betweenresidual FLC <strong>and</strong> an overlying thick gabbroic layer(Fig. 3) may be greater than that estimated by Geryaet al. (2001). Moreover, an additional source <strong>of</strong> gravitypotential may be seen in the presence <strong>of</strong> eclogites in thethickened Ordovician–Silurian AGB layer (50–60 km),overlying the felsic crust at a depth <strong>of</strong> 60–70 km (Sˇtípska´& Powell, 2005a,b).The scenario presented here (Fig. 11d) resembles theexplanation <strong>of</strong> the formation <strong>of</strong> migmatitic domes in hotorogenic belts driven by gravitational collapse <strong>of</strong>thickened continental crust (Rey et al., 2001; V<strong>and</strong>erhaeghe& Teyssier, 2001; V<strong>and</strong>erhaeghe, 2009). All theseconceptual models suggest partial melting <strong>of</strong> the lowercrust to be a trigger mechanism for development <strong>of</strong>gravitational instability in both fossil <strong>and</strong> modern orogenicbelts. The main difference <strong>of</strong> our model is inquantification <strong>of</strong> the gravitational instability, the rate <strong>of</strong>the heating <strong>and</strong> also the rate <strong>of</strong> the development <strong>of</strong> thediapiric structures. The initial position <strong>of</strong> less dense buthighly radioactive felsic material below more densemafic lower crust seems to be the necessary prerequisitefor driving the tectonic evolution <strong>of</strong> the Variscan orogenyin the Bohemian Massif. Importantly, in our<strong>numerical</strong> models, the development <strong>of</strong> granulite domesis caused solely by gravitational instability. However,the <strong>structural</strong> data suggest that domes are initiated byfolding <strong>of</strong> the lower <strong>and</strong> mid-crustal interfaces (e.g.Sˇtípska´ et al., 2004; Schulmann et al., 2005) <strong>and</strong> thelinear distribution <strong>of</strong> elongate granulite belts indicatesthat lateral shortening was an important component(Burg et al., 2004) related to the growth <strong>of</strong> granulitedomes (Fig. 11d). Therefore, the combination <strong>of</strong> gravity<strong>and</strong> laterally forced extrusion <strong>of</strong> orogenic crust may leadto a development <strong>of</strong> gravity overturns more rapidlycompared with the <strong>numerical</strong> models presented herein.Our work shows that only a model <strong>of</strong> internalheating to drive the tectonic <strong>and</strong> gravity redistribution<strong>of</strong> orogenic lower crust can satisfactorily explain thetemporarily restricted orogenic event at c. 340 Ma,which is responsible for most <strong>of</strong> the Variscan tectonothermalevents reported so far in the BohemianMassif. Thus, the timing <strong>of</strong> growth <strong>of</strong> these laterallyforced diapirs seems to be connected with orogeniccollapse <strong>and</strong> the major plate reorganization <strong>of</strong> thewhole Variscan belt (Edel et al., 2003). The co-existence<strong>of</strong> c. 340 Ma ultrapotassic plutons <strong>and</strong> extrudedfelsic HP granulites thus defines a key thermomechanicalevent, which probably signified a rheologicalcollapse <strong>of</strong> the whole Variscan belt in Europe (e.g.Rossi et al., 2009; Rubatto et al., 2010).CONCLUSIONSThe exhumation <strong>of</strong> c. 340 Ma felsic granulites in theBohemian Massif is interpreted in terms <strong>of</strong> tectonicallytriggered gravity redistribution <strong>of</strong> felsic orogenic lowercrust <strong>and</strong> high density mafic crust. The model showsthat radioactive heat production <strong>of</strong> 4 lW m )3 forlower crustal rocks, which is corroborated by calculatedvalues from likely protolith rocks, <strong>and</strong> the calculatedP–T–t evolution satisfy the thermal <strong>and</strong>geochronological evolution <strong>of</strong> the Bohemian Massifgranulites. This radioactive heat production is typical<strong>of</strong> Ordovician felsic igneous rocks in the Fichtelgebirge(Saxothuringian domain), which are believed to havebeen relaminated at the bottom <strong>of</strong> thickened continentalcrust during the early Vise´an continentalunderthrusting. The mutually complementary geochemicalcharacteristics <strong>of</strong> granulites <strong>and</strong> ultrapotassicmagmas highlight the shared thermomechanical historygoverned by radioactive heating <strong>of</strong> the lowercrustal layer, culminating in partial melting <strong>of</strong> both thefelsic lower crustal layer itself <strong>and</strong> its underlying fertile(metasomatized ⁄ contaminated) mantle lithosphere.Gravity-driven redistribution tectonics initiated byinternal heating is interpreted to be the principal agentcontrolling the rheological collapse <strong>of</strong> the Variscanorogenic crust at c. 340 Ma.ACKNOWLEDGEMENTSWe acknowledge grant MSM0021620855 from theMinistry <strong>of</strong> Education <strong>of</strong> the Czech Republic <strong>and</strong>internal research funds from CNRS UMR 7615 forsalary <strong>and</strong> research support <strong>of</strong> Ondrej Lexa, <strong>and</strong> theFrench National Science Foundation ANR projectÔLFO in orogensÕ for additional research support. Wethank C. Clark <strong>and</strong> T. Gerya for their thoroughreviews <strong>and</strong> highly appreciate the comments <strong>and</strong>suggestions, which significantly contributed toimproving the quality <strong>of</strong> the publication. M. Brown isacknowledged for careful editorial work.APPENDIXGoverning equations used for <strong>numerical</strong> <strong>modelling</strong> <strong>of</strong>gravity overturnsTo simulate the thermal evolution, we use a <strong>numerical</strong>model based on the solution <strong>of</strong> the coupled equationsÓ 2010 Blackwell Publishing Ltd199


98 O. LEXA ET AL.<strong>of</strong> momentum (Eqn A.1) <strong>and</strong> energy (Eqn A.3), subjectto an incompressibility constraint (Eqn A.2) known asthe Boussinesq approximation (Hansen & Yuen, 2000):@r ij¼ @P qg i ; ðA:1Þ@x i @x i@v i@x i¼ 0;@T@t þ v @Ti ¼ 1 @ @TjqC p þ A ;@x i qC p @x i @x i qC pðA:2ÞðA:3Þwhere x denotes the coordinates in m, v velocity inms )1 , t time in s; r ij is stress tensor, P is pressure in Pa,g is gravitational acceleration (9.81 m s )2 ), T is temperaturein K, j denotes thermal diffusivity <strong>and</strong> C pdenotes the specific heat capacity. The constitutiverelationship between stress <strong>and</strong> strain is governed bythe transport coefficient g representing viscosity:r ij ¼ 2g_e ij ;ðA:4Þwhere _e ij is the strain-rate tensor in s )1 . For the purpose<strong>of</strong> this work, we used a simplified flow law withonly temperature-dependent viscosity according to thesimple exponential equation:g ¼ C 1 e C 2ð T Þ ;ðA:5Þwhere C 1 <strong>and</strong> C 2 are coefficients used to calculateeffective viscosity from prescribed range (see Table 2).Density q is given by equation <strong>of</strong> state:q ¼ q 0 ½1 aðT T 0 ÞŠ; ðA:6Þwhere a is the coefficient <strong>of</strong> thermal expansion <strong>and</strong> q 0is the reference density at reference temperature T 0 .Thermal diffusivity j <strong>and</strong> specific heat capacity C p arerecalculated according to temperature using the followingequations (Whittington et al., 2009):j ¼ 3:19 10 7 þ 1:214 10 6 eð 273:15285:2T Þ ; ðA:7Þ3:224 10 5 2:714 107C p ¼ 1538:39þTT 2 ; ðA:8Þderived from laser-flash analysis to provide realisticvalues for geologically relevant temperatures. Theseequations are solved for temperature <strong>and</strong> velocity <strong>and</strong>describe the fundamental physics required for <strong>modelling</strong>the thermal evolution during crustal diapirism.Calculation methods used for thermodynamic <strong>modelling</strong>The pseudosections were calculated using THERMOCALC3.30 (Powell et al., 1998) <strong>and</strong> the data set 5.5 (Holl<strong>and</strong>& Powell, 1998; November 2003 upgrade), in the systemNa 2 O–CaO–K 2 O–FeO–MgO–Al 2 O 3 –SiO 2 –H 2 O–TiO 2 –O (NCKFMASHTO) with the biotite <strong>and</strong> meltmodels from White et al. (2007), garnet from Dieneret al. (2008), ilmenite from White et al. (2000), feldsparfrom Holl<strong>and</strong> & Powell (2003), white mica fromCoggon & Holl<strong>and</strong> (2002) <strong>and</strong> cordierite fromTHERMOCALC documentation (Powell & Holl<strong>and</strong>, 2004).The analysed rock composition <strong>of</strong> sample H296 (inwt% SiO 2 =71.98, TiO 2 =0.42, Al 2 O 3 =13.53,FeO=2.1, MnO=0.03, MgO=0.73, CaO=1.93,Na 2 O=2.76, K 2 O=4.01, P 2 O 5 =0.15, H 2 O ) =0.22,H 2 O + =0.56, CO 2 =0.03), was modified for <strong>modelling</strong>by adding 1 mol.% <strong>of</strong> kyanite to enable a smallamount <strong>of</strong> aluminosilicate to be stable at the estimatedpeak metamorphic conditions, as is observed in thinsection.The reconstruction <strong>of</strong> a biotite–muscovite graniteprotolith requires several steps. It involves determination<strong>of</strong> the H 2 O content in the final assemblage, consideration<strong>of</strong> open-system behaviour with respect tomelt, <strong>and</strong> modification <strong>of</strong> the whole-rock compositionby adding melt (White & Powell, 2002; White et al.,2004; Sˇtı´pska´ et al., 2008). Tracking <strong>of</strong> the P–T path istherefore undertaken backwards in time, from thematrix assemblage to the early prograde evolution <strong>and</strong>to the protolith mineralogical composition. Theamount <strong>of</strong> H 2 O for the <strong>modelling</strong> shown in Fig. 12a isset such that it allows the stability <strong>of</strong> the observedmatrix assemblage with garnet rim chemistry oncooling (not shown; see Franeˇk et al., 2011b; seeHasalova´ et al., 2008b for the approach followed). Asthe whole-rock composition is expected to changealong the P–T path as a result <strong>of</strong> loss <strong>of</strong> melt, it has tobe decided from which point on the P–T path the meltcomposition will be taken. Crossing the upper stability<strong>of</strong> muscovite causes an abrupt increase in melt proportion<strong>and</strong> is considered a likely condition for meltloss (White & Powell, 2002; White et al., 2004);therefore, melt composition reintegrated was undertakenat 16 kbar <strong>and</strong> 860 °C.REFERENCESAckerman, L., Jelínek, E., Medaris, G., Jezˇek, J., Siebel, W. &Strnad, L., 2009. Geochemistry <strong>of</strong> Fe-rich peridotites <strong>and</strong>associated pyroxenites from Horní Bory, Bohemian Massif:insights into subduction-related melt-rock reactions. ChemicalGeology, 259, 152–167.Aftalion, M., Bowes, D. & Vrána, S., 1989. Early CarboniferousU-Pb zircon age for garnetiferous, perpotassic granulites,Blansky´ les massif, Czechoslovakia. Neues Jahrbuch fu¨rMineralogie-Monatshefte, 4, 145–152.Avouac, J.P., 2008. Dynamic processes in extensional <strong>and</strong>compressional settings – mountain building: from earthquakesto geological deformation. In: Treatise on Geophysics, Volumes1-11 (ed. Schubert, G.), pp. 377–439. Elsevier, Amsterdam.Babusˇka, V., Fiala, J. & Plomerova´, J., 2010. Bottom to top lithospherestructure <strong>and</strong> evolution <strong>of</strong> western Eger Rift (CentralEurope). International Journal <strong>of</strong> Earth Sciences, 99, 891–907.Bea, F., 1996. Residence <strong>of</strong> REE, Y, Th <strong>and</strong> U in granites <strong>and</strong>crustal protoliths; implications for the chemistry <strong>of</strong> crustalmelts. Journal <strong>of</strong> Petrology, 37, 521–552.Ó 2010 Blackwell Publishing Ltd200


HEAT SOURCES AND EXHUMATION MECHANISMS 99Becker, H., Wenzel, T. & Volker, F., 1999. Geochemistry <strong>of</strong>glimmerite veins in peridotites from Lower Austria – implicationsfor the origin <strong>of</strong> K-rich magmas in collision zones.Journal <strong>of</strong> Petrology, 40, 315–338.Behr, H., 1978. Subfluenz-Prozesse im Grundgebirgs-StockwerkMitteleuropas. Zeitschrift der Deutschen Gesellschaft fu¨r Geowissenschaften,129, 283–318.Burg, J.-P., Van Den Driessche, J. & Brun, J.-P., 1994. Syn-topost-thickening extension in the Variscan Belt <strong>of</strong> WesternEurope: modes <strong>and</strong> <strong>structural</strong> consequences. Ge´ologie de laFrance, 3, 33–51.Burg, J.-P., Kaus, B. & Podladchikov, Y., 2004. Dome structuresin collision orogens. Mechanical investigation <strong>of</strong> the gravity⁄ compression interplay. In: Gneiss Domes in Orogeny, Vol.380 (eds Whitney, D.L., Teyssier, C. & Siddoway, C.S.), pp.47–66, Geological Society <strong>of</strong> America Special Papers, Boulder,Colorado.Chemenda, A., Burg, J.-P. & Mattauer, M., 2000. Evolutionarymodel <strong>of</strong> the Himalaya–Tibet system: geopoem based on new<strong>modelling</strong>, geological <strong>and</strong> geophysical data. Earth <strong>and</strong> PlanetaryScience Letters, 174, 397–409.Coggon, R. & Holl<strong>and</strong>, T.J.B., 2002. Mixing properties <strong>of</strong>phengitic micas <strong>and</strong> revised garnet–phengite thermobarometers.Journal <strong>of</strong> Metamorphic Geology, 20, 683–696.Cooke, R. & OÕBrien, P.J., 2001. Resolving the relationshipbetween high P–T rocks <strong>and</strong> gneisses in collisional terranes: anexample from the Gfo¨ hl gneiss–granulite association in theMoldanubian Zone, Austria. Lithos, 58, 33–54.Dewey, J.F., Ryan, P.D. & Andersen, T.B., 1993. Orogenic uplift<strong>and</strong> collapse, crustal thickness, fabrics <strong>and</strong> metamorphic phasechanges: the role <strong>of</strong> eclogites. In: Magmatic Processes <strong>and</strong>Plate Tectonics, Vol. 76 (eds Prichard, H.M., Alabaster, T.,Harris, N.B.W. & Neary, C.R.), pp. 325–343, GeologicalSociety, Special Publications, London.Diener, J.F.A., Powell, R. & White, R.W., 2008. <strong>Quantitative</strong>phase petrology <strong>of</strong> cordierite–orthoamphibole gneisses <strong>and</strong>related rocks. Journal <strong>of</strong> Metamorphic Geology, 26, 795–814.Do¨ rr, W. & Zulauf, G., 2010. Elevator tectonics <strong>and</strong> orogeniccollapse <strong>of</strong> a Tibetan-style plateau in the European Variscides:the role <strong>of</strong> the Bohemian shear zone. International Journal <strong>of</strong>Earth Sciences, 99, 299–325.Edel, J.B., Schulmann, K. & Holub, F.V., 2003. Anticlockwise<strong>and</strong> clockwise rotations <strong>of</strong> the Eastern Variscides accommodatedby dextral lithospheric wrenching: palaeomagnetic <strong>and</strong><strong>structural</strong> evidence. Journal <strong>of</strong> the Geological Society, 160,209–218.Engl<strong>and</strong>, P.C. & Thompson, A.B., 1984. Pressure temperaturetime paths <strong>of</strong> regional metamorphism. 1. Heat-transfer duringthe evolution <strong>of</strong> regions <strong>of</strong> thickened continental crust. Journal<strong>of</strong> Petrology, 25, 894–928.Fiala, J., Mateˇjovska´, O. & Vanˇkova´, V., 1987a. Moldanubiangranulites <strong>and</strong> related rocks: petrology, geochemistry <strong>and</strong>radioactivity. Rozpravy Cˇeskoslovenske´ Akademie Veˇd, RˇadaMatematicky´ch a Prˇıŕodnıćh Veˇd, 97, 1–102.Fiala, J., Mateˇjovska´, O. & Vanˇkova´, V., 1987b. Moldanubiangranulites: source material <strong>and</strong> petrogenetic considerations.Neues Jahrbuch fu¨r Mineralogie, Abh<strong>and</strong>lungen, 157, 133–165.Finger, F. & Cooke, R., 2004. Evidence for the presence <strong>of</strong> atrace-element-loaded interstitial partial melt in a Moldanubianleucocratic granulite derived from LA-ICP-MS <strong>analyses</strong> onzircons <strong>and</strong> rutiles. In: International Workshop on Petrogenesis<strong>of</strong> Granulites <strong>and</strong> Related Rocks, Na´meˇsˇtˇ nad Oslavou, October1-3, 2004. Excursion Guide & Abstract Volume (ed. Janousˇek,V.), pp. 35–36. Moravian Museum, Brno.Finger, F. & von Quadt, A., 1995. U ⁄ Pb ages <strong>of</strong> zircons from aplagiogranite-gneiss in the south-eastern Bohemian Massif,Austria-further evidence for an important early Paleozoicrifting episode in the eastern Variscides. Schweizerische Mineralogischeund Petrographische Mitteilungen, 75, 265–270.Finger, F., Gerdes, A., Janousˇek, V., Rene´, M. & Riegler, G.,2007. Resolving the Variscan evolution <strong>of</strong> the Moldanubiansector <strong>of</strong> the Bohemian Massif: the significance <strong>of</strong> theBavarian <strong>and</strong> the Moravo–Moldanubian tectonometamorphicphases. Journal <strong>of</strong> Geosciences, 52, 9–28.Franeˇk, J., Schulmann, K. & Lexa, O., 2006. Kinematic <strong>and</strong>rheological model <strong>of</strong> exhumation <strong>of</strong> high pressure granulites inthe Variscan orogenic root: example <strong>of</strong> the Blanský les granulite,Bohemian Massif, Czech Republic. Mineralogy <strong>and</strong>Petrology, 86, 253–276.Franeˇk, J., Schulmann, K., Lexa, O., Tomek, Cˇ. & Edel, J.-B.,2011a. Model <strong>of</strong> syn-convergent extrusion <strong>of</strong> orogenic lowercrust in the core <strong>of</strong> the Variscan belt: implications for exhumation<strong>of</strong> high-pressure rocks in large hot orogens. Journal <strong>of</strong>Metamorphic Geology, 29, 53–78.Franeˇk, J., Schulmann, K., Lexa, O. et al., 2011b. The precursor<strong>of</strong> Variscan felsic granulites, process <strong>of</strong> its granulitization, <strong>and</strong>inversed rheology <strong>of</strong> feldspars <strong>and</strong> quartz in the lower-crustalconditions. Journal <strong>of</strong> Metamorphic Geology, 29, 103–130.Franke, W., 2000. The mid-European segment <strong>of</strong> the Variscides:tectonostratigraphic units, terrane boundaries <strong>and</strong> plate tectonicevolution. Geological Society, London, Special Publications,179, 35–61.Friedl, G., Finger, F., Paquette, J.-L., von Quadt, A.,McNaughton, N.J. & Fletcher, I.R., 2004. Pre-Variscan geologicalevents in the Austrian part <strong>of</strong> the Bohemian Massifdeduced from U–Pb zircon ages. International Journal <strong>of</strong> EarthSciences, 93, 802–823.Fuchs, G., 1976. Zur Entwicklung der Bo¨ hmischen Masse.Jahrbuch der Geologischen Bundesanstalt, 129, 41–49.Gerdes, A., Worner, G. & Finger, F., 2000. Hybrids, magmamixing <strong>and</strong> enriched mantle melts in post-collisional Variscangranitoids: the Rastenberg Pluton, Austria. In: OrogenicProcesses: Quantification <strong>and</strong> Modelling in the Variscan FoldBelt, Vol. 179 (eds Franke, W., Haak, V., Oncken, O. &Tanner, D.), pp. 415–431, Geological Society, Special Publications,London.Gerya, T. & Yuen, D.A., 2003. Characteristics-based marker-incellmethod with conservative finite-differences schemes formodeling geological flows with strongly variable transportproperties. Physics <strong>of</strong> the Earth <strong>and</strong> Planetary Interiors, 140,293–318.Gerya, T., Maresch, W., Willner, A., Van Reenen, D. & Smit,C., 2001. Inherent gravitational instability <strong>of</strong> thickened continentalcrust with regionally developed low-to mediumpressuregranulite facies metamorphism. Earth <strong>and</strong> PlanetaryScience Letters, 190, 221–235.Gerya, T., Perchuk, L., Maresch, W., Willner, A., Van Reenen,D. & Smit, C., 2002. Thermal regime <strong>and</strong> gravitationalinstability <strong>of</strong> multi-layered continental crust: implications forthe buoyant exhumation <strong>of</strong> high-grade metamorphic rocks.European Journal <strong>of</strong> Mineralogy, 14, 687–699.Gerya, T.V., Perchuk, L.L., Maresch, W.V. & Willner, A.P., 2004.Inherent gravitational instability <strong>of</strong> hot continental crust:implications for doming <strong>and</strong> diapirism in granulite facies terrains.In: Gneiss Domes in Orogeny, Vol. 380 (eds Whitney, D.L.,Teyssier, C. & Siddoway, C.S.), pp. 97–115, Geological Society<strong>of</strong> America, Special Papers, Boulder, Colorado.Guy, A., Edel, J.-B., Schulmann, K., Tomek, Cˇ. & Lexa, O.,2010. A geophysical model <strong>of</strong> the Variscan orogenic root(Bohemian Massif): Implications for modern collisional orogens.Lithos, in press. doi:10.1016/j.lithos.2010.08.008.Hacker, B.R., Kelemen, P.B. & Behn, M.D., 2007. Continentalrelamination drives compositional <strong>and</strong> physical-propertychanges in the lower crust, Eos Trans. AGU, 88, Fall MeetingSuppl., Abstract V32A-06.Hansen, U. & Yuen, D.A., 2000. Extended-Boussinesq thermalchemicalconvection with moving heat sources <strong>and</strong> variableviscosity. Earth <strong>and</strong> Planetary Science Letters, 176, 401–411.Hasalova´, P., Schulmann, K., Lexa, O. et al., 2008a. Origin <strong>of</strong>migmatites by deformation-enhanced melt infiltration <strong>of</strong>orthogneiss: a new model based on quantitative micro<strong>structural</strong>analysis. Journal <strong>of</strong> Metamorphic Geology, 26,29–53.Ó 2010 Blackwell Publishing Ltd201


100 O. LEXA ET AL.Hasalova´, P., Sˇtı´pska´, P., Powell, R., Schulmann, K., Janousˇek,V. & Lexa, O., 2008b. Transforming mylonitic metagranite byopen-system interactions during melt flow. Journal <strong>of</strong> MetamorphicGeology, 26, 55–80.Hete´nyi, G., Cattin, R., Brunet, F. et al., 2007. Density distribution<strong>of</strong> the India plate beneath the Tibetan plateau:geophysical <strong>and</strong> petrological constraints on the kinetics <strong>of</strong>lower-crustal eclogitization. Earth <strong>and</strong> Planetary Science Letters,264, 226–244.Holl<strong>and</strong>, T.J.B. & Powell, R., 1998. An internally consistentthermodynamic data set for phases <strong>of</strong> petrological interest.Journal <strong>of</strong> Metamorphic Geology, 16, 309–343.Holl<strong>and</strong>, T.J.B. & Powell, R., 2003. Activity–composition relationsfor phases in petrological calculations: an asymmetricmulticomponent formulation. Contributions to Mineralogy <strong>and</strong>Petrology, 145, 492–501.Holub, F.V., 1997. Ultrapotassic plutonic rocks <strong>of</strong> the durbachiteseries in the Bohemian Massif: petrology, geochemistry<strong>and</strong> petrogenetic interpretation. Journal <strong>of</strong> Geological Sciences-EconomicGeology, Mineralogy, 31, 5–26.Holub, F.V., Cocherie, A. & Rossi, P., 1997. Radiometric dating<strong>of</strong> granitic rocks from the Central Bohemian Plutonic Complex(Czech Republic): constraints on the chronology <strong>of</strong>thermal <strong>and</strong> tectonic events along the Moldanubian–Barr<strong>and</strong>ianboundary. Comptes Rendus de lÕAcademie des SciencesSeries IIA Earth <strong>and</strong> Planetary Science, 325, 19–26.Hrubcova´, P., S´roda, P., Sˇpičák, A. et al., 2005. Crustal <strong>and</strong>uppermost mantle structure <strong>of</strong> the Bohemian Massif based onCELEBRATION 2000 data. Journal <strong>of</strong> Geophysical Research-Solid Earth, 110(B11), B11305.Jakesˇ, P., 1997. Melting in high-P region – case <strong>of</strong> Bohemiangranulites. Acta Universitatis Carolinae Geologica, 41, 113–125.Janousˇek, V., 2006. Saturnin, R language script for application<strong>of</strong> accessory-mineral saturation models in igneous geochemistry.Geologica Carpathica, 57, 131–142.Janousˇek, V. & Holub, F.V., 2007. The causal link betweenHP ⁄ HT metamorphism <strong>and</strong> ultrapotassic magmatism in collisionalorogens: case study from the Moldanubian Zone <strong>of</strong>the Bohemian Massif. Proceedings <strong>of</strong> the GeologistsÕ Association,118, 75–86.Janousˇek, V., Rogers, G. & Bowes, D.R., 1995. Sr-Nd isotopicconstraints on the petrogenesis <strong>of</strong> the Central BohemianPluton, Czech Republic. Geologische Rundschau, 84, 520–534.Janousˇek, V., Holub, F.V. & Gerdes, A., 2003. K-rich magmatismin the Moldanubian Unit, Bohemian Massif – a complexstory featuring variably enriched lithospheric mantle melts<strong>and</strong> their Interaction with the crust. Geolines, 16, 48–49.Janousˇek, V., Braithwaite, C.J.R., Bowes, D.R. & Gerdes, A.,2004a. Magma-mixing in the genesis <strong>of</strong> Hercynian calcalkalinegranitoids: an integrated petrographic <strong>and</strong> geochemicalstudy <strong>of</strong> the Sa´zava intrusion, Central Bohemian Pluton,Czech Republic. Lithos, 78, 67–99.Janousˇek, V., Finger, F., Roberts, M., Frýda, J., Pin, C. &Dolejsˇ, D., 2004b. Deciphering the petrogenesis <strong>of</strong> deeplyburied granites: whole-rock geochemical constraints on theorigin <strong>of</strong> largely undepleted felsic granulites from the MoldanubianZone <strong>of</strong> the Bohemian Massif. Transactions <strong>of</strong> theRoyal Society <strong>of</strong> Edinburgh, Earth Sciences, 95, 141–159.Janousˇek, V., Gerdes, A., Vra´na, S. et al., 2006. Low-pressuregranulites <strong>of</strong> the Lisˇov Massif, Southern Bohemia: Vise´anmetamorphism <strong>of</strong> late Devonian plutonic arc rocks. Journal <strong>of</strong>Petrology, 47, 705–744.Janousˇek, V., Krenn, E., Finger, F., Mı´kova´, J. & Frýda, J.,2007. Hyperpotassic granulites from Blanský les (MoldanubianZone, Bohemian Massif) revisited. Journal <strong>of</strong> Geosciences,52, 73–112.Janousˇek, V., Vrána, S., Erban, V., Vokurka, K. & Dra´bek, M.,2008. Metabasic rocks in the Varied Group <strong>of</strong> the MoldanubianZone, southern Bohemia – their petrology, geochemicalcharacter <strong>and</strong> possible petrogenesis. Journal <strong>of</strong> Geosciences,53, 31–64.Janousˇek, V., Wieg<strong>and</strong>, B.A. & Zˇák, J., 2010. Dating the onset<strong>of</strong> Variscan crustal exhumation in the core <strong>of</strong> the BohemianMassif: new U–Pb single zircon ages from the high-K calcalkalinegranodiorites <strong>of</strong> the Blatna´ suite, Central BohemianPlutonic Complex. Journal <strong>of</strong> the Geological Society, 167, 347–360.Kosˇler, J., Aftalion, M. & Bowes, D.R., 1993. Mid–late Devonianplutonic activity in the Bohemian Massif: U–Pb zirconisotopic evidence from the Staré Sedlo <strong>and</strong> Mirotice gneisscomplexes, Czech Republic. Neues Jahrbuch fu¨r Mineralogie-Monatshefte, 1993, 417–431.Kotkova´, J. & Harley, S., 1999. Formation <strong>and</strong> evolution <strong>of</strong>high-pressure leucogranulites: experimental constraints <strong>and</strong>unresolved issues. Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth, Part A,24, 299–304.Kotkova´, J., Schaltegger, U. & Leichmann, J., 2010. Two types <strong>of</strong>ultrapotassic plutonic rocks in the Bohemian Massif – coevalintrusions at different crustal levels. Lithos, 115, 163–176.Kramers, J., Kreissig, K. & Jones, M., 2001. Crustal heat production<strong>and</strong> style <strong>of</strong> metamorphism: a comparison betweentwo Archean high grade provinces in the Limpopo Belt,southern Africa. Precambrian Research, 112, 149–163.Kro¨ ner, A., OÕBrien, P.J., Nemchin, A. & Pidgeon, R., 2000.Zircon ages for high pressure granulites from South Bohemia,Czech Republic, <strong>and</strong> their connection to Carboniferous hightemperature processes. Contributions to Mineralogy <strong>and</strong>Petrology, 138, 127–142.Kusiak, M., Dunkley, D., Suzuki, K. et al., 2010. Chemical(non-isotopic) <strong>and</strong> isotopic dating <strong>of</strong> Phanerozoic zircon – acase study <strong>of</strong> durbachite from the Trˇebı´cˇ Pluton, BohemianMassif. Gondwana Research, 17, 153–161.Le Pichon, X., Henry, P. & G<strong>of</strong>fe´, B., 1997. Uplift <strong>of</strong> Tibet: fromeclogites to granulites – implications for the Andean Plateau<strong>and</strong> the Variscan belt. Tectonophysics, 273, 57–76.Medaris, L.G., Jelínek, E. & Mı´sarˇ, Z., 1995. Czech eclogites.Terrane settings, interpretation for Variscan tectonic evolution<strong>of</strong> the Bohemian Massif. European Journal <strong>of</strong> Mineralogy, 7,7–28.Mlcˇoch, B. & Konopa´sek, J., 2010. Pre-late Carboniferousgeology along the contact <strong>of</strong> the Saxothuringian <strong>and</strong> Tepla´–Barr<strong>and</strong>ian zones in the area covered by younger sediments<strong>and</strong> volcanics (western Bohemian Massif, Czech Republic).Journal <strong>of</strong> Geosciences, 55, 81–94.OÕBrien, P.J., 2000. The fundamental Variscan problem:high-temperature metamorphism at different depths <strong>and</strong> highpressuremetamorphism at different temperatures. In: OrogenicProcesses: Quantification <strong>and</strong> Modelling in the VariscanBelt, Vol. 179 (eds Franke, W., Haak, V., Oncken, O. &Tanner, D.), pp. 369–386, Geological Society, Special Publications,London.OÕBrien, P.J. & Ro¨ tzler, J., 2003. High-pressure granulites: formation,recovery <strong>of</strong> peak conditions <strong>and</strong> implications fortectonics. Journal <strong>of</strong> Metamorphic Geology, 21, 3–20.OÕBrien, P.J. & Vra´na, S., 1995. Eclogites with a short-livedgranulite-facies overprint in the Moldanubian Zone, CzechRepublic – petrology, geochemistry <strong>and</strong> diffusion modeling <strong>of</strong>garnet zoning. Geologische Rundschau, 84, 473–488.Oncken, O., Hindle, D., Kley, J., Elger, K., Victor, P. &Schemmann, K., 2006. Deformation <strong>of</strong> the Central Andeanupper plate system – facts, fiction, <strong>and</strong> constraints for plateaumodels. In: The Andes–Active Subduction Orogeny (eds Oncken,O.,Chong,G.,Franz,G.,Giese,P.,Go¨ tze, H.-J., Ramos, V.A.,Strecker, M.R. & Wigger, P.), pp. 3–27, Springer-Verlag, Heidelberg.Pearce, J. & Stern, R., 2006. Origin <strong>of</strong> back-arc basin magmas:trace element <strong>and</strong> isotope perspectives. In: Back-Arc SpreadingSystems: Geological, Biological, Chemical, <strong>and</strong> Physical Interactions(eds Christie, D.M., Fisher, C.R., Lee, S.-M. & Givens,S.), pp. 63–86, American Geophysical Union, GeophysicalMonograph Series 166, Washington.Perchuk, L., 1989. P–T–fluid regimes <strong>of</strong> metamorphism <strong>and</strong>related magmatism with specific reference to the granulite-Ó 2010 Blackwell Publishing Ltd202


HEAT SOURCES AND EXHUMATION MECHANISMS 101facies Sharyzhalgay complex <strong>of</strong> Lake Baikal. In: Evolution <strong>of</strong>Metamorphic Belts (eds Daly, J.S., Cliff, R.A. & Yardley,B.W.D.), pp. 275–292, 43, Blackwell Scientific Publications,Oxford.Petrakakis, K., 1997. Evolution <strong>of</strong> Moldanubian rocks inAustria: review <strong>and</strong> synthesis. Journal <strong>of</strong> Metamorphic Geology,15, 203–222.Pitra, P., Burg, J.-P. & Guiraud, M., 1999. Late Variscan strikesliptectonics between the Tepla-Barr<strong>and</strong>ian <strong>and</strong> Moldanubianterranes (Czech Bohemian Massif): petro<strong>structural</strong> evidence.Journal <strong>of</strong> the Geological Society, London, 156, 1003–1020.Plesch, A. & Oncken, O., 1999. Orogenic wedge growth duringcollision – constraints on mechanics <strong>of</strong> a fossil wedge from itskinematic record (Rhenohercynian FTB, Central Europe).Tectonophysics, 309, 117–139.Powell, R. & Holl<strong>and</strong>, T.J.B., 2004. Course Notes for ÔÔTHER-MOCALC Workshop 2004: Calculating Metamorphic PhaseEquilibriaÕÕ (CD ROM). ETH, Zu¨ rich.Powell, R., Holl<strong>and</strong>, T.J.B. & Worley, B., 1998. Calculatingphase diagrams involving solid solutions via non-linearequations, with examples using THERMOCALC. Journal <strong>of</strong>Metamorphic Geology, 16, 577–588.Prince, C.I., Kosˇler, J., Vance, D. & Gu¨ nther, D., 2000. Comparison<strong>of</strong> laser ablation ICP-MS <strong>and</strong> isotope dilution REE<strong>analyses</strong> – implications for Sm-Nd garnet geochronology.Chemical Geology, 168, 255–274.Racek, M., Sˇtı´pska´, P., Pitra, P., Schulmann, K. & Lexa, O.,2006. Metamorphic record <strong>of</strong> burial <strong>and</strong> exhumation <strong>of</strong> orogeniclower <strong>and</strong> middle crust: a new tectonothermal model forthe Drosendorf window (Bohemian Massif, Austria). Mineralogy<strong>and</strong> Petrology, 86, 221–251.Racek, M., Sˇtı´pska´, P. & Powell, R., 2008. Garnet–clinopyroxeneintermediate granulites in the St. Leonhard massif <strong>of</strong> theBohemian Massif: ultrahigh-temperature metamorphism athigh pressure or not? Journal <strong>of</strong> Metamorphic Geology, 26,253–271.Ramberg, H., 1981. Gravity Deformation <strong>and</strong> the EarthÕs Crust:In Theory, Experiments, <strong>and</strong> Geological Application. AcademicPress, London.Rey, P., V<strong>and</strong>erhaeghe, O. & Teyssier, C., 2001. Gravitationalcollapse <strong>of</strong> the continental crust: definition, regimes <strong>and</strong>modes. Tectonophysics, 342, 435–449.Roberts, M. & Finger, F., 1997. Do U–Pb zircon ages fromgranulites reflect peak metamorphic conditions? Geology, 25,319–322.Roering, C., Van Reenen, D., Smit, C. et al., 1992. Tectonicmodel for the evolution <strong>of</strong> the Limpopo Belt. PrecambrianResearch, 55, 539–552.Rossi, P., Oggiano, G. & Cocherie, A., 2009. A restored section<strong>of</strong> the Ôsouthern Variscan realmÕ across the Corsica-Sardiniamicrocontinent. Comptes Rendus Geosciences, 341, 224–238.Rubatto, D., Ferr<strong>and</strong>o, S., Compagnoni, R. & Lombardo, B.,2010. Carboniferous high-pressure metamorphism <strong>of</strong> Ordovicianprotoliths in the Argentera Massif (Italy), SouthernEuropean Variscan belt. Lithos, 116, 65–76.Rudnick, R.L. & Gao, S., 2003. Composition <strong>of</strong> the ContinentalCrust. In: Treatise on Geochemistry (eds Holl<strong>and</strong>, H.D. &Turekian, K.K.), pp. 1–64. Pergamon, Oxford.Rudnick, R.L. & Presper, T., 1990. Geochemistry <strong>of</strong> intermediate-to high-pressure granulites. In: Granulites <strong>and</strong> CrustalEvolution (eds Vielzeuf, D. & Vidal, P.), pp. 523–550. KluwerAcademic Publishers, Dordrecht.Ru˚zˇek, B., Hrubcova´, P., Novotny´, M., Sˇpicˇa´k, A. & Karousova´,O., 2007. Inversion <strong>of</strong> travel times obtained during activeseismic refraction experiments CELEBRATION 2000, ALP2002 <strong>and</strong> SUDETES 2003. Studia Geophysica et Geodaetica,51, 141–164.Scheuvens, D., 2002. Metamorphism <strong>and</strong> microstructures alonga high-temperature metamorphic field gradient: the north–eastern boundary <strong>of</strong> the Kra´lovsky´ Hvozd Unit (BohemianMassif, Czech Republic). Journal <strong>of</strong> Metamorphic Geology, 20,413–428.van Schmus, W., 1995. Natural radioactivity <strong>of</strong> the crust <strong>and</strong>mantle. In: Global Earth Physics: A H<strong>and</strong>book <strong>of</strong> PhysicalConstants (ed. Ahrens, T.J.), pp. 283–291. American GeophysicalUnion, AGU Reference Shelf Series, Washington,DC.Schulmann, K., Kro¨ ner, A., Hegner, E. et al., 2005. Chronologicalconstraints on the pre-orogenic history, burial <strong>and</strong>exhumation <strong>of</strong> deep-seated rocks along the eastern margin <strong>of</strong>the Variscan orogen, Bohemian Massif, Czech Republic.American Journal <strong>of</strong> Science, 305, 407–448.Schulmann, K., Lexa, O., Sˇtı´pska´, P. et al., 2008. Verticalextrusion <strong>and</strong> horizontal channel flow <strong>of</strong> orogenic lower crust:key exhumation mechanisms in large hot orogens? Journal <strong>of</strong>Metamorphic Geology, 26, 273–297.Schulmann, K., Konopa´sek, J., Janousˇek, V. et al., 2009. AnAndean type Palaeozoic convergence in the Bohemian Massif.Comptes Rendus-Ge´oscience, 341, 266–286.Siebel, W., Raschka, H., Irber, W. et al., 1997. Early Palaeozoicacid magmatism in the Saxothuringian belt: new insights froma geochemical <strong>and</strong> isotopic study <strong>of</strong> orthogneisses <strong>and</strong> metavolcanicrocks from the Fichtelgebirge, SE Germany. Journal<strong>of</strong> Petrology, 38, 203–230.Sláma, J., Kosˇler, J., Condon, D. et al., 2008. Plesˇovice zircon –a new natural reference material for U–Pb <strong>and</strong> Hf isotopicmicroanalysis. Chemical Geology, 249, 1–35.Sˇtı´pská, P. & Powell, R., 2005a. Constraining the P-T path <strong>of</strong> aMORB-type eclogite using pseudosections, garnet zoning <strong>and</strong>garnet-clinopyroxene thermometry: an example from theBohemian Massif. Journal <strong>of</strong> Metamorphic Geology, 23, 725–743.Sˇtı´pská, P. & Powell, R., 2005b. Does ternary feldspar constrainthe metamorphic conditions <strong>of</strong> high-grade meta-igneousrocks? Evidence from orthopyroxene granulites, BohemianMassif Journal <strong>of</strong> Metamorphic Geology, 23, 627–647.Sˇtı´pská, P., Schulmann, K. & Kro¨ ner, A., 2004. Vertical extrusion<strong>and</strong> middle crustal spreading <strong>of</strong> omphacite granulite: amodel <strong>of</strong> syn-convergent exhumation (Bohemian Massif,Czech Republic). Journal <strong>of</strong> Metamorphic Geology, 22, 179–198.Sˇtı´pská, P., Schulmann, K. & Powell, R., 2008. Contrastingmetamorphic histories <strong>of</strong> lenses <strong>of</strong> high-pressure rocks <strong>and</strong>host migmatites with a flat orogenic fabric (Bohemian Massif,Czech Republic): a result <strong>of</strong> tectonic mixing within horizontalcrustal flow? Journal <strong>of</strong> Metamorphic Geology, 26, 623–646.Tajčmanova´, L., Konopa´sek, J. & Schulmann, K., 2006. Thermalevolution <strong>of</strong> the orogenic lower crust during exhumationwithin a thickened Moldanubian root <strong>of</strong> the Variscan belt <strong>of</strong>Central Europe. Journal <strong>of</strong> Metamorphic Geology, 24, 119–134.Tajčmanova´, L., Soejono, I., Konopa´sek, J., Kosˇler, J. &Klo¨ tzli, U., 2010. Structural position <strong>of</strong> high-pressure felsic tointermediate granulites from NE Moldanubian domain(Bohemian Massif). Journal <strong>of</strong> the Geological Society, 167,329–345.Tomek, Cˇ., Dvorˇákova´, V. & Vra´na, S., 1997. Geologicalinterpretation <strong>of</strong> the 9HR <strong>and</strong> 503M seismic pr<strong>of</strong>iles inwestern Bohemia. In: Geological Model <strong>of</strong> Western BohemiaRelated to the KTB Borehole in Germany (eds Vra´na, S. &Sˇtědrá, V.), pp. 43–50, Journal <strong>of</strong> Geological Sciences, Geology,47, Prague.V<strong>and</strong>erhaeghe, O., 2009. Migmatites, granites <strong>and</strong> orogeny: flowmodes <strong>of</strong> partially-molten rocks <strong>and</strong> magmas associated withmelt ⁄ solid segregation in orogenic belts. Tectonophysics, 477,119–134.V<strong>and</strong>erhaeghe, O. & Teyssier, C., 2001. Partial melting <strong>and</strong> flow<strong>of</strong> orogens. Tectonophysics, 342, 451–472.Vellmer, C., 1992. St<strong>of</strong>fbest<strong>and</strong> und Petrogenese von Granulitenund Granitischen Gesteinen der Su¨dlichen Bo¨hmischen Masse inNiedero¨sterreich. Unpublished PhD thesis, Georg-August-Universita¨ t, Go¨ ttingen.Venera, Z., Schulmann, K. & Kro¨ ner, A., 2000. Intrusion withina transtensional tectonic domain: the Cˇista´ granodioriteÓ 2010 Blackwell Publishing Ltd203


102 O. LEXA ET AL.(Bohemian Massif) – structure <strong>and</strong> rheological <strong>modelling</strong>.Journal <strong>of</strong> Structural Geology, 22, 1437–1454.Verner, K., Zˇák, J., Nahodilova´, R. & Holub, F., 2008. Magmaticfabrics <strong>and</strong> emplacement <strong>of</strong> the cone-sheet-bearingKnízˇecı´ Stolec durbachitic pluton (Moldanubian Unit, BohemianMassif): implications for mid-crustal reworking <strong>of</strong>granulitic lower crust in the Central European Variscides.International Journal <strong>of</strong> Earth Sciences, 97, 19–33.Vigneresse, J., Barbey, P. & Cuney, M., 1996. Rheologicaltransitions during partial melting <strong>and</strong> crystallization withapplication to felsic magma segregation <strong>and</strong> transfer. Journal<strong>of</strong> Petrology, 37, 1879–1600.Vrána, S., 1989. Perpotassic granulites from southern Bohemia.Contributions to Mineralogy <strong>and</strong> Petrology, 103, 510–522.Vrána, S. & Jakesˇ, P., 1982. Orthopyroxene <strong>and</strong> two-pyroxenegranulites from a segment <strong>of</strong> charnockitic crust in southernBohemia. Bulletin <strong>of</strong> the Czech Geological Survey, 57, 129–143.Watson, E., Vicenzi, E. & Rapp, R., 1989. Inclusion ⁄ host relationsinvolving accessory minerals in high-grade metamorphic<strong>and</strong> anatectic rocks. Contributions to Mineralogy <strong>and</strong> Petrology,101, 220–231.Weber, K., 1984. Variation in tectonic style with time (Variscan<strong>and</strong> Proterozoic systems). In: Patterns <strong>of</strong> Change in EarthEvolution (eds Holl<strong>and</strong>, H.D. & Trendall, A.F.), pp. 371–386.Springer Verlag, Berlin.Wendt, J.I., Kro¨ ner, A., Fiala, J. & Todt, W., 1993. Evidencefrom zircon dating for existence <strong>of</strong> approximately 2.1 Ga oldcrystalline basement in southern Bohemia, Czech Republic.Geologische Rundschau, 82, 42–50.Wenzel, T., Mertz, D., Oberha¨ nsli, R., Becker, T. & Renne, P.,1997. Age, geodynamic setting, <strong>and</strong> mantle enrichment processes<strong>of</strong> a K-rich intrusion from the Meissen Massif (northernBohemian Massif) <strong>and</strong> implications for related occurrencesfrom the mid-European Hercynian. Geologische Rundschau,86, 556–570.Wenzel, T., Oberha¨ nsli, R. & Mezger, K., 2000. K-rich plutonicrocks <strong>and</strong> lamprophyres from the Meissen Massif (northernBohemian Massif): geochemical evidence for variably enrichedlithospheric mantle sources. Neues Jahrbuch fu¨r Mineralogie,Abh<strong>and</strong>lungen, 175, 249–293.White, R.W. & Powell, R., 2002. Melt loss <strong>and</strong> the preservation<strong>of</strong> granulite facies mineral assemblages. Journal <strong>of</strong> MetamorphicGeology, 20, 621–632.White, R.W., Powell, R., Holl<strong>and</strong>, T.J.B. & Worley, B.A., 2000.The effect <strong>of</strong> TiO 2 <strong>and</strong> Fe 2 O 3 on metapelitic assemblages atgreenschist <strong>and</strong> amphibolite facies conditions: mineral equilibriacalculations in the system K 2 O-FeO-MgO-Al 2 O 3 -SiO 2-H 2 O-TiO 2 -Fe 2 O 3 . Journal <strong>of</strong> Metamorphic Geology, 18, 497–511.White, R.W., Powell, R. & Halpin, J.A., 2004. Spatiallyfocussedmelt formation in aluminous metapelites fromBroken Hill, Australia. Journal <strong>of</strong> Metamorphic Geology, 22,825–845.White, R.W., Powell, R. & Holl<strong>and</strong>, T.J.B., 2007. Progressrelating to calculation <strong>of</strong> partial melting equilibria formetapelites. Journal <strong>of</strong> Metamorphic Geology, 25, 511–527.Whittington, A.G., H<strong>of</strong>meister, A.M. & Nabelek, P.I., 2009.Temperature-dependent thermal diffusivity <strong>of</strong> the EarthÕscrust <strong>and</strong> implications for magmatism. Nature, 458(7236),319–321.Wieg<strong>and</strong>, B., 1997. Isotopengeologische und geochemischeUntersuchungen zur pra¨ variszischen magmatischen und sedimenta¨ren Entwicklung im saxothuringisch–moldanubischenU¨ bergangsbereich (Grenzgebiet BRD ⁄ CR). GeotektonischeForschungen, 88, 1–177.Ža´k, J., Holub, F. & Verner, K., 2005. Tectonic evolution <strong>of</strong> acontinental magmatic arc from transpression in the uppercrust to exhumation <strong>of</strong> mid-crustal orogenic root recorded byepisodically emplaced plutons: the Central Bohemian PlutonicComplex (Bohemian Massif). International Journal <strong>of</strong> EarthSciences, 94, 385–400.Ža´k, J., Kratinova´, Z., Trubač, J., et al., 2010. Structure,emplacement, <strong>and</strong> tectonic setting <strong>of</strong> Late Devonian granitoidplutons in the Tepla´-Barr<strong>and</strong>ian unit, Bohemian Massif.International Journal <strong>of</strong> Earth Sciences, in press. doi:10.1007 ⁄s00531-010-0565-7.Za´vada, P., Schulmann, K., Konopa´sek, J., Ulrich, S. & Lexa,O., 2007. Extreme ductility <strong>of</strong> feldspar aggregates – melt-enhancedgrain boundary sliding <strong>and</strong> creep failure: rheologicalimplications for felsic lower crust. Journal <strong>of</strong> GeophysicalResearch, 112(B10), 1–15.Received 16 March 2010; revision accepted 18 August 2010.Ó 2010 Blackwell Publishing Ltd204


J. metamorphic Geol., 2011, 29, 53–78 doi:10.1111/j.1525-1314.2010.00903.xModel <strong>of</strong> syn-convergent extrusion <strong>of</strong> orogenic lower crust in thecore <strong>of</strong> the Variscan belt: implications for exhumation <strong>of</strong>high-pressure rocks in large hot orogensJ. FRANĚK, 1,2 K. SCHULMANN, 3 O. LEXA, 1,2 Č.TOMEK 1 AND J.-B. EDEL 31 Czech Geological Survey, Klárov 3, 118 21 Prague, Czech Republic (jan.franek@geology.cz)2 Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University, Albertov 6, 128 43 Prague, Czech Republic3 Institut de Physique du Globe de Strasbourg, IPGS – UMR 7516, CNRS et Université de Strasbourg (EOST), 1 Rue Blessig,67084 Strasbourg, FranceABSTRACTReflection seismic section, field <strong>structural</strong> analysis <strong>and</strong> gravimetric <strong>modelling</strong> <strong>of</strong> orogenic lower crust inthe core <strong>of</strong> a Carboniferous orogenic root reveal details <strong>of</strong> the polyphase process <strong>of</strong> exhumation.Subvertical amphibolite facies fabrics strike parallel to former plate margins that collided in the NW.The fabrics are developed in both mid-crustal <strong>and</strong> lower crustal high-pressure granulite units as a result<strong>of</strong> intensive NW–SE intraroot horizontal shortening driven probably by the west-directed collision. Ingranulites, the steep fabrics originated as a result <strong>of</strong> extrusion <strong>of</strong> orogenic lower crust in a 20 km widevertical ascent channel from lower crustal depths at 350–340 Ma. The large granulite bodies preserveolder granulite facies fabrics documenting a two-stage evolution during the exhumation process. Surfaceexposures <strong>of</strong> granulites coincide with the absence <strong>of</strong> subhorizontal seismic reflectors at depth, suggestingpreservation <strong>of</strong> the 20 km wide subvertical tabular structure reaching Moho depths. Horizontalseismic reflectors surrounding the vertical channel structure corroborate a dominant flat migmatiticfabric developed in all tectonic units. This <strong>structural</strong> pattern is interpreted in terms <strong>of</strong> subhorizontalspreading <strong>of</strong> partially molten orogenic lower crust in mid-crustal levels (765 °C <strong>and</strong> 0.76 GPa) at 342–337 Ma. Large massifs <strong>of</strong> extruded <strong>and</strong> progressively dismembered felsic granulites disturbed midcrustalfabrics in the surrounding horizontally flowing partially molten crust. The horizontal mid-crustalflow resulted in collapse <strong>of</strong> the supra-crustal Teplá-Barr<strong>and</strong>ian Unit (interpreted as the orogenic lid)along a large-scale crustal detachment above the extruded lower crustal dome. The presence <strong>of</strong> felsicgranulites at the bottom <strong>of</strong> the orogenic root is considered to be a key factor controlling the exhumation<strong>of</strong> orogenic lower crust in large hot orogens.Key words: exhumation; granulites; Moldanubian domain; orogenic lower crust; Variscan belt.INTRODUCTIONCurrent models for the exhumation <strong>of</strong> deep-seatedcrustal rocks <strong>of</strong>ten focus on processes along suturezones that <strong>of</strong>fer a pre-existing discontinuity alongwhich the deep-seated rocks can be transported upwards.These models mimic a classical concept <strong>of</strong>buoyancy-driven exhumation <strong>of</strong> subducted crustal slicesalong the subduction channel (e.g. Chemendaet al., 1995) simulated <strong>numerical</strong>ly by Gerya &Stockhert (2006), Gerya et al. (2008) <strong>and</strong> Beaumontet al. (2009) for orogens characterized by continentalsubduction. In contrast, Platt (1993) proposed a cornerflow regime operating on a crustal scale between asubduction zone <strong>and</strong> a rigid buttress <strong>of</strong> the overridingplate, which evokes rise <strong>of</strong> lower crustal material alongits edge. Burg & Podladchikov (1999) <strong>of</strong>fered <strong>numerical</strong>models <strong>of</strong> lithospheric-scale buckling, where thebuckled crust may undergo fragmentation via thrustzones, which can exhume particular blocks >10 kmvertically (Sokoutis et al., 2005).In many cases at the end <strong>of</strong> an orogeny the thickenedcrust undergoes rapid tectonic thinning rather thanslow erosion-related decrease <strong>of</strong> crustal thickness (e.g.Dewey et al., 1993). This process results from gravitydriven extensional collapse <strong>of</strong> the orogen causingdevelopment <strong>of</strong> pervasive flat crustal fabrics (Koyiet al., 1999). Additionally, heat redistribution leadingto weakening <strong>of</strong> the deep crustal rocks via partialmelting may facilitate collapse <strong>of</strong> a hot <strong>and</strong> thickenedorogenic root (Rey et al., 2001, 2009). Subsequentlateral ductile- to channel-flow, driven by a combination<strong>of</strong> tectonic <strong>and</strong> overburden gravity force, mayresult in exhumation from high pressure (HP) conditionsas modelled by Beaumont et al. (2004) <strong>and</strong>Jamieson et al. (2007).The Variscan orogen in centralEurope is characterized by numerous occurrences <strong>of</strong>HP rocks located far from a suture. The HP rocks areÓ 2010 Blackwell Publishing Ltd 53205


54 J. FRANĚK ET AL.represented mostly by large bodies <strong>of</strong> HP felsic granuliteswhich form areas sometimes several 100 km 2 insize (OÕBrien & Carswell, 1993; Janousˇek & Holub,2007). These HP granulites are associated with otherHP rocks like mafic granulites, eclogites <strong>and</strong> garnetperidotites collectively forming the so-called Gfo¨ hlUnit (Fuchs, 1976). This unit does not form a continuousHP body, but occurs in three major belts, whichare surrounded by medium-grade rocks (Schulmannet al., 2009). The protolith <strong>of</strong> these rocks, their positionat lower crustal depths prior to exhumation <strong>and</strong>the exhumation mechanisms represent major problems<strong>of</strong> the Variscan belt (e.g. Behr, 1961, 1980). Schulmannet al. (2005) proposed a multistage model <strong>of</strong> rapidexhumation <strong>of</strong> orogenic lower crust associated withdevelopment <strong>of</strong> subvertical ascent channels triggeredby rapid amplification <strong>of</strong> initial instabilities in front <strong>of</strong>an advancing continental buttress. Such a process isenabled by the extremely low viscosity <strong>of</strong> the Variscanorogenic lower crust at the bottom <strong>of</strong> the orogenic rootdue to its unusually hot thermal structure (e.g. OÕBrien,2008) combined with its felsic composition (Ru˚zˇeket al., 2007). This model is fundamentally similar tothat proposed by Weber & Behr (1983) who explainedthe exhumation <strong>of</strong> the Saxonian HP granulites byÔdiapiric foldingÕ. In their model, the deep granulitelayer tends to amplify <strong>and</strong> pierce through the weakermiddle crust during crustal shortening, to form largescalesteep folds bringing HP rocks to mid-crustallevels.The study region is located in the central part <strong>of</strong> theorogenic root, far from both the easterly continentalbuttress <strong>and</strong> westerly suture zone (Fig. 1). Here, weexamine a large portion <strong>of</strong> orogenic lower <strong>and</strong> middlecrust using both the published <strong>and</strong> unpublished part <strong>of</strong>the 9HR seismic line <strong>of</strong> Tomek et al. (1997). Theseismic pr<strong>of</strong>ile is combined with a detailed <strong>structural</strong>study along a SE–NW traverse evaluating mutualrelationships between middle orogenic crust <strong>and</strong> thethree major granulite massifs <strong>of</strong> South Bohemia, whichcompletes an earlier study <strong>of</strong> one <strong>of</strong> these massifs(Franeˇk et al., 2006). The pr<strong>of</strong>ile line drawing is combinedwith detailed gravity forward <strong>modelling</strong> allowingestimation <strong>of</strong> a vertical ascent channel <strong>of</strong> orogeniclower crustal rocks. Finally, we discuss the development<strong>of</strong> flat-lying amphibolite facies fabrics throughoutthe traverse, which provides evidence <strong>of</strong> mid-crustalhorizontal flow. It is shown that this episode is connectedwith the collapse <strong>of</strong> the orogenic suprastructurerelated to final upwelling <strong>of</strong> a lower crustal dome.Despite disruption <strong>of</strong> the linear Variscan orogenictrend by late large-scale transcurrent shear zones (Edelet al., 2003), the wealth <strong>of</strong> quantitative geological datamakes the Bohemian Massif a suitable field laboratoryfor underst<strong>and</strong>ing processes related to exhumation <strong>of</strong>orogenic lower crust in large hot orogens. The fieldresults are suitable for testing the results from largescaleexhumation-related <strong>numerical</strong> models referencedabove.GEOLOGICAL SETTINGThe Bohemian Massif (Fig. 1) is traditionally dividedinto the Saxothuringian domain to the west, the Tepla´-Barr<strong>and</strong>ian <strong>and</strong> Moldanubian domains in the centralpart <strong>of</strong> the Massif (Kossmat, 1927) <strong>and</strong> the Brunovistulian(Brunia) Neoproterozoic continent to the east(Dudek, 1980). Schulmann et al. (2005, 2009) interpretedthe Bohemian Massif as a Gondwana-derivedcollisional domain characterized by: (i) relicts <strong>of</strong> a twostageSE-directed subduction at the Saxothuringian–Tepla´-Barr<strong>and</strong>ian boundary; (ii) a magmatic arcgenetically related to the subduction represented bythe Central Bohemian Plutonic Complex in the centre;<strong>and</strong> (iii) the rigid forel<strong>and</strong> represented by the Bruniamicroplate in the SE. In this concept, the Tepla´-Barr<strong>and</strong>ian between the suture zone <strong>and</strong> the magmaticarc represents the fore-arc domain <strong>and</strong> the Moldanubi<strong>and</strong>omain between the magmatic arc <strong>and</strong> the Bruniamicroplate constitutes the shortened <strong>and</strong> thickenedintracontinental back-arc region. Large Variscanstrike-slip zones (e.g. the Elbe Fault Zone, seeFig. 1a,b) strike NW–SE <strong>and</strong> dismember the NNEtrending Variscan structure <strong>of</strong> the Bohemian Massif(Edel & Weber, 1995).Geology <strong>of</strong> the Saxothuringian, Teplá-Barr<strong>and</strong>ian <strong>and</strong>Moldanubian domainsThe SE part <strong>of</strong> the Saxothuringian domain consists <strong>of</strong>the antiformal Erzgebirge Crystalline Complex, whichcan be divided into a lowermost para-autochtonousdomain that is overlain by crystalline nappes thatrecord peak HP metamorphism at 345–340 Ma(e.g. Franke, 2000; Konopa´sek & Schulmann, 2005).Towards the NW, the antiform passes into a Cambrianto Lower Carboniferous volcano-sedimentarysequence, which crops out in several large-scale latetectonic antiforms <strong>and</strong> synforms (Franke, 1993). Thesesupracrustal rocks are overthrust by oceanic crustalunits (Mu¨ nchberg, Frankenberg <strong>and</strong> Wildenfels klippen),metamorphosed at eclogite facies at c. 395–380 Ma (e.g. Dallmayer et al., 1995; Franke, 2000).The Saxonian granulite massif emerges as a large-scaleNE–SW elongated dome structure from below theSaxothuringian Basin (Du¨ rbaum et al., 1999) <strong>and</strong>consists <strong>of</strong> HP felsic granulite, with subordinate lenses<strong>of</strong> mafic granulite <strong>and</strong> serpentinized peridotite (Ro¨ tzler& Romer, 2001). Kro¨ ner & Willner (1998) obtainedages <strong>of</strong> 485–470 Ma from zircon cores interpreted todate the protolith, while c. 340 Ma overgrowths wereinterpreted to date the time <strong>of</strong> peak metamorphism.The Tepla´-Barr<strong>and</strong>ian domain is separated from theSaxothuringian domain by the Variscan SE-dippingsuture zone represented by the Maria´nske´ La´zněComplex (e.g. Zulauf et al., 1997). This unit consists<strong>of</strong> serpentinites <strong>and</strong> metagabbros <strong>of</strong> Cambrian <strong>and</strong>Ordovician age (Timmermann et al., 2004), which werein part eclogitized before Devonian exhumation. TheÓ 2010 Blackwell Publishing Ltd206


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 55(a)(b)Fig. 1. (a) Principal divisions <strong>of</strong> the Variscan chain. RH, Rhenohercynian domain; SX, Saxothuringian domain; MO, Moldanubi<strong>and</strong>omain. (b) Simplified geological map <strong>of</strong> the Bohemian Massif modified after Franke (2000).Tepla´-Barr<strong>and</strong>ian domain consists <strong>of</strong> primitiveNeoproterozoic siltstones to greywackes <strong>and</strong> volcanicrocks deposited on an oceanic or transitional crust(e.g. Cháb & Pelc, 1973) <strong>and</strong> weakly metamorphosedduring the Cadomian orogeny (Kettner, 1918). ThisCadomian basement is unconformably overlain byCambrian <strong>and</strong> Ordovician to mid-Devonian sedimentarysequences that consist <strong>of</strong> greywackes, shales,s<strong>and</strong>stones <strong>and</strong> limestones. The Variscan tectonometamorphicprocesses are most pronounced in thewestern part, where they are restricted to 385–360 Ma(Fig. 2; Appendix S1). The boundary to the east withthe migmatites <strong>of</strong> the Moldanubian domain is markedby intense strike-slip deformation along the so-calledCentral Bohemian Shear Zone (Rajlich, 1988; Pitraet al., 1999; Scheuvens & Zulauf, 2000), however, thetrue contact is <strong>of</strong>ten masked by 375–336 Ma (Fig. 2)intrusions <strong>of</strong> the Central Bohemian Plutonic Complex(e.g. Zˇa´k et al., 2005a). The boundary betweenMoldanubian migmatites <strong>and</strong> the SW part <strong>of</strong> theTepla´-Barr<strong>and</strong>ian domain is marked by the so-calledWest-Bohemian Shear Zone (Zulauf et al., 2002; Do¨ rr& Zulauf, 2008).The Moldanubian domain corresponds to theinternal orogenic root zone <strong>of</strong> the Variscan orogen(Suess, 1926). It is intruded by numerous Variscanplutons ranging from I-type (e.g. specifically K–Mgrich syenites or calc-alkaline arc-related rocks, Janousˇeket al., 2000) to S-type granitoids (Finger & Steyer,1995). The Moldanubian–Brunia boundary is markedby a several-kilometres-wide zone <strong>of</strong> highly deformedme´lange derived from both the units (Konopa´seket al., 2002). The Moldanubian domain is traditionallydivided into three tectonic units (Fuchs, 1976). The<strong>structural</strong>ly deepest medium-grade MonotonousGroup is overlain by the medium-grade Varied Group<strong>and</strong> the <strong>structural</strong>ly highest high-grade Gfo¨ hl Unit.The Monotonous Group consists <strong>of</strong> biotite-plagioclaseparagneiss with minor orthogneiss, quartzite,amphibolite <strong>and</strong> locally eclogite bodies (Medaris et al.,1995; OÕBrien & Vra´na, 1995). It contains metagranitoids<strong>of</strong> Early Palaeozoic age (see Fig. 2), suggesting aPrecambrian age for the protolith <strong>of</strong> the surroundingparagneiss (Finger & von Quadt, 1995; Friedl et al.,2004). The Varied Group includes more pelitic protoliths<strong>and</strong> abundant amphibolite, quartzite, marble <strong>and</strong>Ó 2010 Blackwell Publishing Ltd207


56 J. FRANĚK ET AL.Fig. 2. Summary <strong>of</strong> geochronology for the individual units in the Moldanubian, Tepla´-Barr<strong>and</strong>ian <strong>and</strong> Saxothuringian domains.Histograms <strong>of</strong> larger datasets are mixed with shaded columns that contain individual age data combined with the stratigraphy <strong>of</strong>associated Palaeozoic sedimentary rocks. Histograms represent the sum <strong>of</strong> Gaussian probability curves for each radiometric ageincluded. Citations for the 263 plotted ages are given in the Supporting information, Appendix S1. Significant ages are plottedindividually with corresponding blocking temperature. Apatite fission track dating in the Tepla´-Barr<strong>and</strong>ian was performed byGlasmacher et al. (2002), the monazite U–Pb dating from the Moldanubian pegmatite is from Novák et al. (1998). Sx-TB, Saxothuringian-Tepla´-Barr<strong>and</strong>ian;MLC, Maria´nské La´zně Complex; MK, Mu¨ nchberg Klippe; ZEV, Zone Erbendorf-Vohenstrauss; ECC,Eger Crystalline Complex; CBPC, Central Bohemian Plutonic Complex; BPSZ, Bavarian Moldanubicum with Pfahl Shear Zone.calc-silicate intercalations. The protolith <strong>of</strong> the VariedGroup metasedimentary rocks is supposed to be atleast partly Devonian (Friedl et al., 1993) <strong>and</strong> Cambrian,based on the 509 ± 27 Ma Nd model age forthe amphibolite layers (Janousˇek et al., 1997). TheGfo¨ hl Unit consists <strong>of</strong> kyanite-bearing felsic granulite,peridotite, eclogite <strong>and</strong> migmatitic Gfo¨ hl orthogneiss<strong>of</strong> Cambrian to Early Ordovician protolith age (Friedlet al., 2004; Schulmann et al., 2005).Relevant P–T estimatesThe petrological studies <strong>of</strong> granulites in the BohemianMassif (Fig. 3) have focused traditionally onpeak metamorphic conditions <strong>and</strong> amphibolite faciesretrogression involving almost isothermal decompressionfollowed by near isobaric cooling (e.g.Vra´na et al., 1995; OÕBrien & Ro¨ tzler, 2003). Conventionalthermobarometry yields 1000 °C ⁄ 1.6 GPafor the metamorphic peak, followed by retrogressionat 800–900 °C ⁄ 0.8–1.2 GPa (Vra´na, 1989; OÕBrien &Seifert, 1992; Carswell & O’Brien, 1993; Cooke et al.,2000) <strong>and</strong> 700–800 °C ⁄ 0.5–0.8 GPa (OÕBrien &Carswell, 1993; Vra´na, 1997). However, Sˇtı´pska´ &Powell (2005), Racek et al. (2006) <strong>and</strong> Tajcˇmanova´et al. (2006) proposed significantly lower temperatureconditions <strong>of</strong> 750–850 °C at 1.6–1.8 GPa for thepeak metamorphic event followed by retrogression at700–800 °C <strong>and</strong> 0.5–0.7 GPa. The mantle-derivedrocks enclosed in granulites yield a broad range <strong>of</strong>P–T conditions <strong>of</strong> 800–1350 °C at 2.0–6.0 GPa (seesummary by Medaris et al., 2006). The P–T estimatesfor the Moldanubian mid-crustal rocks yield peakconditions <strong>of</strong> 750 °C ⁄ 0.7–1.2 GPa for the VariedGroup (Petrakakis, 1997; Racek et al., 2006) <strong>and</strong>780 °C ⁄ 0.75 GPa for the Monotonous Group(Scheuvens, 2002), with retrogression to 650–720 °C ⁄0.4–0.5 GPa recorded by both the mid-crustal groups(Vra´na et al., 1995; Pitra et al., 1999; Racek et al.,2006).Ó 2010 Blackwell Publishing Ltd208


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 57Fig. 3. Relevant P–T data from Bohemiangranulites <strong>and</strong> surrounding metasedimentaryrocks; from the Moldanubian domain,if not specified from some other unitbelow. The P–T data <strong>and</strong> related radiometricages (Ma) are assembled from thepublications 1–11, according to the labelsat each polygon. Citations for P–T data –1, Kalt et al. (1999); 2, Kotkova´ (1993) –Eger Crystalline Complex; 3, Kro¨ ner et al.(2000); 4, Linner (1996); 5, Pitra et al.(1999); 6, Racek et al. (2006); 7, Sˇtı´pska´et al. (2004); 8, Sˇtı´pská & Powell (2005);9, Tajcˇmanova´ et al. (2006); 10, Verneret al. (2008); 11, Ro¨ tzler & Romer (2001) –Saxonian Granulites. Citations forgeochronology: 1, Kalt et al. (1997); 2,Kotkova´ et al. (1996) – Eger CrystallineComplex; 3, Kro¨ ner et al. (2000); 5,Gebauer et al. (1989); 7, Sˇtı´pska´ et al.(2004); 9, Schulmann et al. (2005); 10,Verner et al. (2008); 11, Romer & Ro¨ tzler(2001) – Saxonian Granulites.Moldanubian geochronologyMost <strong>of</strong> the existing U–Pb zircon ages from theBohemian Massif granulites cluster at c. 340 Ma(Fig. 2; for review, see e.g. Janousˇek & Holub, 2007),interpreted as a time <strong>of</strong> HP metamorphism <strong>and</strong> subsequentretrogression. The older ages, widely scatteredbetween 450 <strong>and</strong> 350 Ma, are interpreted usually asprotolith ages (e.g. Wendt et al., 1994; Kro¨ ner et al.,2000; Friedl et al., 2003). Sm–Nd garnet data revealslightly older ages than 340 Ma (e.g. Prince et al.,2000), but their significance is problematic (Romer &Ro¨ tzler, 2001). Janousˇek et al. (2004) suggested anOrdovician granitic protolith with a model age <strong>of</strong>c. 450 Ma for the Variscan felsic granulites, supportedalso by radiometric U–Pb zircon ages <strong>of</strong> Kro¨ ner et al.(2000) <strong>and</strong> Friedl et al. (2004). The Rb–Sr biotite <strong>and</strong>muscovite cooling ages span between 330 <strong>and</strong> 310 Ma(Van Breemen et al., 1982; Svojtka et al., 2002) similarto 40 Ar– 39 Ar ages on amphibole, muscovite <strong>and</strong> biotite(Kosˇler et al., 1999). The exposed Moldanubian rocksreached the surface during the Permian, when theywere locally covered by undeformed sedimentaryrocks. Moldanubian Varied <strong>and</strong> Monotonous Groupsrecord metamorphic events from 355 ± 2 Ma (U–Pbon sphene, Wendt et al., 1993) or the 367 ± 19 Ma forthe Kaplice paragneiss (U–Pb on zircon, Kro¨ ner et al.,1988). A minimum age for the ductile deformationis restricted by syndeformational pegmatites that yieldan age <strong>of</strong> 331 ± 5 Ma (Rb–Sr on muscovite, VanBreemen et al., 1982).There is a range <strong>of</strong> syn- to post-tectonic granitoidsintruded into the Moldanubian rocks. The oldest calcalkalineintrusions <strong>of</strong> the Central Bohemian PlutonicComplex were syntectonically emplaced into uppercrustallevels during regional transpression from c. 354Ó 2010 Blackwell Publishing Ltd209


58 J. FRANĚK ET AL.to c. 346 Ma (Janousˇek et al., 2004) <strong>and</strong> extension at c.340 Ma (Ža´k et al., 2005a), followed by the intrusion<strong>of</strong> undeformed granitoids along the boundary betweenthis complex <strong>and</strong> the Moldanubian domain at c.337 Ma (e.g. Janousˇek & Gerdes, 2003). This corroboratesages <strong>of</strong> 338–314 Ma for intrusion into theMonotonous Group <strong>of</strong> mostly post-tectonic granitoids<strong>of</strong> the Moldanubian Pluton (e.g. Finger et al., 1997).In northern Bavaria, the Moldanubian rocks exhibityounger metamorphism associated with widespreadmigmatitization dated by U–Pb on monazite at c.326 Ma (Kalt et al., 1997) <strong>and</strong> c. 315 Ma (Schulzschmalschlager,1984). Syntectonic Bavarian granitoidsyield ages comparable to the MoldanubianPluton (335–322 Ma; e.g. Siebel et al., 2006).GEOLOGY AND STRUCTURAL GEOLOGY OFSOUTH BOHEMIAN MOLDANUBIAN DOMAINThis study focuses on a 130 · 30 km traverse throughthe whole South Bohemian Moldanubian domain fromthe Moldanubian Pluton in the SE to the CentralBohemian Plutonic Complex in the NW following the9HR seismic line (Fig. 4a). The SE part is characterizedby the occurrence <strong>of</strong> three large neighbouringbodies <strong>of</strong> felsic granulites (Blansky´ les, Krˇisˇtˇanov <strong>and</strong>Prachatice granulite massifs, see Fig. 5a). Discontinuousstripes <strong>of</strong> garnet or spinel peridotites <strong>and</strong> garnetpyroxenites are systematically found along the margins<strong>of</strong> all the granulite massifs, <strong>and</strong> the Blansky´ les granulitemassif also contains numerous large bodies <strong>of</strong>mantle-derived rocks in its interior.The adjacent Moldanubian domain consists <strong>of</strong> VariedGroup rocks to the SE <strong>and</strong> Monotonous Group rocks tothe NW. The individual massifs are separated by narrowzones <strong>of</strong> medium to high-grade paragneisses withaffinity to both Monotonous <strong>and</strong> Varied Groups. Thenorth-eastern margin <strong>of</strong> the Blansky´ les granulite massifis marked by a complex association <strong>of</strong> medium <strong>and</strong>high-grade rocks (Fig. 4a). Farther to the east all theunits become parallel to a major thrust zone defined bythe occurrence <strong>of</strong> eclogite bodies marking the ductilethrust <strong>of</strong> the Varied Group over the underlyingMonotonous Group (Rajlich et al., 1986; Vra´na &Sˇra´mek, 1999; Faryad et al., 2006). A zone <strong>of</strong> lowergrade muscovite-biotite paragneiss intruded by granitoids<strong>of</strong> the Moldanubian Pluton occurs in the easternmostextremity <strong>of</strong> the studied area. The MonotonousGroup in the north-western part <strong>of</strong> the traverse isintercalated with several large NE–SW elongated bodies<strong>of</strong> Varied Group <strong>and</strong> Gfo¨ hl orthogneiss (Fig. 4a).Earliest granulite facies fabrics S1 <strong>and</strong> S2The oldest identified fabrics (S1 <strong>and</strong> S2) are exceptionallywell preserved in an 8.4 · 2.5 km wide ellipticaldomain in the southern part <strong>of</strong> the Blansky´ lesgranulite massif (Figs 5a,b & 6a). Small relicts arefurther present throughout this massif <strong>and</strong> rarely alsoin the Prachatice granulite massif. In particular, in theBlansky´ les relict elliptical area the granulites exhibit apenetrative mylonitic foliation S2 (Fig. 5b) whichcontains relicts <strong>of</strong> S1 compositional layering, bothbearing a stable granulite facies mineral assemblage<strong>of</strong> Qtz + Kfs + Pl + Grt + Ky + Bt (Sˇtípska´ &Powell, 2005; Franěk et al., 2006). Mineral abbreviationsare after Kretz (1983).The onset <strong>of</strong> S2 fabric development is characterizedby pervasive recrystallization <strong>of</strong> the former coarsegrained S1 layered orthogneiss into a fine-grainedmosaic <strong>of</strong> plagioclase, K-feldspar <strong>and</strong> quartz, containinga small volume <strong>of</strong> former partial melt, nowindicated, for example, by cuspate shapes <strong>of</strong> theseminerals, distributed along feldspar boundaries(Franeˇk et al., 2010). The subsequent folding <strong>of</strong> S1 isaccompanied by progressive development <strong>of</strong> S2 axialplanecleavage <strong>and</strong> penetrative reworking <strong>of</strong> the S1fabric into the fine-grained, K-feldspar dominatedmatrix containing large quartz ribbons <strong>and</strong> biotiteflakes oriented into strong subhorizontal stretchinglineation (fig. 3e in Franěk et al., 2006). Plagioclasecoronas developed around kyanite <strong>and</strong> garnet in thematrix were interpreted as a product <strong>of</strong> reaction duringdecompression by Tajcˇmanova´ et al. (2007). The S2foliation from the Blansky´ les granulite massif strikesuniformly N–S showing variable dip <strong>of</strong> 50–90° to theW or less commonly to the E (Figs 4e, 5b & 7).Amphibolite facies D3 deformationIn all the Moldanubian rocks except the granulites, thefirst well-defined metamorphic fabric is a steepNE–SW trending amphibolite facies foliation(Fig. 6b,d). It is assigned as S3 due to a common<strong>structural</strong> concordance with a steep retrogressive S3fabric developed in the granulites.Because <strong>of</strong> strong later reworking, the S3 fabrics areusually well preserved only in competent lithologies.The only exceptions represent the southern part <strong>of</strong> theKaplice Zone muscovite paragneiss <strong>and</strong> the boundary<strong>of</strong> the Moldanubian <strong>and</strong> the Tepla´-Barr<strong>and</strong>i<strong>and</strong>omains where steep NW-dipping S3 is present. Rarerelicts <strong>of</strong> S3 in metasedimentary rocks throughout thestudy area exhibit the same attitude. The S3 foliationoccasionally bears a weak mineral lineation or corrugations,both shallowly plunging NE <strong>and</strong> SW. In thevicinity <strong>of</strong> the regionally folded granulite massifs, thegenerally stable orientation <strong>of</strong> the S3 foliation ismodified by large wavelength late F3 folds.The S3 is defined by preferred orientation <strong>of</strong> mica<strong>and</strong> elongated quartz–feldspar aggregates in the Bt +Qtz + Pl + Kfs ± Sil ± Grt ± Ms paragneiss <strong>of</strong>the Monotonous <strong>and</strong> Varied Groups <strong>and</strong> by preferredorientation <strong>of</strong> leucosomes in migmatitic paragneiss<strong>and</strong> the anatectic Gfo¨ hl gneiss. In the medium-gradeKaplice Zone paragneiss the S3 is defined by alignment<strong>of</strong> muscovite <strong>and</strong> biotite <strong>and</strong> segregation <strong>of</strong> quartzlenses with cordierite or aluminosilicates (Vra´na et al.,Ó 2010 Blackwell Publishing Ltd210


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 59(a)(b)(c)(d)(e)(f)Fig. 4. (a) Structural map <strong>of</strong> the study region. Structural data from granitoids adopted from Zˇa´k et al. (2005a). (b) Location <strong>of</strong>geological pr<strong>of</strong>iles across the granulites. All pr<strong>of</strong>iles are <strong>of</strong> the same scale <strong>and</strong> not exaggerated. (c) W–E view <strong>of</strong> Krˇisˇťanov granulitemassif. (d) N–S pr<strong>of</strong>ile across Prachatice granulite massif, Libı´n Zone <strong>and</strong> Krˇisˇtˇanov granulite massif. (e) NW–SE Section <strong>of</strong>Prachatice <strong>and</strong> Blansky´ les granulite massifs, roughly following the seismic reflection line 9HR. (f) N–S section <strong>of</strong> Blansky´ les granulitemassif. Part <strong>of</strong> pr<strong>of</strong>ile (e) <strong>and</strong> pr<strong>of</strong>ile (f) are modified after Franeˇk et al. (2006).Ó 2010 Blackwell Publishing Ltd211


60 J. FRANĚK ET AL.(a)(b)(c)(d)Fig. 5. (a) Structural map <strong>of</strong> the SE part <strong>of</strong> transect, which is dominated by granulites. The S2 fabrics are depicted in the inset (b)in a detail <strong>of</strong> the relict granulitic domain inside the Blansky´ les granulite massif. (c) Intensity <strong>of</strong> the L4 lineation in the Prachaticegranulite massif <strong>and</strong> surroundings interpolated from field data. (d) Steepness <strong>of</strong> the S4 foliation in the Prachatice granulite massif <strong>and</strong>surroundings interpolated from field data.1995; Vra´na & Ba´rtek, 2005). At the boundary <strong>of</strong> theMoldanubian <strong>and</strong> Tepla´-Barr<strong>and</strong>ian domains the S3 isdefined by a higher modal muscovite in conjunctionwith a decreasing intensity <strong>of</strong> partial melting.In the granulites, the S3 represents the dominantplanar fabric in the Blansky´ les <strong>and</strong> Krˇisˇtˇanov granulitemassifs, whereas in the Prachatice granulite massif,S3 is preserved only along its southern <strong>and</strong> easternmargins. The S3 results from transposition <strong>of</strong> S2 viaoutcrop-scale folding accompanied by development<strong>of</strong> S3 axial-plane foliation. The stable assemblage <strong>of</strong>Qtz + Kfs + Pl + Bt ± Sil ± Grt indicates syndeformationalamphibolite facies retrogression commonlyresulting in weak compositional layering.Syn-D3 partial melting occurred in the outer parts <strong>of</strong>the Blansky´ les <strong>and</strong> to some extent in the Krˇisˇťanovgranulite massif. At the transition from S2 to S3, theD3 microstructures still resemble the granulitic structure,suggesting that the onset <strong>of</strong> the D3 deformationoccurred at the granulite–amphibolite facies transition.The timing <strong>of</strong> D3 in the granulites is best bracketed bythe 337 ± 0.3 Ma crystallization age <strong>of</strong> a granulitefacies hyperpotassic dyke deformed by D3 (Aftalionet al., 1989; Sla´ma et al., 2008) <strong>and</strong> the crystallizationage <strong>of</strong> 340 ± 3 Ma for post-D4 cordierite-bearingleucosomes (Kro¨ ner et al., 2000).In granulites the S3 exhibits arcuate geometry(Figs 5a & 7) with moderate to steep dips (Kodym, 1972;Franeˇk et al., 2006). The arrangement <strong>of</strong> poles <strong>of</strong> S3along a great circle in each <strong>of</strong> the granulite massifssuggests a cylindrical fold geometry. Changes in attitude<strong>of</strong> S3 in all the granulite bodies appear as km-scaleflexures, or rarely as several km wide sequences <strong>of</strong> outcrop-scaleparasitic folds (Fig. 4b–f). Axial planes <strong>of</strong>these large folds in all three granulite bodies revealsimilar orientation (Figs 5a & 7). For the tight fold <strong>of</strong> thePrachatice granulite massif, we have constructed asubvertical NNW–SSE trending axial surface oriented262 ⁄ 87 (dip direction ⁄ dip in degrees) which is similar inorientation to the 246 ⁄ 89 axial planes <strong>of</strong> large-scale foldsdefined previously in the Blansky´ les granulite massifby Franěk et al. (2006). For the wide arcuate geometry<strong>of</strong> S3 in the Krˇisˇťanov granulite massif we are only ableto estimate a steep axial-plane dipping to the W.Ó 2010 Blackwell Publishing Ltd212


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 61(a)(b)(c)(d) (e) (f)Fig. 6. Field photographs <strong>of</strong> structures. (a) Development <strong>of</strong> S2 axial fabric during F2 passive folding in granulites. (b) Transposition<strong>of</strong> S2–S3 via mesoscopic similar folds in granulites. (c) Initial S4 in discrete shear zones affecting S3 in the Prachatice granulite massif.(d) Steep S3 only weakly affected by D4, lower grade Kaplice paragneiss in the SE <strong>of</strong> the transect. (e) S4 crenulation cleavage in amigmatized paragneiss, Varied Group in the NW <strong>of</strong> the transect. (f) S3 transposed into S4 via crenulations <strong>and</strong> shear b<strong>and</strong>s, whichare commonly filled with granitic leucosome. Orthogneiss in NW part <strong>of</strong> the transect.The biotite or sillimanite L3 mineral lineationplunges north to northwest, parallel to the hinges <strong>of</strong>these large folds in the Blansky´ les granulite massifexcept for the northern part <strong>of</strong> this body, where theyvary from south to east. The well-developed L3 lineationin the Prachatice granulite massif plunges shallowlysubparallel to the strike <strong>of</strong> the S3. In contrast,the L3 lineation in the Krˇisˇťanov granulite massif ispoorly developed except in the NW corner, where theplunge is generally to the SW (Fig. 8).In the metasedimentary rocks separating the granulitemassifs, the S3 fabrics are well preserved only inthe south, whereas towards the north <strong>and</strong> west onlyrare outcrop-scale relicts <strong>of</strong> S3 occur. Ubiquitous syndeformationalpartial melting <strong>of</strong> paragneisses resultedin stromatitic migmatitic layering parallel to the S3schistosity. The S3 in these metasedimentary rocksstrikes parallel to the S3 in neighbouring granulites; inthe Lhenice Zone the S3 forms a tight vertical N–Selongated fan-like pattern (Figs 4e & 5a) whereas inthe Libín Zone it dips steeply to the southwestunderneath the Krˇisˇťanov granulite massif (Figs 4d &5a). Rocks <strong>of</strong> the Lhenice Zone have an intense shallowlyN–S-plunging lineation defined by biotite alignment,whereas in the Libı´n Zone a L3 lineation is notdeveloped.Subhorizontal amphibolite facies fabric S4The most prominent fabric in rocks <strong>of</strong> the traverseis the S4 foliation that generally strikes NE–SWsubparallel to the Moldanubian–Tepla´-Barr<strong>and</strong>ianboundary, in general dipping gently to the NW. Onlyalong the southern edge <strong>of</strong> the Central BohemianPlutonic Complex <strong>and</strong> 10 km northwest <strong>of</strong> thePrachatice granulite massif does the S4 fabric dip tothe SE, forming large-scale open fold-like flexures inthe Monotonous <strong>and</strong> Varied Groups; a N–S to NE–SW stretching lineation on the S4 fabric is subparallelto the hinges <strong>of</strong> these flexures. Similar to the S3 fabric,in the vicinity <strong>of</strong> the granulite massifs, the S4 fabricexhibits significant perturbation as a result <strong>of</strong> deformationpartitioning <strong>and</strong> locally attains a steep attitudesubparallel to the S3 fabric (e.g. at the eastern margin<strong>of</strong> the Blansky´ les granulite massif).Partial melting associated with the S4 fabric in theMonotonous <strong>and</strong> Varied metasedimentary rocks(Fig. 6e,f) is <strong>of</strong> lower volume than that associated withÓ 2010 Blackwell Publishing Ltd213


62 J. FRANĚK ET AL.Fig. 7. Equal area lower hemisphere projections <strong>of</strong> S2, S3 <strong>and</strong> S4 fabrics in the granulite-dominated region <strong>of</strong> the transect contrastedagainst the same fabrics developed in the Moldanubian domain outside the granulite-dominated region.Fig. 8. A krigging interpolation to depictthe prevalence <strong>of</strong> particular fabrics inferredfrom field study <strong>of</strong> individual outcrops.Grey levels show the extent <strong>of</strong> reworkingby particular deformation phases (darkest =S2, lightest = post-D4 late NW-SE crenulationcleavage). Four generations <strong>of</strong>corresponding lineations are depicted.Ó 2010 Blackwell Publishing Ltd214


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 63the S3 fabric. In the granulites the S4 is marked bytransposition <strong>of</strong> the typical dm scale compositionallayering <strong>and</strong> syntectonic replacement <strong>of</strong> garnet bybiotite <strong>and</strong> sillimanite. The transition from S3 to S4proceeds via development <strong>of</strong> shallowly dipping S4-parallel shear zones, <strong>of</strong>ten accompanied by buckling <strong>of</strong>S3 at various scales. Occasional cm to dm scale intrafoliationpinch <strong>and</strong> swell structures with steep E–Wtrending tensional cracks filled by undeformed graniticleucosome develop in the S4 indicating verticalshortening <strong>and</strong> N–S-oriented stretching operating attemperatures around the granite solidus.The S4 intensity systematically increases from the SEto the NW towards the Central Bohemian PlutonicComplex. Here, the deformation culminates in highlynon-coaxial shearing manifested by a shear zoneseveral kilometres wide intruded by syntectonicgranitoids (Ža´k et al., 2005a) dated at 342–337 Ma,which corroborates a U–Pb age <strong>of</strong> 341 ± 3 Ma forzircon from leucosome patches cross-cutting the S4fabric <strong>of</strong> the Prachatice granulite massif (Kro¨ ner et al.,2000). The ambient P–T conditions in this regionduring formation <strong>of</strong> a flat fabric comparable with theS4 fabric <strong>of</strong> this study were determined as >720 °C at0.4 GPa by Pitra et al. (1999) <strong>and</strong> Scheuvens (1999).In the hangingwall paragneisses, at the Moldanubian–Tepla´-Barr<strong>and</strong>ian boundary, the D4 imprint rapidlydies out, documenting a block several kilometres wide<strong>of</strong> lower grade rocks that are inferred to have sunkbetween the Moldanubian <strong>and</strong> Tepla´-Barr<strong>and</strong>i<strong>and</strong>omains.In the east, the intensity <strong>of</strong> the D4 deformationgradually decreases so that only the western part <strong>of</strong> thegranulite bodies shows relatively strong reworking bythe flat S4 fabrics at 765 °C <strong>and</strong> 0.76 GPa (Fig. 6c;Verner et al., 2008). Whereas most <strong>of</strong> the Prachaticegranulite massif is strongly reworked by the subhorizontalS4 foliation <strong>and</strong> weak N–S lineation (Figs 5c,d& 8), the Krˇisˇťanov granulite massif exhibits onlymoderate reworking by a shallow WNW-dipping S4foliation (Verner et al., 2008) <strong>and</strong> the Blansky´ lesgranulite massif to the east is almost unaffected by D4(Fig. 4b–f). In the Lhenice Zone the S4 forms an opensynform with horizontal N–S axis, <strong>and</strong> the rocks exhibita strongly developed, shallow N–S-plungingstretching lineation. Within the Libı´n Zone the S4fabric dips moderately to the SW below the Krˇisˇtˇanovgranulite massif; a weak lineation <strong>of</strong> variable attitude ispresent. North <strong>of</strong> the Blansky´ les granulite massif theS4 dips shallowly to the SSW. Further southeast the S4is variable <strong>and</strong> generally has a steeper attitude indicatingthat the D4 intensity decreases. Finally, in thesouth <strong>of</strong> the medium-grade paragneiss <strong>of</strong> the KapliceZone the S4 is only weakly developed.NW–SE trending steep cleavage frontThe SW part <strong>of</strong> the traverse is affected by late ductiledeformation resulting in development <strong>of</strong> a crenulationcleavage, which develops into a new penetrative foliationdipping moderately to steeply to the NE (Fig. 4a).The intensity <strong>of</strong> this fabric generally increases southwestward<strong>and</strong> it becomes the dominant fabric in theBavarian region characterized by coeval dextral shearzones (e.g. Behrmann & Tanner, 1997; Finger et al.,2007). There is no expression <strong>of</strong> such structures insidethe granulite massifs implying that the geometry <strong>and</strong>internal structure <strong>of</strong> the granulites were not affected bythis late deformational event.REFLECTION SEISMIC PROFILEMost <strong>of</strong> the traverse is parallel to a deep seismicreflection pr<strong>of</strong>ile (9HR) which was shot in a NW–SEdirection through western half <strong>of</strong> the Bohemian Massifusing explosive sources. The seismic pr<strong>of</strong>ile, which isonly partly published (Tomek et al., 1997), extendsacross the region between the eastern boundary <strong>of</strong> theSaxothuringian domain in the NW to the eastern edge<strong>of</strong> the Blansky´ les granulite massif in the SE. Thisstudy examines the SE half <strong>of</strong> the pr<strong>of</strong>ile (Figs 9 & 10),which <strong>of</strong>fers a good quality record from 2 km belowthe surface to the Moho at 40 km depth, exhibitingin places also reflections in the upper mantle.Characterization <strong>of</strong> reflection seismic dataFor most <strong>of</strong> the traverse the seismic pr<strong>of</strong>ile runsperpendicular to the strike <strong>of</strong> rock fabrics, whichsuggests that the dips <strong>of</strong> the reflectors are close totrue dips, at least in the upper crust. The line drawing(Fig. 10) based on the migrated seismic record revealsa striking difference between the seismic properties <strong>of</strong>the Tepla´-Barr<strong>and</strong>ian <strong>and</strong> the Moldanubian crust.Whereas the Moldanubian crust exhibits severalseismically different domains, the highly reflectiveTepla´-Barr<strong>and</strong>ian upper crust shows one single set <strong>of</strong>parallel <strong>and</strong> very strong reflection packages (B1–B6)dipping to the SE. These packages are interrupted atdepth along a NW-dipping virtual line, below whichthe Tepla´-Barr<strong>and</strong>ian lower crust shows very poorreflectivity, being constrained by subhorizontal, butdiscontinuous Moho reflections (M1) at 11 s TWT(two-way time). The Tepla´-Barr<strong>and</strong>ian–Moldanubianboundary zone, intruded by granitoids <strong>of</strong> the CentralBohemian Plutonic Complex, shows a sharp decreasein reflectivity <strong>of</strong> the whole crust in a 10 km widecolumn.The Moldanubian crust shows contrasting seismiccharacteristics in the eastern part, which is dominatedby granulites, <strong>and</strong> the western part, which is devoid <strong>of</strong>these rocks. The western part is characterized bystronger reflection packages (K1 <strong>and</strong> K2, V1–V4) <strong>and</strong>anastomosing reflections down to 5 s TWT, whichresemble pinch <strong>and</strong> swell structures <strong>of</strong> 20 km wavelength.The upper crust in the east <strong>of</strong> the seismic pr<strong>of</strong>ileis cross-cut by a strong <strong>and</strong> unusually straight reflectionpackage (G2) dipping moderately to the SE.Ó 2010 Blackwell Publishing Ltd215


64 J. FRANĚK ET AL.Fig. 9. The 9HR seismic pr<strong>of</strong>ile showing the migrated raw data. Vertical scale approximately equals the horizontal scale. Position <strong>of</strong> the pr<strong>of</strong>ile is shown in Fig. 4a, it coincideswith section 3 <strong>of</strong> the gravity model shown in Fig. 11.Ó 2010 Blackwell Publishing Ltd216


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 65(a)(b)Fig. 10. The 9HR seismic pr<strong>of</strong>ile showing line drawing <strong>of</strong> the migrated reflections. Position <strong>of</strong> the pr<strong>of</strong>ile is shown in Fig. 4a. (a) Upper crust with <strong>structural</strong> pr<strong>of</strong>ile that is basedonly on field <strong>structural</strong> study. Surface extent <strong>of</strong> major lithological units is plotted above. CBPC refers to Central Bohemian Plutonic Complex, TBU to Tepla´-Barr<strong>and</strong>ianUnit. (b) Seismic reflections with package description, thickness <strong>and</strong> darkness <strong>of</strong> lines express intensity <strong>of</strong> reflections. The reflections in (b) west <strong>of</strong> the granulites fit very well withthe independent geological interpretation down to 5 km depth. Vertical scale approximately equals the horizontal scale.Ó 2010 Blackwell Publishing Ltd217


66 J. FRANĚK ET AL.At the top, the package terminates abruptly at 1 sTWT below the western edge <strong>of</strong> the Prachatice granulitemassif, probably on a local N–S fault. Thestrongest reflections in the pr<strong>of</strong>ile appear below thePrachatice granulite massif; they diminish to the SE at5 s TWT, below the centre <strong>of</strong> the Blansky´ les granulitemassif. A parallel weaker package (G1) <strong>of</strong>12 km length appears attached to the top <strong>of</strong> thispackage <strong>of</strong> strong reflections. Except for these features,the crust below the granulites exhibits almost noreflectivity <strong>and</strong> there is only a weakly defined Moho(M4) at 12–13 s TWT. The reflections along the SEedge <strong>of</strong> the seismic pr<strong>of</strong>ile are probably just artefacts <strong>of</strong>the migration procedure. This lack <strong>of</strong> reflections contrastswith a package <strong>of</strong> strong NW-dipping mantlereflections below 12s TWT (Cˇ. Tomek, personal communication;not shown in Fig. 10). The western part <strong>of</strong>the Moldanubian middle to lower crust exhibits uniformlydistributed horizontal reflections designated asLC, which appear in a lens-shaped domain60 · 15 km in size. There are several packages <strong>of</strong>strong reflections (Z1–Z4) in the underlying lowercrust that show apparent dips <strong>of</strong> 30–40° to the SE <strong>and</strong>continue through discontinuous MOHO reflections(M2 <strong>and</strong> M3) into the uppermost mantle.Interpretation <strong>of</strong> reflection seismic dataIn the high-grade Moldanubian domain, which recordspolyphase deformation, the reflection packages mayrepresent lithological layering, penetrative foliation,faults or boundaries <strong>of</strong> intrusive bodies. We try toovercome this ambiguity by careful comparison withthe surface extent <strong>of</strong> major lithological units <strong>and</strong> withour <strong>structural</strong> interpretations.The upper-crustal reflection packages (B1–B6) in theTepla´-Barr<strong>and</strong>ian Unit correspond to Proterozoic lowgradesequence <strong>of</strong> siltstones interlayered with basaltsexposed on the surface that show SE-dipping schistosity.Termination <strong>of</strong> these packages at the margin <strong>of</strong>the Tepla´-Barr<strong>and</strong>ian Unit may document sidewardintrusions <strong>of</strong> granitoid bodies related to the CentralBohemian Plutonic Complex, or reworking by laterVariscan deformation. The low reflectivity in thecrustal column directly below the Tepla´-Barr<strong>and</strong>ian–Moldanubian boundary can be best explained by theoccurrence <strong>of</strong> Central Bohemian Plutonic Complexgranitoid intrusions at depth below the MonotonousGroup exposed at the surface.The anastomosing reflections in the upper crust <strong>of</strong>the western Moldanubian domain correlate well withthe geometry <strong>of</strong> the S4 fabrics as determined by ourfield research (Fig. 10a). This conformity suggests thatall the anastomosing reflections can be related to thesubsurface continuation <strong>of</strong> the S4 fabrics. The V1–V4reflection packages correlate well with surface exposures<strong>of</strong> km-scale lenses <strong>of</strong> Varied Group rockssuggesting that the reflectors in this case representinterlayering <strong>of</strong> paragneisses with marbles, amphibolites<strong>and</strong> other lithologies, which are oriented parallelto the S4 fabric. The K1 <strong>and</strong> K2 reflection packagesdocument probable subsurface occurrence <strong>of</strong> variedlithological intercalations at the western side <strong>of</strong> theMoldanubian domain. The nature <strong>of</strong> the straightreflection package (G2) dipping below the granulitemassifs remains ambiguous, because the observed highamplitude <strong>and</strong> length <strong>of</strong> reflections may be causedeither by lithological layering similar to that <strong>of</strong> theVaried Group or by a thick zone <strong>of</strong> intensive mylonitization.Based on the single seismic section, theorientation <strong>of</strong> the G2 reflectors can vary between asteeper dip to the NE through medium dips to the SEto steeper dips to the SW. Such orientations cannot bedirectly correlated to any structure mapped on thesurface. The straight geometry <strong>of</strong> the G2 packagesuggests that the corresponding reflectors were notsignificantly affected by the D4 deformation, or thatthey may represent an anomalously oriented D4structure. The low reflectivity in the crustal columnbelow the granulite massifs may be caused either bylack <strong>of</strong> reflection inducing inhomogeneities or extremesteepness <strong>of</strong> reflectors, which would not be detectableby seismic survey. The low energy <strong>of</strong> the shots cannotbe responsible for the lack <strong>of</strong> crustal reflections,because the much deeper mantle reflections have beenrecorded below this seismically transparent zone. Therest <strong>of</strong> the Moldanubian crystalline crust exhibits significantreflectivity, suggesting that the Moldanubiancrust in general contains abundant reflectors thatwould be recorded if they had a suitable attitude.Below the granulites the reflectors may have beenpartially destroyed by the rise <strong>of</strong> these HP rocks, but itis argued that this crustal section likely still containsnumerous reflectors similar to the rest <strong>of</strong> the Moldanubi<strong>and</strong>omain, but the attitude <strong>of</strong> these reflectors istoo steep to be detected by the seismic method. Incombination with the dominance <strong>of</strong> steep fabrics in theexposed granulites, the lack <strong>of</strong> reflections serves asindirect evidence for the dominance <strong>of</strong> steep foliationsin the seismically transparent zone throughout thecrust below the granulite massifs. The homogenouslydistributed reflections in the middle to lower crust,designated as LC in Fig. 10b, do not reach the surfacealong the 9HR line. Such middle to lower crustalreflections are classically, but without direct evidence,ascribed to subhorizontal intrusions. These horizontalreflections cannot represent the deeper continuation <strong>of</strong>the S4 fabric, because <strong>of</strong> the sharp transition zone withthe anastomosing upper-crustal S4-related reflections<strong>and</strong> the general dip <strong>of</strong> the S4 to the NW. The LCreflections are unlikely to represent a fabric older thanS4, because the D3 phase <strong>of</strong> horizontal shorteningleads to development <strong>of</strong> steep NE–SW trending mostlymigmatitic fabrics along the traverse. The abrupttransition in seismic fabric may also mark a rheologicaltransition between the Variscan upper <strong>and</strong> middle tolower crust, with the age <strong>of</strong> the LC reflectionsremaining unknown.Ó 2010 Blackwell Publishing Ltd218


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 67(a)(c)(b)(d)Fig. 11. (a, b) Three-dimensional gravity model <strong>of</strong> the Prachatice granulite massif, comparison <strong>of</strong> measured with modelled gravityfield. (c) Two most significant model sections through the granulite massif. Curves above depict the values <strong>of</strong> the measured Bougueranomalies (thick dashed) <strong>and</strong> those modelled in three-dimensional (solid). The thin dashed line shows the two-dimensional gravityeffect <strong>of</strong> the individual sections. Only 20 km <strong>of</strong> the horizontal extent <strong>of</strong> each section relevant to the Prachatice granulite massif areshown. Section 3, which coincides with the reflection seismic pr<strong>of</strong>ile, contains line drawing <strong>of</strong> corresponding reflections. White values insection 4 refer to final densities used in the gravimetric model. (d) Map <strong>of</strong> Bouguer gravity anomalies (modified after J. Sˇvancara,unpublished data) with outlines <strong>of</strong> geological units, location <strong>of</strong> gravity model pr<strong>of</strong>iles <strong>and</strong> the 9HR seismic line.The lower crustal reflections Z1–Z4, which areconcentrated in a triangular region in the NW part <strong>of</strong>the Moldanubian domain, dip moderately to the SE<strong>and</strong> their lower tips appear to cut through the Mohointo the uppermost mantle. Such characteristics arecomparable with seismic images <strong>of</strong> subduction sutures,Ó 2010 Blackwell Publishing Ltd219


68 J. FRANĚK ET AL.Fig. 12. Interpretative geochronological sketch <strong>of</strong> Variscan evolution in the Bohemian Massif, as it is related to the traverse.but the speculation that the observed seismic structuremay represent the remnant <strong>of</strong> a suture zone betweenan unknown plate <strong>and</strong> the overlying Moldanubi<strong>and</strong>omain is very poorly constrained. The discontinuous<strong>and</strong> weakly defined Moho reflections (M1–M4) belowthe Moldanubian crust document systematic deepening<strong>of</strong> the Moho from 35 km in the NW to 42 kmin the SE. This deepening is in agreement with theMoho geometry defined by the refraction study <strong>of</strong>Hrubcova´ et al. (2005) <strong>and</strong> by the passive seismologyexperiments <strong>of</strong> Plomerova´ et al. (2005).GRAVITY MODELLING AND VERTICAL EXTENTOF THE FELSIC GRANULITESTo decipher the vertical extent <strong>of</strong> the granulite massifs,an interpolated grid <strong>of</strong> Bouguer anomalies coveringmost <strong>of</strong> the traverse (provided by J. Sˇvancara,unpublished data; Fig. 11d) has been examined. Thearea is dominated by large gravity lows related togranitoids <strong>of</strong> the Moldanubian Pluton in the SE or tounknown sources in the centre <strong>and</strong> NW parts. Positiveanomalies are all smaller scale <strong>and</strong> appear in thesouthern part <strong>of</strong> the area.Forward gravity <strong>modelling</strong> in three-dimensions hasbeen undertaken using the IGMAS s<strong>of</strong>tware (Schmidt& Go¨ tze, 1999), where geological bodies are representedby polyhedrons triangulated between manuallydrawn sections. The only suitable body for such anapproach is the Prachatice granulite massif, the surfaceextent <strong>of</strong> which correlates with a pronounced circularnegative anomaly. Its lithological homogeneity <strong>and</strong> thesmoothness <strong>of</strong> the gravity anomaly suggest that nosignificantly denser bodies (e.g. ultrabasites) areinvolved in this massif. The Krˇisˇťanov granulite massifinduces only a weak negative anomaly possiblyaffected by neighbouring intrusion <strong>of</strong> denser K–Mgsyenites. The Blansky´ les granulite massif partiallycorrelates with a positive anomaly, despite the density<strong>of</strong> the felsic granulites being lower than the surroundingmetasedimentary rocks (Appendix S2),probably as a result <strong>of</strong> abundance <strong>of</strong> ultrabasic bodiesÓ 2010 Blackwell Publishing Ltd220


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 69(a) (b)(c)(d)Fig. 13. Mid-crustal evolution <strong>of</strong> the region focused on detailed behaviour during the D3 <strong>and</strong> D4 phases. (a) Amplification <strong>of</strong> the complex granulite bulge under dextraltranspression. The driving force is transferred probably from the Saxothuringian subduction <strong>and</strong> collision. KG, PG <strong>and</strong> BLG refer to the ancestors <strong>of</strong> the Krˇisˇťanov, Prachatice<strong>and</strong> Blansky´ les granulite massifs, TBU marks the Tepla´-Barr<strong>and</strong>ian Unit <strong>and</strong> CBPC the Central Bohemian Plutonic Complex. (b) Hardening <strong>of</strong> felsic granulites caused by coolingresults in amplification <strong>of</strong> several F3 regional-scale folds due to the dextral shear. (c) Final stage <strong>of</strong> fold development accompanied by intrusion <strong>of</strong> Mg-K rich syenites. (d)Immediate change <strong>of</strong> regional kinematic regime to vertical flattening. Pervasive development <strong>of</strong> flat S4 fabrics in the NW part <strong>of</strong> the traverse is combined with normal movement <strong>of</strong>the Tepla´-Barr<strong>and</strong>ian block in the NW partially along the flat S4 foliation.Ó 2010 Blackwell Publishing Ltd221


70 J. FRANĚK ET AL.<strong>and</strong> underlying dense Varied Group rocks. Theanomalies involved in our gravity model have smallareal extent <strong>and</strong> exhibit high gradients at their edges,implying that all <strong>of</strong> them are caused by upper-crustalheterogeneities. Removal <strong>of</strong> regional long-wavelengthgravity components caused by deeper heterogeneities isthen unnecessary <strong>and</strong> the original values <strong>of</strong> theBouguer anomalies have been used in the <strong>modelling</strong>that follows. For the same reason, the model concernsonly upper-crustal bodies down to 10 km depth.The geometry <strong>of</strong> the Prachatice granulite massif hasbeen approximated in six vertical sections 40 km long<strong>and</strong> 10 km deep, trending NW–SE parallel to theseismic reflection line 9HR (Fig. 11a–c). Unlike thetwo-dimensional gravity model <strong>of</strong> Vra´na & Sˇra´mek(1999), the three-dimensional approach was chosenbecause the relationship <strong>of</strong> this granulite massif with itssurroundings suggests a complex non-cylindricalgeometry <strong>of</strong> structures on the scale <strong>of</strong> our model <strong>and</strong> acircular shape <strong>of</strong> the anomaly also means that a twodimensionalapproach is unsuitable. The model isconstrained by the surface extent <strong>of</strong> the geologicalunits, <strong>structural</strong> data extrapolated to 1–2 km depth<strong>and</strong> the reflection seismic line that cross-cuts thePrachatice granulite massif <strong>and</strong> the centre <strong>of</strong> the negativeanomaly. The database <strong>of</strong> rock densities(Appendix S2) from Hana´k et al. (unpublished data),Chlupa´cˇova´ et al. (unpublished data) <strong>and</strong> Blizˇkovsky´et al. (unpublished data) provided starting guesses,which were carefully adjusted when the constraintsdiscussed earlier in the article restricted changes <strong>of</strong>geometry for the modelled bodies. This is a commontechnique because rock densities vary in space, e.g. theVaried Group in surface section exhibits significantlithological change on a kilometre scale. The model wasdesigned to be as simple as possible to avoid ambiguoussolutions, but its simplicity naturally causes divergence<strong>of</strong> the modelled gravity values from the measured valuesat the model edges. In our model, the Prachaticegranulite massif reaches a maximum depth <strong>of</strong> 5.7 km,having a thick lensoidal shape. It cannot extend belowthe strong reflection package, because such a package<strong>of</strong> reflectors could not be expected in a lithologicallyhomogenous granulite massif. It also cannot form athinner lens, because then an unreasonably low densityfor the felsic granulite would be needed to produce themeasured gravity low. A spectacular detail is the cusp<strong>of</strong> high-density Varied Group rocks introduced into thegranulite massif from below in the hinge region <strong>of</strong> thelarge fold (e.g. at 15 km in section 3 in Fig. 11c).Inferring this structure is the only reasonable way toproduce the observed cranked gravity curves. It isprobably <strong>structural</strong>ly analogous to a cusp at the SWedge <strong>of</strong> the Blansky´ les granulite massif, where a sheet<strong>of</strong> amphibolites with ultrabasites penetrates deeply insidethe granulite massif along the large fold axialplane. Our results differ from the two-dimensionalgravity model <strong>of</strong> the Prachatice granulite massif publishedby Vra´na & Sˇra´mek (1999) who expected arectangular rather than lensoidal shape <strong>of</strong> the massif,reaching to 9 km depth. The discrepancy is causedpartially by the different rock densities used <strong>and</strong>partially by the two-dimensional approach. It is arguedthat our model approximates better the reality as aresult <strong>of</strong> calculation in three-dimensions as wellas using better <strong>structural</strong> constraints from the nearsurfacegeology.DISCUSSION – KINEMATIC AND MECHANICALSIGNIFICANCE OF THE DEFORMATION FABRICSTectonic significance <strong>of</strong> granulite facies fabricsThe granulite facies S1–S2 foliations represent aunique example <strong>of</strong> lower crustal fabrics in respect <strong>of</strong>the whole Bohemian Massif. The decompressional P–Tpath indicates that the S2 fabric originated duringexhumation. The highly discordant relationshipbetween the S2 granulite facies fabric <strong>and</strong> the S3amphibolite facies foliation is the most important<strong>structural</strong> observation from the South Bohemiangranulites (Franeˇk et al., 2006). The HP granulitesrecord the early deformation, which are not recordedin rocks <strong>of</strong> either the middle- or the upper-crustalunits. Therefore, there is a possibility that the granulitefacies D2 mylonitization is older <strong>and</strong> occurred in acrustal unit that was geographically remote comparedto the Lower Carboniferous D3–D4 history recordedin lower-, middle- <strong>and</strong> upper-crustal units (Fig. 12).Horizontal shortening <strong>and</strong> development <strong>of</strong> the regionalvertical fabricThe distribution <strong>of</strong> the steep S3 foliation in thegranulite massifs shows important variations, whichcan be interpreted as a result <strong>of</strong> large-scale folding <strong>of</strong>granulite massifs with steep N–S axial planes (Franěket al., 2006). The granulite bodies were morecompetent than the surrounding metasedimentaryrocks due to cooling after exhumation at later stages<strong>of</strong> D3, which enabled fold amplification. The differencebetween the asymmetrically folded granulite S3fabrics <strong>and</strong> the straight NE–SW S3 fabric homogeneouslydeveloped in middle-crustal host rocks indicatesa non-coaxial dextral shear operating duringD3 in both the middle- <strong>and</strong> lower crustal units. Thesteep attitude <strong>of</strong> N–S axial planes rules out thepossibility that the S3 folding proceeded during thedevelopment <strong>of</strong> the flat S4 foliation. The axial planeslie close to the plane <strong>of</strong> maximum flattening <strong>of</strong> theinstantaneous D3 strain ellipsoid, while the generalNE–SW trend <strong>of</strong> the S3 fabric reflects the finitestrain orientation.When the Z-shaped Blansky´ les, the symmetricalKrˇisˇťanov <strong>and</strong> the V-shaped Prachatice granulitemassifs are unfolded to yield originally NE–SW strikingsheets, then the S3 generally dips to the west in theBlansky´ les, subvertically in the Krˇisˇtˇanov <strong>and</strong> to theÓ 2010 Blackwell Publishing Ltd222


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 71(a)(b)(c)(d)Fig. 14. Geotectonic model <strong>of</strong> granulite lower crustal exhumation with subsequent stacking <strong>and</strong> subhorizontal mid-crustal flowdepicted along a two-dimensional NW–SE lithospheric-scale pr<strong>of</strong>ile. (a) Hypothetical formation <strong>of</strong> granulites, which are suggested tohave formed during SE-directed subduction leading to intermixing <strong>of</strong> distinct crustal levels, probably <strong>of</strong> Late Devonian age (based on<strong>numerical</strong> models <strong>of</strong>, e.g. Gerya & Stockhert, 2006). (b) Early Carboniferous subduction stage with intensive arc magmatism <strong>and</strong>shortening in the arc <strong>and</strong> back-arc domain. (c) Prolonged shortening leads to extrusion <strong>of</strong> the lower crustal rocks from below the arcregion, while the fore-arc block slides to the NW from the amplifying extrusion. (d) Detail <strong>of</strong> the (c) stage focused on the traversestudied for this paper. Abbreviations used: UC, upper crust (above root); MC, middle crust; LC, lower crust; CC, continental crust(Armorica); EC, eclogitized crust (Armorica); UM, upper mantle; Ast, asthenosphere; PG, Prachatice granulite massif; BLG, Blansky´les granulite massif.east in the Prachatice granulite massif. Such spatialvariations in dip may be interpreted to reflect a20 km wide positive subvertical fan-like structurestriking NE–SW (Figs 13 & 14d). The distribution <strong>of</strong>serpentinite stripes inside the Blansky´ les granulitemassif accompanied sometimes by Varied Groupmetasedimentary rocks was interpreted by Franeˇket al. (2006) as a result <strong>of</strong> isoclinal folding <strong>of</strong> a granulitesheet <strong>and</strong> the host rocks accompanied by imbricationduring early D3. This observation indicatesoverall isoclinal folding <strong>of</strong> lower- <strong>and</strong> mid-crustal unitsthat dismembered a coherent mass <strong>of</strong> exhuming granuliteinto several massifs.The original NE–SW trending positive fan-likestructure along the boundary between the Varied <strong>and</strong>Monotonous Groups can be interpreted as the core <strong>of</strong>crustal-scale vertical extrusion <strong>of</strong> orogenic lower crustover medium-grade rocks to the NW <strong>and</strong> SE (Figs 13& 14). The fan-like geometry may result either fromdevelopment <strong>of</strong> a ductile positive flower structureduring extrusion or due to subsequent flattening <strong>and</strong>ductile thinning, that is responsible for the followingD4 subhorizontal flow.The D3 vertical fabric in the Moldanubian domain isgeometrically <strong>and</strong> kinematically consistent with thefabrics that are developed in the arc-related CentralÓ 2010 Blackwell Publishing Ltd223


72 J. FRANĚK ET AL.Bohemian Plutonic Complex as well as in the adjacentsupracrustal Tepla´-Barr<strong>and</strong>ian domain. The steepNE–SW trending fabrics preserved in slates <strong>and</strong> volcanicrocks <strong>of</strong> the Neoproterozoic unit as well as insyntectonic calc-alkaline (354–346 Ma) intrusions areconsistent with dextral transpression <strong>and</strong> coeval withsyntectonic contact metamorphism at 349–341 Ma(Scheuvens & Zulauf, 2000; Zˇa´k et al., 2005a,b). Thespatial distribution <strong>of</strong> fabrics <strong>and</strong> metamorphic rocksdocumenting D3 deformation reflects a significantexhumation event <strong>of</strong> HP rocks up to 0.7 GPa. Deeperparts <strong>of</strong> the subvertical granulite fan-like structureare expressed in the seismic section as a low reflectivityregion below the granulite massifs, which indicates theprobable existence <strong>of</strong> steep extrusion-related S3 fabricsdown to the Moho <strong>and</strong> marks the exhumation path <strong>of</strong>the granulite-dominated belt. Thus, the present crustalstructure preserves the geometry <strong>of</strong> the subverticalextrusion channel related to the flow <strong>of</strong> HP rocks fromthe base <strong>of</strong> the crust. The existence <strong>of</strong> other granulitebodies deeper in this channel is indirectly supportedby gravity <strong>modelling</strong> suggesting a significant thicknessfor individual granulite bodies (Fig. 11c). In conclusion,the D3 deformation reflects mechanical couplingamong the upper, middle <strong>and</strong> lower levels <strong>of</strong> theorogenic crust approximately at 354–342 Ma.Mid-crustal horizontal flow <strong>and</strong> crustal-scale detachmentThe good correlation <strong>of</strong> the S4 fabric with the subsurfaceseismic reflections, especially in the NW part <strong>of</strong>the traverse (Fig. 10a), implies that S4 extends at leastto 10 km depth. The surface fabric trajectories combinedwith the anastomosing pattern <strong>of</strong> seismicreflectors indicate that the large-scale geometry can beinterpreted as a set <strong>of</strong> 10–20 km wide pinch <strong>and</strong> swellstructures with subhorizontal NE–SW axes <strong>and</strong> shallowlyNW-dipping median planes. A characteristicfeature is the systematic decrease <strong>of</strong> D4 intensity fromthe northwest to the southeast so that the central parts<strong>of</strong> the granulite massifs are unaffected by this deformationevent, which is in accord with very limitedflattening <strong>of</strong> the Prachatice granulite massif inferredfrom the gravity <strong>modelling</strong>. This tectonometamorphicpattern can be interpreted as a result <strong>of</strong> a large-scaledetachment zone weakened by partial melting, alongwhich the supra-crustal Tepla´-Barr<strong>and</strong>ian Unit sliddown to the W–NW. Scheuvens & Zulauf (2000) <strong>and</strong>Do¨ rr & Zulauf (2008) suggested that sliding along theCentral Bohemian Shear Zone, supported by limiteddiapiric ascent <strong>of</strong> migmatites, occurred at c. 340 Maunder dextral transtension. We suggest that a notableamount <strong>of</strong> vertical movement was achieved along S4planes in the Moldanubian migmatites, based on theextreme intensity <strong>of</strong> the S4 development <strong>and</strong> correlationwith the results <strong>of</strong> Zˇa´k et al. (2005a). The verticaldifferential motion at the Moldanubian–Tepla´-Barr<strong>and</strong>ianboundary is estimated to 10–15 km, basedon data <strong>of</strong> Scheuvens & Zulauf (2000). The flatgeometry indicates also a component <strong>of</strong> lateral ductiletransport, which is consistent with <strong>numerical</strong> models <strong>of</strong>late evolution in the core <strong>of</strong> hot orogens (e.g. Beaumontet al., 2006; Jamieson et al., 2007), but difficultto quantify in the Moldanubian case.D4 fabric maps (Figs 5c,d & 8) show that rheologicallyweaker host gneisses exhibit a <strong>structural</strong> patternconsistent with viscous flow around granulite bodies,which behaved as rigid ellipsoidal objects (e.g. in thecentral part <strong>of</strong> the Prachatice granulite massif, the S4exhibits a flat attitude interpreted as the apical part <strong>of</strong>an ellipsoidal crustal-scale granulite boudin). At themargins <strong>of</strong> this massif, the steeper <strong>and</strong> in the east alsosynformal S4 fabrics represent 2 km wide neck zones.The Varied <strong>and</strong> Monotonous Group rocks surroundingthe granulites are devoid <strong>of</strong> (U)HP relicts. Thus,the HP granulite boudins do not represent relicts <strong>of</strong> anoriginally coherent HP metamorphic unit where theVaried <strong>and</strong> Monotonous Group rocks would be completelyretrogressed, unlike other orogenic root systems(e.g. Engvik & Andersen, 2000). In conclusion, the S4fabric corresponds to zone <strong>of</strong> coaxial vertical shortening,with the exception <strong>of</strong> a narrow highly noncoaxialzone at the Moldanubian–Central BohemianPlutonic Complex boundary in the NW. Here, the D4deformation is partially responsible for decoupling<strong>of</strong> the Neoproterozoic upper-crustal lid from theMoldanubian middle crust. Such deformation partitioningis <strong>of</strong> common occurrence across large-scalenormal shear zones related to exhumation (e.g. Lawet al., 1994; Little et al., 1994). The subhorizontal flowis responsible for vertical exhumation from 0.7–0.8GPa to 0.3–0.4 GPa at 342–337 Ma.TECTONIC MODELDevonian subduction <strong>of</strong> continental crustIn addition to the Carboniferous events, the followingevidence supports an earlier independent Variscantectonic regime throughout the N–NW BohemianMassif in the Devonian. Relicts <strong>of</strong> a mid-Devonianoceanic subduction zone, e.g. the Mu¨ nchberg Massif,occur in the hangingwall <strong>of</strong> the main continentalSaxothuringian subduction zone (e.g. OÕBrien, 1997;Konopa´sek & Schulmann, 2005). Also, there are felsicgranulites <strong>of</strong> Devonian age with a granulite facieslayering reported from several places in the BohemianMassif (e.g. Sowie Go´ry Mountains, OÕBrien et al.,1997). Finally, there is a range <strong>of</strong> Devonian zirconU–Pb <strong>and</strong> Pb–Pb ages from granulites (Wendt et al.,1994; Schulmann et al., 2005).Based on this evidence, it is suggested that thegranulite facies rocks in the Moldanubian domainoriginated during a subduction event prior to theSaxothuringian continental subduction, which culminatedin the Carboniferous collisional event at c.340 Ma. The S1–S2 fabrics, which are discordant tothe overall NE–SW Moldanubian trend, could haveÓ 2010 Blackwell Publishing Ltd224


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 73developed during such an older episode, possibly inrelation to the early arc history. This model is stronglysupported by the <strong>numerical</strong> simulations <strong>of</strong> Gerya &Stockhert (2006), who proposed an origin for HPgranulites in a lithosphere-scale subduction wedge. Inthis model, the rocks <strong>of</strong> the footwall plate are draggedto depths corresponding to 2.0–2.5 GPa, exhumedbackwards by a return flow <strong>and</strong> accreted to the base <strong>of</strong>the hangingwall crust (Fig. 14a). Other possible modelsinclude those <strong>of</strong> Gerya et al. (2008), Warren et al.(2008) <strong>and</strong> Beaumont et al. (2009) for buoyancy-drivenexhumation <strong>of</strong> weakened continental rocks in a subductionchannel during the early stages <strong>of</strong> continentalcollision. Janousˇek & Holub (2007) suggested a similarearly evolution <strong>and</strong> pointed out the geochemicalaffinity <strong>of</strong> the felsic granulites with numerous pre-Variscan granitoids <strong>of</strong> the Saxothuringian domain, i.e.in the footwall <strong>of</strong> the Saxothuringian subduction. Weinfer that during the D2 deformation the granuliteswere partially exhumed in such a tectonic setting fromthe peak pressure <strong>of</strong> 1.8–2.0 GPa <strong>and</strong> accreted to thebase <strong>of</strong> the hangingwall lower crust (tentativelydepicted in Fig. 14a), because the following D3 deformationcommenced at lower granulite facies conditions(Fig. 3). Buoyancy <strong>of</strong> the weak felsic granulites presumablyplayed an important role during their transport<strong>and</strong> exhumation. Alternatively, the low viscosity<strong>and</strong> low density <strong>of</strong> the granulites could have allowedtheir rise vertically, through the mantle wedge above thesubduction zone, as suggested for example by Oncken(1998) for the Saxonian granulites. During the ascent,the granulites would also trap the characteristic bodies<strong>of</strong> garnet <strong>and</strong> spinel peridotites. The very limited record<strong>of</strong> the early S1 <strong>and</strong> S2 fabrics precludes distinctionbetween exhumation along the suture zone or thissecond case, as well as a more precise description <strong>of</strong> theearly granulite transport path.Carboniferous extrusion <strong>of</strong> granulites <strong>and</strong> collapse <strong>of</strong>crustal lidThe parallelism <strong>of</strong> the S3 <strong>and</strong> S4 strike with the twomain sutures in the Bohemian Massif – the Saxothuringiansubduction zone in the NW <strong>and</strong> the Moldanubian–Bruniaboundary in the SE (Fig. 1b), indicatesthat at least one <strong>of</strong> these collisional zones generatedhorizontal shortening inducing the D3 <strong>and</strong> D4 events.Both <strong>of</strong> them could have acted as a stiff indentorunderthrusting below the already juxtaposed Tepla´-Barr<strong>and</strong>ian <strong>and</strong> Moldanubian domains. Geochronologicalarguments (e.g. Schma¨ dicke et al., 1995)indicate activity <strong>of</strong> the Saxothuringian subduction <strong>and</strong>subsequent collision between 385 <strong>and</strong> 335 Ma (Figs 2& 12; e.g. Konopa´sek & Schulmann, 2005), a timerange involving the radiometric ages <strong>of</strong> Moldanubi<strong>and</strong>eformation. The easterly Brunia margin documentsprolonged underthrusting from 330 to 310 Ma (e.g.Hartley & Otava, 2001), too young to cause theMoldanubian shortening.The spatial abundance <strong>of</strong> S3 steep fabrics (Fig. 5)<strong>and</strong> their vertical extent (Fig. 10b) suggest homogeneousdevelopment <strong>of</strong> the steep foliation throughoutthe arc (the Central Bohemian Plutonic Complex) <strong>and</strong>back-arc domains (Moldanubian domain in the sense<strong>of</strong> Schulmann et al., 2005, 2009) in a NW–SE dextraltranspressional regime. In the Tepla´-Barr<strong>and</strong>ian Unit,S3 is developed only at the eastern edge (e.g. Zˇa´k et al.,2005a) suggesting that the remote part <strong>of</strong> western forearcupper crust behaved as a stiff block during the D3.Strain localization below the magmatic arc wasresponsible for the unusually massive amplification <strong>of</strong>F3 folds in the lower crust that resulted in localizedexhumation <strong>of</strong> the deep-seated granulites to midcrustallevels in the form <strong>of</strong> a 20 km wide ductileisoclinally folded subvertical fan-like structure (Figs 13& 14b–d). The adjacent mid-crustal units in this casewould be forced to develop marginal synclines aroundthe dome, to balance the voluminous vertical masstransfer through the ductile crust (Fig. 14c,d), draggingthe upper-crustal Palaeozoic Varied sequences tomid-crustal depths. This model for exhumation <strong>of</strong> thegranulites in many aspects is similar to that proposedby Behr (1978), Weber (1984) or Franke & Stein (2000)for exhumation <strong>of</strong> the Saxonian granulites, but it differsfrom the wide range <strong>of</strong> <strong>numerical</strong> models focusedon exhumation in collisional domains (e.g. Burg &Podladchikov, 1999; Gerya & Stockhert, 2006; Jamiesonet al., 2007).The specific tectonic history during exhumation <strong>of</strong>granulites in the Moldanubian domain results from acombination <strong>of</strong> buoyancy <strong>and</strong> the felsic composition,which together are responsible for their distinct <strong>and</strong>transient mechanical behaviour (Franeˇk et al., 2011;Lexa et al., 2011). The abundance <strong>of</strong> felsic granulitesamong Variscan lower crustal rocks suggests thattheir presence at the base <strong>of</strong> continental crust isresponsible for the unique tectonic style presented inthis work. The D4 vertical shortening followedimmediately after the D3 granulite ascent, being mostpronounced near the margin <strong>of</strong> the upper-crustalTepla´-Barr<strong>and</strong>ian block (Figs 13 & 14c,d). Here, theS4 was lubricated by late arc-related magmas <strong>and</strong>syenite magma <strong>and</strong> attained characteristics <strong>of</strong> a normalshear zone with a dip-slip lineation (Zˇa´k et al.,2005a). This intensive vertical shortening in thecentral parts <strong>of</strong> the orogenic root can be explained bya ductile thinning mechanism in which the Tepla´-Barr<strong>and</strong>ian suprastructure slid along the normalshear zone to the NW from the mid- <strong>and</strong> lowercrustal D3 dome described above.The D4 sliding was a mid-crustal expression <strong>of</strong>gravitational spreading <strong>of</strong> the ro<strong>of</strong> (suprastructure) <strong>of</strong>the crustal-scale dome cored by the granulites (infrastructure).The sliding was localized mainly at thethermally weakened volcanic arc, being driven by thefinal D3 dome amplification (Fig. 14c,d). Indeed,recent work by Gerya et al. (2008) <strong>and</strong> Beaumontet al. (2009) shows that rapidly exhumed, buoyant,Ó 2010 Blackwell Publishing Ltd225


74 J. FRANĚK ET AL.weak lower <strong>and</strong> ⁄ or subducted HP crust is mechanicallydecoupled from middle-crustal units until thelater stages <strong>of</strong> its ascent. This is potentially indicatedby the discordance <strong>of</strong> early structures in the granuliteswith respect to the Monotonous <strong>and</strong> Varied Grouprocks. In these <strong>numerical</strong> models, rapid emplacement<strong>of</strong> these weak, hot rocks as <strong>structural</strong> domes in themiddle crust drives lateral flow, leading to extension<strong>and</strong> ductile thinning above <strong>and</strong> adjacent to the dome.With these calculations in mind, the fan-like geometry<strong>of</strong> the S3 fabrics could have been due to the ductilethinning. Exhumation <strong>of</strong> the Moldanubian rocksfrom below the Tepla´-Barr<strong>and</strong>ian Unit was suggestedby earlier authors (e.g. Scheuvens & Zulauf, 2000;Do¨ rr & Zulauf, 2008), but without the support <strong>of</strong><strong>structural</strong> or seismic data covering the broaderMoldanubian domain.CONCLUSIONSThe <strong>structural</strong> evolution <strong>of</strong> the South Bohemiangranulites reveals a complexity not seen in the easternpart <strong>of</strong> the Moldanubian domain. The SouthBohemian granulites were exhumed along two distinctfabrics, the S2 <strong>and</strong> S3, instead <strong>of</strong> in a single verticalchannel known from the eastern Moldanubian(Schulmann et al., 2005, 2008). The S2 indicates adistinct older tectonic episode possibly related to anearly subduction period <strong>and</strong> emplacement <strong>of</strong> orogeniclower crust at the bottom <strong>of</strong> the orogenic root (Franeˇket al., 2011). The gravimetry <strong>and</strong> <strong>structural</strong> geologyindicate that the individual South Bohemian granulitemassifs reach several kilometres depth. A reflectionseismic pr<strong>of</strong>ile additionally depicts a region <strong>of</strong> probablesteep fabrics through the crust below granulites.This vertical region <strong>of</strong> low reflectivity represents thetrace <strong>of</strong> the deformed granulite D3 crustal-scale ascentchannel that probably consists <strong>of</strong> additional deepergranulite bodies.Partially molten lower crust dominated by felsicgranulites was incorporated into middle crustal levelsin the form <strong>of</strong> a D3 syn-compressional crustal-scaledome at 342–337 Ma. After emplacement, the dome<strong>of</strong> felsic granulite deformed together with the surroundingmiddle crust. The subhorizontal S4 fabricimmediately reworked the S3 during the interval342–337 Ma, being induced by gravitational spreading<strong>of</strong> the growing dome. During D4, the Tepla´-Barr<strong>and</strong>ianupper crust slid to the NW away from the domeregion <strong>and</strong> allowed ductile thinning <strong>of</strong> the mid-crustallevel in the Moldanubian domain.This two-stage exhumation mechanism from HPconditions through the orogenic crust for felsic granulitesmay be applicable to exhumation <strong>of</strong> other HPfelsic rocks. The initial low viscosity probably enabledthe buoyancy-driven rise, while cooling <strong>and</strong> continuedshearing resulted in development <strong>of</strong> a large-scaleme´lange <strong>of</strong> granulite bodies dismembered withinmid-crustal rocks.The presence <strong>of</strong> low-density felsic granulites at thebottom <strong>of</strong> the crustal root is a key factor controllingexhumation <strong>of</strong> the orogenic lower crust <strong>and</strong> the tectonicstyle, both driven by buoyancy in addition totectonic far-field forces. The Bohemian Massif representsa field laboratory for many conceptual models toexplain the exhumation <strong>of</strong> orogenic lower crust ingeneral.ACKNOWLEDGEMENTSThe work was supported by a grant from the CzechScience Foundation (GACˇR 205 ⁄ 05 ⁄ 2187) <strong>and</strong> aninternal project <strong>of</strong> the Czech Geological Survey(326700). Visits by J. Franeˇk to ULP Strasbourg werefunded by the French Government Foundation(BGF). K. Schulmann <strong>and</strong> O. Lexa acknowledgefinancial support from the French National GrantAgency (no. 06-1148784) <strong>and</strong> from the Ministry <strong>of</strong>Education <strong>of</strong> the Czech Republic (grant MSM-0021620855). Pr<strong>of</strong>. G. Manatschal is acknowledged forinspiring discussions. We are grateful for the constructive<strong>and</strong> inspiring reviews <strong>of</strong> R. Jamieson <strong>and</strong>M. Pierce, <strong>and</strong> we thank M. Brown for careful editorialwork.REFERENCESAftalion, M., Bowes, D.R. & Vrána, S., 1989. Early carboniferousU-Pb Zircon age for Garnetiferous, Perpotassic Granulites,Blanský Les Massif, Czechoslovakia. Neues JahrbuchFur Mineralogie-Monatshefte, 4, 145–152.Beaumont, C., Jamieson, R.A., Nguyen, M.H. & Medvedev, S.,2004. Crustal channel flows: 1. Numerical models with applicationsto the tectonics <strong>of</strong> the Himalayan-Tibetan orogen.Journal <strong>of</strong> Geophysical Research B: Solid Earth, 109(B6),B06406.Beaumont, C., Nguyen, M.H., Jamieson, R.A. & Ellis, S., 2006.Crustal flow modes in large hot orogens. Geological Society <strong>of</strong>London Special Publication, 268, 91–145.Beaumont, C., Jamieson, R.A., Butler, J.P. & Warren, C.J.,2009. Crustal structure: a key constraint on the mechanism <strong>of</strong>ultra-high-pressure rock exhumation. Earth <strong>and</strong> PlanetaryScience Letters, 287, 116–129.Behr, H.J., 1961. Beitrage zur petrographischen und tektonischenAnalyse des sachsischen Granulitgebirges. FrebiergerForschungshefte C, 119, 5–118.Behr, H.J., 1978. Subfluenz-Prozesse im Grundgebirgs-StockwerkMitteleuropas. Zeitschrift der Deutschen Gesellschaft fu¨rGeowissenschaften, 129, 283–318.Behr, H.J., 1980. Polyphase shear zones in the granulite beltsalong the margins <strong>of</strong> the Bohemian Massif. Journal <strong>of</strong> StructuralGeology, 2, 249–254.Behrmann, J.H. & Tanner, D.C., 1997. Carboniferous tectonics<strong>of</strong> the Variscan basement collage in eastern Bavaria <strong>and</strong>western Bohemia. Geologische Rundschau, 86, S15–S27.Burg, J.P. & Podladchikov, Y., 1999. Lithospheric-scale folding:<strong>numerical</strong> <strong>modelling</strong> <strong>and</strong> application to the Himalayansyntaxes. International Journal <strong>of</strong> Earth Sciences, 88, 190–200.Carswell, D.A. & O’Brien, P.J., 1993. Thermobarometry <strong>and</strong>geotectonic significance <strong>of</strong> high-pressure granulites – examplesfrom the Moldanubian Zone <strong>of</strong> the Bohemian Massif inLower Austria. Journal <strong>of</strong> Petrology, 34, 427–459.Ó 2010 Blackwell Publishing Ltd226


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 75Cha´b, J. & Pelc, Z., 1973. Proterozoische Grauwacken im NW-Teil des Barr<strong>and</strong>iums (Tschechisch). Sbornı´k Geologicky´chVeˇd, 25, 7–84.Chemenda, A.I., Mattauer, M., Malavielle, J. & Bokun, A.N.,1995. A mechanism for syn-collisional rock exhumation <strong>and</strong>associated normal faulting: results from physical <strong>modelling</strong>.Earth <strong>and</strong> Planetary Science Letters, 132, 225–232.Cooke, R.A., OÕBrien, P.J. & Carswell, D.A., 2000. Garnetzoning <strong>and</strong> the identification <strong>of</strong> equilibrium mineral compositionsin high-pressure-temperature granulites from theMoldanubian Zone, Austria. Journal <strong>of</strong> Metamorphic Geology,18, 551–569.Dallmayer, R.D., Franke, W. & Weber, K., 1995. Pre-PermianGeology <strong>of</strong> Central <strong>and</strong> Eastern Europe. Springer-Verlag,Berlin Heidelberg.Dewey, J.F., Ryan, P.D. & Andersen, T.B., 1993. Orogenic uplift<strong>and</strong> collapse, crustal thickness, fabrics <strong>and</strong> metamorphic phasechanges: the role <strong>of</strong> eclogites. In: Magmatic Processes <strong>and</strong>Plate Tectonics (eds Prichard, H.M., Alabaster, T., Harris,N.B.W. & Neary Christopher, R.) Geological Society SpecialPublications, pp. 325–343. Geological Society <strong>of</strong> London,London, UK.Do¨ rr, W. & Zulauf, G., 2008. Elevator tectonics <strong>and</strong> orogeniccollapse <strong>of</strong> a Tibetan-style plateau in the European Variscides:the role <strong>of</strong> the Bohemian shear zone. International Journal <strong>of</strong>Earth Sciences, 99, Doi: 10.1007/s00531-008-0389-x.Dudek, A., 1980. The crystalline basement block <strong>of</strong> the OuterCarpathians in Moravia. Rozpravy Cˇeskoslovenske´ AkademieVeˇd, 90, 1–85.Du¨ rbaum, H.J., Franke, W., Stein, E. et al., 1999. Structure <strong>of</strong>the Saxonian Granulites: geological <strong>and</strong> geophysical constraintson the exhumation <strong>of</strong> high-pressure ⁄ high-temperaturerocks in the mid-European Variscan belt. Tectonics, 18, 756–773.Edel, J.B. & Weber, K., 1995. Cadomian terranes, wrenchfaulting <strong>and</strong> thrusting in the central Europe Variscides: geophysical<strong>and</strong> geological evidence. Geologische Rundschau, 84,412–432.Edel, J.B., Schulmann, K. & Holub, F.V., 2003. Anticlockwise<strong>and</strong> clockwise rotations <strong>of</strong> the Eastern Variscides accommodatedby dextral lithospheric wrenching: palaeomagnetic <strong>and</strong><strong>structural</strong> evidence. Journal <strong>of</strong> the Geological Society, 160,209–218.Engvik, A.K. & Andersen, T.B., 2000. Evolution <strong>of</strong> Caledoni<strong>and</strong>eformation fabrics under eclogite <strong>and</strong> amphibolite facies atVardalsneset, Western Gneiss Region, Norway. Journal <strong>of</strong>Metamorphic Geology, 18, 241–257.Faryad, S.W., Perraki, M. & Vrána, S., 2006. P-T evolution <strong>and</strong>reaction textures in retrogressed eclogites from Sveˇtlı´k, theMoldanubian Zone (Czech Republic). Mineralogy <strong>and</strong>Petrology, 88, 297–319.Finger, F. & Steyer, H.P., 1995. A tectonic model for the easternVariscides: indications from a chemical study <strong>of</strong> amphibolitesin the south-eastern Bohemian Massif, Austria. GeologicaCarpathica, 46, 1–14.Finger, F. & von Quadt, A., 1995. U ⁄ Pb ages from zircons froma plagiogranite-gneiss in the south-eastern Bohemian Massif,Austria – further evidence for an important early Paleozoicrifting episode in the eastern Variscides. Schweizerische Mineralogischeund Petrographische Mitteilungen, 75, 265–270.Finger, F., Roberts, M.P., Haunschmid, B., Schermaier, A. &Steyrer, H.P., 1997. Variscan granitoids <strong>of</strong> central Europe:their typology, potential sources <strong>and</strong> tectonothermal relations.Mineralogy <strong>and</strong> Petrology, 61, 67–96.Finger, F., Gerdes, A., Janousˇek, V., Rene´, M. & Riegler, G.,2007. Resolving the Variscan evolution <strong>of</strong> the Moldanubiansector <strong>of</strong> the Bohemian Massif: the significance <strong>of</strong> theBavarian <strong>and</strong> the Moravo–Moldanubian tectonometamorphicphases. Journal <strong>of</strong> Geosciences, 52, 9–28.Franeˇk, J., Schulman, K. & Lexa, O., 2006. Kinematic <strong>and</strong>rheological model <strong>of</strong> exhumation <strong>of</strong> high pressure granulites inthe Variscan orogenic root: example <strong>of</strong> the Blansky´ les granulite,Bohemian Massif, Czech Republic. Mineralogy <strong>and</strong>Petrology, 86, 253–276.Franeˇk, J., Schulmann, K., Lexa, O. et al., 2011. Origin <strong>of</strong> felsicgranulite microstructure by heterogeneous decomposition<strong>of</strong> alkali feldspar <strong>and</strong> extreme weakening <strong>of</strong> orogenic lowercrust during the Variscan orogeny. Journal <strong>of</strong> MetamorphicGeology, 29, 103–130.Franke, W., 1993. The Saxonian granulites: a metamorphic corecomplex? Geologische Rundschau, 82, 505–515.Franke, W., 2000. The mid-European segment <strong>of</strong> the Variscides:tectonostratigraphic units, terrane boundaries <strong>and</strong> plate tectonicevolution. Geological Society Special Publication, 179,35–56.Franke, W. & Stein, E., 2000. Exhumation <strong>of</strong> high-grade rocksin the Saxo-Thuringian Belt: geological constraints <strong>and</strong> geodynamicconcepts. Geological Society, London, Special Publications,179, 337–354.Friedl, G., von Quadt, A., Ochsner, A. & Finger, F., 1993.Timing <strong>of</strong> the Variscan orogeny in the Southern BohemianMassif (NE Austria) deduced from new U-Pb zircon <strong>and</strong>monazite dating, Terra Abstracts, 5, 235–236.Friedl, G., Cooke, R., Finger, F., McNaughton, N.J. & Fletcher,I., 2003. U-Pb shrimp dating <strong>and</strong> trace element investigationson multiple zoned zircons from a South-Bohemian granulite. Journal <strong>of</strong> the Czech Geological Society,48, 51.Friedl, G., Finger, F., Paquette, J.L., von Quadt, A.,McNaughton, N.J. & Fletcher, I.R., 2004. Pre-Variscan geologicalevents in the Austrian part <strong>of</strong> the Bohemian Massifdeduced from U-Pb zircon ages. International Journal <strong>of</strong> EarthSciences, 93, 802–823.Fuchs, G., 1976. Zur Entwicklung der Bo¨ hmischen Masse.Jahrbuch des Geologischen Bundesanstalts, 119, 45–61.Gebauer, D., Williams, I.S., Compston, W. & Grunenfelder, M.,1989. The development <strong>of</strong> the Central European continentalcrust since the Early Archaean based on conventional <strong>and</strong> ionmicroprobedating <strong>of</strong> up to 3.84 b.y. old detrital zircons.Tectonophysics, 157, 81–96.Gerya, T. & Stockhert, B., 2006. Two-dimensional <strong>numerical</strong>modeling <strong>of</strong> tectonic <strong>and</strong> metamorphic histories at activecontinental margins. International Journal <strong>of</strong> Earth Sciences,95, 250–274.Gerya, T.V., Perchuk, L.L. & Burg, J.-P., 2008. Transient hotchannels: perpetrating <strong>and</strong> regurgitating ultrahigh-pressure,high-temperature crust–mantle associations in collision belts.Lithos, 103, 236–256.Glasmacher, U.A., Mann, U. & Wagner, G.A., 2002. Thermotectonicevolution <strong>of</strong> the Barr<strong>and</strong>ian, Czech Republic, as revealedby apatite fission-track analysis. Tectonophysics, 359,381–402.Hartley, A.J. & Otava, J., 2001. Sediment provenance <strong>and</strong> dispersalin a deep marine forel<strong>and</strong> basin: the Lower CarboniferousCulm Basin, Czech Republic. Journal <strong>of</strong> the GeologicalSociety, 158, 137–150.Hrubcova´, P., Sroda, P., Sˇpicˇák, A. et al., 2005. Crustal <strong>and</strong>uppermost mantle structure <strong>of</strong> the Bohemian Massif based onCELEBRATION 2000 data. Journal <strong>of</strong> Geophysical ResearchB: Solid Earth, 110, 1–21.Jamieson, R.A., Beaumont, C., Nguyen, M.H. & Culshaw,N.G., 2007. Synconvergent ductile flow in variable-strengthcontinental crust: <strong>numerical</strong> models with application to thewestern Grenville orogen. Tectonics, 26, TC5005.Janousˇek, V. & Gerdes, A., 2003. Timing the magmatic activitywithin the Central Bohemian Pluton, Czech Republic: conventionalU-Pb ages for the Sazava <strong>and</strong> Tabor intrusions <strong>and</strong>their geotectonic significance. Journal <strong>of</strong> the Czech GeologicalSociety, 48, 70–71.Janousˇek, V. & Holub, F.V., 2007. The causal link between HP-HT metamorphism <strong>and</strong> ultrapotassic magmatism in collisionalorogens: case study from the Moldanubian Zone <strong>of</strong> theBohemian Massif. Proceedings <strong>of</strong> the Geologists Association,118, 75–86.Ó 2010 Blackwell Publishing Ltd227


76 J. FRANĚK ET AL.Janousˇek, V., Vokurka, K. & Vra´na, S., 1997. Izotopy stroncia aneodymu v amfibolitech pestre´ skupiny moldanubika v okolı´Chýnova, Zpra´vy o Geologicky´ch Vy´zkumech, 1996, 80–81.Janousˇek, V., Bowes, D.R., Rogers, G., Farrow, C.M. & Jelinek,E., 2000. Modelling diverse processes in the petrogenesis <strong>of</strong> acomposite batholith: the Central Bohemian Pluton, CentralEuropean Hercynides. Journal <strong>of</strong> Petrology, 41, 511–543.Janousˇek, V., Finger, F., Roberts, M., Frýda, J., Pin, C. &Dolejs, D., 2004. Deciphering the petrogenesis <strong>of</strong> deeplyburied granites: whole-rock geochemical constraints on theorigin <strong>of</strong> largely undepleted felsic granulites from the MoldanubianZone <strong>of</strong> the Bohemian Massif. Transactions <strong>of</strong> theRoyal Society <strong>of</strong> Edinburgh-Earth Sciences, 95, 141–159.Kalt, A., Corfu, F. & Wijbrams, J., 1997. P-T-t Pfade variszicherHochtemperatur-Metamorphite des Bayerischen Waldes.European Journal <strong>of</strong> Mineralogy, 9, 19.Kalt, A., Berger, A. & Blumel, P., 1999. Metamorphic evolution<strong>of</strong> cordierite-bearing migmatites from the Bayerische Wald(Variscan Belt, Germany). Journal <strong>of</strong> Petrology, 40, 601–627.Kettner, R., 1918. Versuch einer stratigraphischen Einteilungdes bo¨ hmischen Algonkiums. Geologische Rundschau, 8, 169–188.Kodym, O., 1972. Multiphase deformation in the Blansky´ lesgranulite massif (South Bohemia). Krystalinikum, 9, 91–105.Konopa´sek, J. & Schulmann, K., 2005. Contrasting Early Carboniferousfield geotherms: evidence for accretion <strong>of</strong> a thickenedorogenic root <strong>and</strong> subducted Saxothuringian crust(Central European Variscides). Journal <strong>of</strong> the GeologicalSociety, 162, 463–470.Konopa´sek, J., Schulmann, K. & Johan, V., 2002. Eclogite-faciesmetamorphism at the eastern margin <strong>of</strong> the Bohemian Massif –subduction prior to continental underthrusting? EuropeanJournal <strong>of</strong> Mineralogy, 14, 701–713.Kosˇler, J., Kelley, S.P., Vance, D. & Svojtka, M., 1999. Independentdating <strong>of</strong> cooling <strong>and</strong> decompression <strong>of</strong> high-graderocks in the southern Bohemian Massif with Ar-Ar, Sm-Nd<strong>and</strong> U-Pb techniques. Journal <strong>of</strong> Conference Abstracts, 4, 39.Kossmat, F., 1927. Gliederung der varistischen Gebirgsbaues.Abh<strong>and</strong>lungen des Sa¨chsischen Geologischen L<strong>and</strong>esamts, 1,1–39.Kotkova´, J., 1993. Tectonometamorphic history <strong>of</strong> lower crustin Bohemian Massif: example <strong>of</strong> north Bohemian granulites.Special paper 2, Czech Geological Survey, Prague, 42 pp.Kotkova´, J., Kro¨ ner, A., Todt, W. & Fiala, J., 1996. Zircondating <strong>of</strong> north Bohemian granulites, Czech Republic: furtherevidence for the lower carboniferous high-pressure event in theBohemian Massif. Geologische Rundschau, 85, 154–161.Koyi, H.A., Milnes, A.G., Schmeling, H., Talbot, C.J., Juhlin,C. & Zeyen, H., 1999. Numerical models <strong>of</strong> ductile rebound <strong>of</strong>crustal roots beneath mountain belts. Geophysical JournalInternational, 139, 556–562.Kretz, R., 1983. Symbols for rock forming minerals. AmericanMineralogist, 68, 277–279.Kro¨ ner, A. & Willner, A.P., 1998. Time <strong>of</strong> formation <strong>and</strong> peak<strong>of</strong> Variscan HP-HT metamorphism <strong>of</strong> quartz-feldspar rocks inthe central Erzgebirge, Saxony, Germany. Contributions toMineralogy <strong>and</strong> Petrology, 132, 1–20.Kro¨ ner, A., Wendt, I., Liew, T.C. et al., 1988. U-Pb zircon <strong>and</strong>Sm-Nd model ages <strong>of</strong> high-grade Moldanubian metamsediments,Bohemian Massif, Czechoslovakia. Contributions toMineralogy <strong>and</strong> Petrology, 99, 257–266.Kro¨ ner, A., OÕBrien, P.J., Nemchin, A.A. & Pidgeon, R.T.,2000. Zircon ages for high pressure granulites from SouthBohemia, Czech Republic, <strong>and</strong> their connection to Carboniferoushigh temperature processes. Contributions to Mineralogy<strong>and</strong> Petrology, 138, 127–142.Law, R.D., Miller, E.L., Little, T.A. & Lee, J., 1994. Extensionalorigin <strong>of</strong> ductile fabrics in the Schist Belt, Central BrooksRange, Alaska-II. Micro<strong>structural</strong> <strong>and</strong> petr<strong>of</strong>abric evidence.Journal <strong>of</strong> Structural Geology, 16, 919–940.Lexa, O., Schulmann, K., Janousˇek, V., Sˇtı´pska´, P., Guy, A. &Racek, M., 2011. Heat sources <strong>and</strong> trigger mechanisms <strong>of</strong>exhumation <strong>of</strong> HP granulites in Variscan orogenic root.Journal <strong>of</strong> Metamorphic Geology, 29, 79–102.Linner, M., 1996. Metamorphism <strong>and</strong> partial melting <strong>of</strong> paragneisses<strong>of</strong> the Monotonous Group, SE Moldanubicum(Austria). Mineralogy <strong>and</strong> Petrology, 58, 215–234.Little, T.A., Miller, E.L., Lee, J. & Law, R.D., 1994. Extensionalorigin <strong>of</strong> ductile fabrics in the Schist Belt, Central BrooksRange, Alaska-I. Geologic <strong>and</strong> <strong>structural</strong> studies. Journal <strong>of</strong>Structural Geology, 16, 899–918.Medaris, G., Jelı´nek, E. & Mı´sarˇ, Z., 1995. Czech eclogites –terrane settings <strong>and</strong> implications for Variscan tectonic evolution<strong>of</strong> the Bohemian Massif. European Journal <strong>of</strong> Mineralogy,7, 7–28.Medaris, L.G., Beard, B.L. & Jelı´nek, E., 2006. Mantle-derived,UHP garnet pyroxenite <strong>and</strong> eclogite in the MoldanubianGfohl nappe, Bohemian Massif: a geochemical review, newP-T determinations, <strong>and</strong> tectonic interpretation. InternationalGeology Review, 48, 765–777.Novák, M., Cˇerny´, P., Kimbrough, D.L., Taylor, M.C. & Ercit,T.S., 1998. U-Pb ages <strong>of</strong> monazite from granitic pegmatites inthe Moldanubian Zone <strong>and</strong> their geological implications. ActaUniversitatis Carolinae-Geologica, 42, 309–310.OÕBrien, P.J., 1997. Garnet zoning <strong>and</strong> reaction textures inoverprinted eclogites, Bohemian Massif, European Variscides:a record <strong>of</strong> their thermal history during exhumation. Lithos,41, 119–133.OÕBrien, P.J., 2008. Challenges in high-pressure granulite metamorphismin the era <strong>of</strong> pseudosections: reaction textures,compositional zoning <strong>and</strong> tectonic interpretation with examplesfrom the Bohemian Massif. Journal <strong>of</strong> MetamorphicGeology, 26, 235–251.OÕBrien, P.J. & Carswell, D.A., 1993. Tectonometamorphicevolution <strong>of</strong> the Bohemian Massif – evidence from highpressuremetamorphic rocks. Geologische Rundschau, 82, 531–555.OÕBrien, P.J. & Ro¨ tzler, J., 2003. High-pressure granulites: formation,recovery <strong>of</strong> peak conditions <strong>and</strong> implications fortectonics. Journal <strong>of</strong> Metamorphic Geology, 21, 3–20.OÕBrien, P.J. & Seifert, K., 1992. P–T–t paths as records <strong>of</strong>orogenic processes: examples <strong>and</strong> problems from the crystalline<strong>of</strong> the Bohemian Massif. Terra Nostra, 1992, 58–59.OÕBrien, P.J. & Vra´na, S., 1995. Eclogites with a short-livedgranulite facies overprint in the Moldanubian zone, CzechRepublic: petrology, geochemistry <strong>and</strong> diffusion <strong>modelling</strong> <strong>of</strong>garnet zoning. Geologische Rundschau, 84, 473–488.OÕBrien, P.J., Kro¨ ner, A., Jaeckel, P., Hegner, E., Zelazniewicz,A. & Kryza, R., 1997. Petrological <strong>and</strong> isotopic studies onpalaeozoic high-pressure granulites, Gory Sowie Mts, PolishSudetes. Journal <strong>of</strong> Petrology, 38, 433–456.Oncken, O., 1998. Tectonic mass transfer prior to collision –subduction erosion versus accretion in the Mid-EuropeanVariscides. Acta Universitatis Carolinae-Geologica, 42, 313–314.Petrakakis, K., 1997. Evolution <strong>of</strong> Moldanubian rocks in Austria:review <strong>and</strong> synthesis. Journal <strong>of</strong> Metamorphic Geology,15, 203–222.Pitra, P., Burg, J.P. & Guiraud, M., 1999. Late Variscan strikesliptectonics between the Tepla´- Barr<strong>and</strong>ian <strong>and</strong> Moldanubianterranes (Czech Bohemian Massif): petro<strong>structural</strong>evidence. Journal <strong>of</strong> the Geological Society, 156, 1003–1020.Platt, J.P., 1993. Exhumation <strong>of</strong> high-pressure rocks: a review <strong>of</strong>concepts <strong>and</strong> processes. Terra Nova, 5, 119–133.Plomerová, J., Vecsey, L., Babuska, V., Granet, M. & Achauer,U., 2005. Passive seismic experiment Mosaic – a pilot study <strong>of</strong>mantle lithosphere anisotropy <strong>of</strong> the Bohemian Massif. StudiaGeophysica et Geodaetica, 49, 541–560.Prince, C.I., Kosˇler, J., Vance, D. & Gunther, D., 2000. Comparison<strong>of</strong> laser ablation ICP-MS <strong>and</strong> isotope dilution REE<strong>analyses</strong> – implications for Sm-Nd garnet geochronology.Chemical Geology, 168, 255–274.Racek, M., Sˇtı´pska´, P., Pitra, P., Schulmann, K. & Lexa, O.,2006. Metamorphic record <strong>of</strong> burial <strong>and</strong> exhumation <strong>of</strong>Ó 2010 Blackwell Publishing Ltd228


EXTRUSIONOFLOWERCRUSTINVARISCANOROGEN 77orogenic lower <strong>and</strong> middle crust: a new tectonothermal modelfor the Drosendorf window (Bohemian Massif, Austria).Mineralogy <strong>and</strong> Petrology, 86, 221–251.Rajlich, P., 1988. Tektonika sz. okraje strˇedocˇeske´ho plutonu avariska´ transprese v bloku bohemika. Sbornı´k Geologicky´chVeˇd, 43, 9–81.Rajlich, P., Synek, J., Sarbach, M. & Schulmann, K., 1986.Hercynian-thrust related shear zones <strong>and</strong> deformation <strong>of</strong> thevaried group on the contact <strong>of</strong> granulites Southern Moldanubian,Bohemian Massif. Geologische Rundschau, 75, 665–683.Rey, P., V<strong>and</strong>erhaeghe, O. & Teyssier, C., 2001. Gravitationalcollapse <strong>of</strong> the continental crust: definition, regimes <strong>and</strong>modes. Tectonophysics, 342, 435–449.Rey, P., Teyssier, C. & Whitney, D.L., 2009. The role <strong>of</strong> partialmelting <strong>and</strong> extensional strain rates in the development <strong>of</strong>metamorphic core complexes. Tectonophysics, 477, 135–144.Romer, R.L. & Ro¨ tzler, J., 2001. P-T-t evolution <strong>of</strong> ultrahightemperaturegranulites from the Saxon Granulite Massif,Germany. Part II: geochronology. Journal <strong>of</strong> Petrology, 42,2015–2032.Ro¨ tzler, J. & Romer, R.L., 2001. P-T-t evolution <strong>of</strong> ultrahightemperaturegranulites from the Saxon Granulite Massif,Germany. Part I: petrology. Journal <strong>of</strong> Petrology, 42, 1995–2013.Ru˚zˇek, B., Hrubcova´, P., Novotna´, M., Sˇpicˇa´k, A. & Karousova´,O., 2007. Inversion <strong>of</strong> travel times obtained during activeseismic refraction experiments CELEBRATION 2000, ALP2002 <strong>and</strong> SUDETES 2003. Studia Geophysica et Geodaetica,51, 141–164.Scheuvens, D., 1999. Die tektonometamorphe und kinematischeEntwicklung im Westteil der Zentralbohmischen Scherzone(Bohmische Masse): evidenz fur variscischen Kollaps. FrankfurterGeowissenschaftliche Arbeiten, 18, 273.Scheuvens, D., 2002. Metamorphism <strong>and</strong> microstructures alonga high-temperature metamorphic field gradient: the northeasternboundary <strong>of</strong> the Kralovsky hvozd unit (BohemianMassif, Czech Republic). Journal <strong>of</strong> Metamorphic Geology, 20,413–428.Scheuvens, D. & Zulauf, G., 2000. Exhumation, strain localization,<strong>and</strong> emplacement <strong>of</strong> granitoids along the western part<strong>of</strong> the Central Bohemian shear zone (Bohemian Massif).International Journal <strong>of</strong> Earth Sciences, 89, 617–630.Schma¨ dicke, E., Mezger, K., Cosca, M.A. & Okrusch, M., 1995.Variscan Sm-Nd <strong>and</strong> Ar-Ar ages <strong>of</strong> eclogite facies rocks fromthe Erzgebirge, Bohemian Massif. Journal <strong>of</strong> MetamorphicGeology, 13, 537–552.Schmidt, S. & Go¨ tze, H.J., 1999. Integration <strong>of</strong> data constraints<strong>and</strong> potential field <strong>modelling</strong> – an example from southernlower saxony, Germany. Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth,Part A: Solid Earth <strong>and</strong> Geodesy, 24, 191–196.Schulmann, K., Kro¨ ner, A., Hegner, E. et al., 2005. Chronologicalconstraints on the pre-orogenic history, burial <strong>and</strong>exhumation <strong>of</strong> deep-seated rocks along the eastern margin <strong>of</strong>the Variscan Orogen, Bohemian Massif, Czech Republic.American Journal <strong>of</strong> Science, 305, 407–448.Schulmann, K., Lexa, O., Sˇtı´pska´, P. et al., 2008. Verticalextrusion <strong>and</strong> horizontal channel flow <strong>of</strong> orogenic lower crust:key exhumation mechanisms in large hot orogens? Journal <strong>of</strong>Metamorphic Geology, 26, 273–297.Schulmann, K., Konopa´sek, J., Janousˇek, V. et al., 2009. AnAndean type Palaeozoic convergence in the Bohemian Massif.Comptes Rendus – Geoscience, 341, 266–286.Schulzschmalschlager, M., 1984. U-Pb Studies <strong>of</strong> zircons <strong>and</strong>monazites <strong>of</strong> the anterior Bavarian Forest. Fortschritte derMineralogie, 62, 223.Siebel, W., Thiel, M. & Chen, F., 2006. Zircon geochronology<strong>and</strong> compositional record <strong>of</strong> late- to post-kinematic granitoidsassociated with the Bavarian Pfahl zone (Bavarian Forest).Mineralogy <strong>and</strong> Petrology, 86, 45–62.Sla´ma, J., Kosˇler, J., Condon, D.J. et al., 2008. Plesˇovice zircon – anew natural reference material for U-Pb <strong>and</strong> Hf isotopicmicroanalysis. Chemical Geology, 249, 1–35.Sokoutis, D., Burg, J.P., Bonini, M., Corti, G. & Cloetingh, S.,2005. Lithospheric-scale structures from the perspective <strong>of</strong>analogue continental collision. Tectonophysics, 406, 1–15.Sˇtı´pská, P. & Powell, R., 2005. Does ternary feldspar constrainthe metamorphic conditions <strong>of</strong> high-grade meta-igneousrocks? Evidence from orthopyroxene granulites, BohemianMassif Journal <strong>of</strong> Metamorphic Geology, 23, 627–647.Sˇtı´pská, P., Schulmann, K. & Kro¨ ner, A., 2004. Vertical extrusion<strong>and</strong> middle crustal spreading <strong>of</strong> omphacite granulite: a model<strong>of</strong> syn-convergent exhumation (Bohemian Massif, CzechRepublic). Journal <strong>of</strong> Metamorphic Geology, 22, 179–198.Suess, F.E., 1926. Intrusionstektonik und W<strong>and</strong>ertektonik imVariszichen Grundgebirge. Verlag Bornttrager, Berlin.Svojtka, M., Kosˇler, J. & Venera, Z., 2002. Dating granulitefaciesstructures <strong>and</strong> the exhumation <strong>of</strong> lower crust in theMoldanubian Zone <strong>of</strong> the Bohemian Massif. InternationalJournal <strong>of</strong> Earth Sciences, 91, 373–385.Tajčmanova´, L., Konopa´sek, J. & Schulmann, K., 2006. Thermalevolution <strong>of</strong> the orogenic lower crust during exhumationwithin a thickened Moldanubian root <strong>of</strong> the Variscan belt <strong>of</strong>Central Europe. Journal <strong>of</strong> Metamorphic Geology, 24, 119–134.Tajčmanova´, L., Konopa´sek, J. & Connolly, J.A.D., 2007. Diffusion-controlleddevelopment <strong>of</strong> silica-undersaturated domainsin felsic granulites <strong>of</strong> the Bohemian Massif (Variscanbelt <strong>of</strong> Central Europe). Contributions to Mineralogy <strong>and</strong>Petrology, 153, 237–250.Timmermann, H., Sˇteˇdra´, V., Gerdes, A., Noble, S.R., Parrish,R.R. & Do¨ rr, W., 2004. The problem <strong>of</strong> dating high-pressuremetamorphism: a U-Pb isotope <strong>and</strong> geochemical study oneclogites <strong>and</strong> related rocks <strong>of</strong> the Marianske Lazne Complex,Czech Republic. Journal <strong>of</strong> Petrology, 45, 1311–1338.Tomek, Cˇ., Dvorˇákova´, V. & Vra´na, S., 1997. Geologicalinterpretation <strong>of</strong> the 9HR <strong>and</strong> 503M seismic pr<strong>of</strong>iles in WesternBohemia. In: Journal <strong>of</strong> Geological Sciences, GeologicalModel <strong>of</strong> Western Bohemia Related to the KTB Borehole inGermany (eds Vra´na, S. & Sˇtědrá, V.), pp. 43–50. CzechGeological Survey, Prague.Van Breemen, O., Aftalion, M., Bowes, D.R. et al., 1982. Geochronologicalstudies <strong>of</strong> the Bohemian massif, Czechoslovakia,<strong>and</strong> their significance in the evolution <strong>of</strong> Central Europe.Transactions <strong>of</strong> the Royal Society <strong>of</strong> Edinburgh: Earth Sciences,73, 89–108.Verner, K., Zˇa´k, J., Nahodilova´, R. & Holub, F.V., 2008.Magmatic fabrics <strong>and</strong> emplacement <strong>of</strong> the cone-sheet-bearingKnı´zˇecı´ Stolec durbachitic pluton (Moldanubian Unit, BohemianMassif): implications for mid-crustal reworking <strong>of</strong>granulitic lower crust in the Central European Variscides.International Journal <strong>of</strong> Earth Sciences, 97, 19–33.Vrána, S., 1989. Perpotassic granulites from Southern Bohemia –a new rock-type derived from partial melting <strong>of</strong> crustal rocksunder upper mantle conditions. Contributions to Mineralogy<strong>and</strong> Petrology, 103, 510–522.Vrána, S., 1997. Geology <strong>and</strong> petrology <strong>of</strong> the Moldanubianzone. In: Geological Model <strong>of</strong> Western Bohemia (eds Vra´na, S.&Sˇteˇdrá, V.), pp. 109–113. Czech Geological Survey, Prague.Vrána, S. & Ba´rtek, J., 2005. Retrograde metamorphism in aregional shear zone <strong>and</strong> related chemical changes: The KapliceUnit <strong>of</strong> muscovite-biotite gneisses in the Moldanubian Zone<strong>of</strong> southern Bohemia, Czech Republic. Journal <strong>of</strong> the CzechGeological Society, 50, 43–57.Vrána, S. & Sˇra´mek, J., 1999. Geological interpretation <strong>of</strong>detailed gravity survey <strong>of</strong> the granulite complex in southernBohemia <strong>and</strong> its structure. Bulletine <strong>of</strong> Czech GeologicalSurvey, 74, 261–277.Vrána, S., Blu¨ mel, P. & Petrakakis, K., 1995. Moldanubianzone: metamorphic evolution. In: Pre-Permian Geology <strong>of</strong> theCentral <strong>and</strong> Western Europe (eds Dallmeyer, D., Franke, W. &Weber, K.), pp. 453–466. Springer-Verlag, Berlin.Warren, C.J., Beaumont, C. & Jamieson, R.A., 2008. Deepsubduction <strong>and</strong> rapid exhumation: role <strong>of</strong> crustal strength <strong>and</strong>strain weakening in continental subduction <strong>and</strong> ultrahighpressurerock exhumation. Tectonics, 27, TC6002.Ó 2010 Blackwell Publishing Ltd229


78 J. FRANĚK ET AL.Weber, K., 1984. Variation in tectonic style with time (Variscan<strong>and</strong> Proterozoic systems, Europe). In: Patterns <strong>of</strong> Change inEarth Evolution. Report <strong>of</strong> the Dahlem workshop (eds Holl<strong>and</strong>,H.D. & Trendall, A.F.), pp. 371–386. Springer-Verlag, Berlin.Weber, K. & Behr, H.J., 1983. Geodynamic interpretation <strong>of</strong> theVariscides. In: Intracontinental Fold Belts (eds Martin, H. &Eder, F.W.), pp. 427–469. Springer-Verlag, Berlin.Wendt, J.I., Kro¨ ner, A., Fiala, J. & Todt, W., 1993. Evidencefrom zircon dating for existence <strong>of</strong> approximately 2.1 Ga oldcrystalline basement in southern Bohemia, Czech Republic.Geologische Rundschau, 82, 42–50.Wendt, J.I., Kro¨ ner, A., Fiala, J. & Todt, W., 1994. U-Pb zircon<strong>and</strong> Sm-Nd dating <strong>of</strong> Moldanubian HP ⁄ HT granulites fromsouth Bohemia, Czech Republic. Journal <strong>of</strong> the GeologicalSociety, London, 151, 83–90.Zˇák, J., Holub, F.V. & Verner, K., 2005a. Tectonic evolution <strong>of</strong>a continental magmatic arc from transpression in the uppercrust to exhumation <strong>of</strong> mid-crustal orogenic root recorded byepisodically emplaced plutons: the Central Bohemian PlutonicComplex (Bohemian Massif). International Journal <strong>of</strong> EarthSciences, 94, 385–400.Zˇák, J., Schulmann, K. & Hrouda, F., 2005b. Multiple magmaticfabrics in the Sa´zava pluton (Bohemian Massif, CzechRepublic): a result <strong>of</strong> superposition <strong>of</strong> wrench-dominatedregional transpression on final emplacement. Journal <strong>of</strong>Structural Geology, 27, 805–822.Zulauf, G., Do¨ rr, W., Fiala, J. & Vejnar, Z., 1997. Late Cadomiancrustal tilting <strong>and</strong> Cambrian transtension in the Tepla-Barr<strong>and</strong>ian unit (Bohemian Massif, Central European Variscides).Geologische Rundschau, 86, 571–584.Zulauf, G., Bues, C., Do¨ rr, W. & Vejnar, Z., 2002. 10 km minimumthrow along the West Bohemian shear zone: evidencefor dramatic crustal thickening <strong>and</strong> high topography inBohemian Massif (European Variscides). International Journal<strong>of</strong> Earth Sciences, 91, 850–864.SUPPORTING INFORMATIONAdditional Supporting Information may be found inthe online version <strong>of</strong> this article:Appendix S1. Radiometric ages used for histogramcalculations in Fig. 2.Appendix S2. Mineralogical densities used as startingvalues in three-dimensional gravity <strong>modelling</strong>.Please note: Wiley-Blackwell are not responsible forthe content or functionality <strong>of</strong> any supporting materialssupplied by the authors. Any queries (other thanmissing material) should be directed to the correspondingauthor for the article.Received 26 February 2010; revision accepted 16 August 2010.Ó 2010 Blackwell Publishing Ltd230


J. metamorphic Geol., 2005, 23, 649–666 doi:10.1111/j.1525-1314.2005.00601.xContrasting textural record <strong>of</strong> two distinct metamorphic events <strong>of</strong>similar P–T conditions <strong>and</strong> different durationsO. LEXA, 1,2 P. ŠTÍPSKÁ, 1,2 K. SCHULMANN, 1,2 L. BARATOUX 1 AND A. KRÖNER 31 Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University, Albertov 6, 128 43 Prague, Czech Republic(lexa@natur.cuni.cz)2 Université Louis Pasteur, CGS/EOST, UMR 7517, 1 rue Blessig, Strasbourg 67084, France3 Institut für Geowissenschaften, Universität Mainz, 55099 Mainz, GermanyABSTRACTA <strong>structural</strong>, metamorphic <strong>and</strong> geochronological study <strong>of</strong> the Stare´ Meˇsto belt implies the existence <strong>of</strong>two distinct metamorphic events <strong>of</strong> similar peak P–T conditions (700–800 °C, 8–10 kbar) during theCambro-Ordovician <strong>and</strong> the Carboniferous tectonometamorphic events. The hypothesis <strong>of</strong> two distinctperiods <strong>of</strong> metamorphism was suggested on the basis <strong>of</strong> <strong>structural</strong> discordance between an undoubtedlyCarboniferous granodiorite sill intrusion <strong>and</strong> earlier Cambro-Ordovician fabrics <strong>of</strong> a b<strong>and</strong>edamphibolite complex. The analysis <strong>of</strong> crystal size distribution (CSD) shows high nucleation density(N 0 ) <strong>and</strong> low average growth rate (Gt) for Carboniferous mylonitic metagabbros <strong>and</strong> myloniticgranodiorites. The parameter N 0 decreases whereas the quantity Gt increases towards highertemperatures progressively approaching the values obtained from the Cambro-Ordovician b<strong>and</strong>edamphibolite complex. The spatial distribution <strong>of</strong> amphibole <strong>and</strong> plagioclase shows intense mechanicalmixing for lower-temperature mylonitic metagabbros. In high-temperature mylonites a strong aggregatedistribution is developed. Cambro-Ordovician amphibolites unaffected by Carboniferous deformationshow a regular to anticlustered spatial distribution resulting from heterogeneous nucleation <strong>of</strong> individualphases. This pattern, together with CSD, was subsequently modified by the grain growth <strong>and</strong> texturalequilibration controlled by diffusive mass transfer during Carboniferous metamorphism. The differencesbetween the observed textures <strong>of</strong> the amphibolites are interpreted to be a consequence <strong>of</strong> the differentdurations <strong>of</strong> the Carboniferous <strong>and</strong> Cambro-Ordovician thermal events.Key words: Cambro-Ordovician <strong>and</strong> Carboniferous metamorphism; quantitative textural analysis;crystal size distributions; grain contact frequencies.INTRODUCTIONIdentifying distinct metamorphic episodes in domainswith a polymetamorphic history is possible providingthe P–T conditions <strong>of</strong> a younger metamorphic eventare markedly different from those <strong>of</strong> a preceding one.The problem becomes more complex when the morerecent metamorphic event affects units which previouslysuffered a metamorphic event <strong>of</strong> a similar grade.In this particular case it is only the geological context<strong>and</strong> geochronological data that can indicate the existence<strong>of</strong> distinct tectonometamorphic episodes.Polymetamorphic domains are commonly studiedusing analysis <strong>of</strong> polyphase deformations <strong>and</strong> individualtectonic events are attributed to distinctdeformational phases (Turner & Weiss, 1963). Infavourable situations tectonic regimes responsible forthe formation <strong>of</strong> two distinct deformational phases canbe distinguished. However, polyphase structures commonlyresult from the continuous activation <strong>of</strong> localmechanical instabilities during a single deformationevent (Burg, 1999).The tools <strong>of</strong> metamorphic petrology are able toprovide P–T estimates <strong>of</strong> peak metamorphic conditions,<strong>and</strong> important fragments <strong>of</strong> P–T paths can bereconstructed when thermodynamic <strong>modelling</strong> isapplied (Powell & Holl<strong>and</strong>, 1988). However, metamorphic<strong>and</strong> phase petrology do not reveal informationabout the duration <strong>of</strong> metamorphic events.The time span between individual metamorphicdeformationevents <strong>and</strong> the duration <strong>of</strong> metamorphicreworking can be determined in principle bygeochronology.When metamorphic rocks are reworked by a followingtectonometamorphic event after a significantperiod <strong>of</strong> time but under similar P–T conditions, it isunlikely that the duration <strong>of</strong> events, strain rates <strong>and</strong>kinematics <strong>of</strong> deformation are also the same. However,the texture <strong>of</strong> metamorphic tectonites results fromnucleation, grain growth <strong>and</strong> various recrystallizationmechanisms that are strongly controlled by temperature,time <strong>and</strong> strain rate/stress ratio (Hickey & Bell,1996). Therefore, the analysis <strong>of</strong> metamorphic texturesis a method that is capable <strong>of</strong> distinguishing betweentwo metamorphic events under similar P–T conditions<strong>of</strong> different durations.Here, we consider an example <strong>of</strong> Cambro-Ordovicianmetamorphism associated with crustal thinningÓ 2005 Blackwell Publishing Ltd 649231


650 O. LEXA ET AL.followed by moderate crustal thickening <strong>and</strong> magmaunderplating during the Carboniferous (Variscan)orogeny (Sˇtı´pska´ et al., 2001). The results <strong>of</strong> a <strong>structural</strong>,petrological <strong>and</strong> geochronological study allow apreliminary distinction <strong>of</strong> two metamorphic events,which exhibit similar P–T conditions. The <strong>structural</strong><strong>and</strong> geochronological arguments presented here arenot sufficiently unambiguous to distinguish betweenmetamorphic fabrics developed during the Cambro-Ordovician thinning <strong>and</strong> the Carboniferous moderatethickening. Therefore, quantitative textural analysis <strong>of</strong>pairs <strong>of</strong> rock samples with similar mineralogical compositionshas been used to show the influence <strong>of</strong> theCambro-Ordovician <strong>and</strong> Carboniferous events on finaltextures.GEOLOGICAL SETTINGThe Stare´ Meˇsto belt was a Cambro-Ordovician rift,separating high-grade gneisses <strong>of</strong> a thickened continentalcrust (Orlica-Snieznik dome) in the west from theNeo-Proterozoic continental margin (Silesian domain)in the east (Fig. 1), at the eastern margin <strong>of</strong> theBohemian Massif (Kro¨ ner et al., 2000a,b). The Stare´Meˇsto belt consists <strong>of</strong> a lower crustal complex whichwas later affected by the Variscan collisional tectonics(Kro¨ ner et al., 2000a,b; Sˇtípska´ et al., 2001). Theoverall structure <strong>of</strong> this unit is marked by NE–SWtrending lithologies generally dipping to the west(Figs 2 & 3). The top <strong>of</strong> the tectonic sequence <strong>and</strong> theboundary with the rocks <strong>of</strong> the Orlica-Snieznik dome isrepresented by a layer <strong>of</strong> strongly sheared metagabbro(504.9 ± 1.0 Ma; Kro¨ ner et al., 2000a,b). Thisboundary represents a ductile shear zone along which aVariscan granodiorite sill was emplaced (Parry et al.,1997). Zircon from the northern <strong>and</strong> southern parts <strong>of</strong>the granodiorite sill were dated at 339.4 ± 1.1 <strong>and</strong>344.5 ± 0.4 Ma, respectively, reflecting the time <strong>of</strong>magma crystallization. The b<strong>and</strong>ed amphibolite complexlocated <strong>structural</strong>ly below comprises a layeredsequence <strong>of</strong> alternating b<strong>and</strong>ed amphibolites <strong>and</strong> finegrainedquartz<strong>of</strong>eldspathic rocks (503 ± 2 <strong>and</strong>502.1 ± 1.7 Ma; Kro¨ ner et al., 2000a,b), subordinatetonalitic gneisses (503.3 ± 0.8 <strong>and</strong> 501.9 ± 0.6 Ma;Kro¨ ner et al., 2000a,b) <strong>and</strong> high-grade metasedimentsshowing evidence <strong>of</strong> anatexis (age <strong>of</strong> metamorphism –c. 504 Ma; Kro¨ ner et al., 2000a,b). In the northernpart <strong>of</strong> the Stare´ Meˇsto belt the mylonitic metagabbrosare tectonically repeated underneath the b<strong>and</strong>edamphibolite complex (Fig. 1).(a)(b)Fig. 1. (a) The schematic outlines <strong>of</strong> the major units <strong>of</strong> the Bohemian Massif. The location <strong>of</strong> the studied area is indicated. Upper leftinset shows the position <strong>of</strong> the studied area in the frame <strong>of</strong> European Variscides. (b) Geological map <strong>of</strong> the Stare´ Město belt based ongeological maps 1:25 000 provided by courtesy <strong>of</strong> the Czech Geological Survey. Important thrust faults <strong>and</strong> normal faults areindicated. Frames mark location <strong>of</strong> maps shown in Fig. 2.Ó 2005 Blackwell Publishing Ltd232


CONTRASTING TEXTURAL RECORD OF TWO METAMORPHIC EVENTS 651(a)(b)Fig. 2. Geological <strong>and</strong> <strong>structural</strong> maps <strong>of</strong> the Stare´ Město belt. Foliations, mineral lineations <strong>and</strong> major thrust <strong>and</strong> normal faults areindicated. (a) Northern area, see Fig. 1 for location. A–A¢ is the cross-section shown in Fig. 4. (b) Southern area, see Fig. 1 forlocation. B–B¢ is the cross-section shown in Fig. 3.STRUCTURAL CHARACTERIZATION OFVARISCAN AND CAMBRO-ORDOVICIANFABRICSStructural relationships between the granodiorite sill,mylonitic metagabbros <strong>and</strong> b<strong>and</strong>ed amphibolitecomplex combined with geochronological data byKro¨ ner et al. (2000a,b) allowed Sˇtípska´ et al. (2001)to distinguish Cambro-Ordovician <strong>and</strong> Variscantectonometamorphic events in the study area. Thishypothesis was based on <strong>structural</strong> relationshipsbetween the b<strong>and</strong>ed amphibolite complex <strong>and</strong> thesuperposed granodiorite intrusion in the northernpart <strong>of</strong> the Stare´ Město belt (Sˇtı´pska´ et al., 2001). Inthe south, however, the presumed Variscan <strong>and</strong>Cambro-Ordovician fabrics are concordant, <strong>and</strong> arealmost indistinguishable from each other in the field(Fig. 3b).Ó 2005 Blackwell Publishing Ltd233


652 O. LEXA ET AL.(a)(b)Fig. 3. Geological cross-sections A–A¢ <strong>and</strong> B–B¢ (shown in Fig. 2) <strong>of</strong> the Stare´ Město belt disclosing major structures, lithology <strong>of</strong> individual units <strong>and</strong> major tectonic boundaries.Equal-area, lower-hemisphere stereoplots show D 1 ,D 2 <strong>and</strong> D 3 planar <strong>and</strong> linear structures. Each stereoplot is contoured at regular multiples <strong>of</strong> distribution. Drawings are madeaccording to field photographs <strong>and</strong> field notes <strong>and</strong> show principal <strong>structural</strong> features.Ó 2005 Blackwell Publishing Ltd234


CONTRASTING TEXTURAL RECORD OF TWO METAMORPHIC EVENTS 653Structures in Variscan granodioriteThe Variscan granodiorite sill is syntectonicallyemplaced between the mylonitic metagabbro to thewest, <strong>and</strong> the b<strong>and</strong>ed amphibolite complex to the east(Parry et al., 1997). The <strong>structural</strong> pattern <strong>of</strong> thegranodiorite sill varies from SW to NE along thelength <strong>of</strong> the intrusion (Figs 2 & 3). The magmatic tosubsolidus foliation (S 2 ) dips to the NW <strong>and</strong> bears anintense subhorizontal mineral lineation (L 2 ) defined byan alignment <strong>of</strong> amphibole (Fig. 3a,b). Rare lock-upshear-b<strong>and</strong>s filled with residual melt indicate a dextralsense <strong>of</strong> shear. In the south, the granodiorite does notform a continuous body but occurs as numerous sills insurrounding amphibolites <strong>and</strong> anatectic metasediments(Fig. 3b). The granodiorite locally forms dykesemplaced along conjugate steep brittle–ductile NW–SEor NE–SW trending shear zones at high angle to theinitially flat-lying S 1 foliation (Fig. 3a,b).Structures <strong>of</strong> mylonitic metagabbrosAlthough the western metagabbro belt was stronglyreworked during the Variscan orogeny, magmaticstructures are preserved in the low-strain domains(Fig. 4c). Here, the metagabbro exhibits either amedium-grained isotropic texture <strong>and</strong> homogeneouscomposition, or layering marked by alternating layers<strong>of</strong> variable grain size ranging from a few mm up to 2 cm(Fig. 4c). The mylonitic metagabbros exhibit thedevelopment <strong>of</strong> S 2 mylonitic foliation, defined byalternating monomineralic ribbons <strong>of</strong> recrystallizedplagioclase <strong>and</strong> amphibole (Figs 4c & 5c). This foliationdips at medium to high angles to the WNW.Towards the south, the foliation tends to becomesubvertical (Fig. 2b). A strong mineral L 2 lineation isdefined by recrystallized aggregates <strong>of</strong> amphibole <strong>and</strong> isassociated with numerous kinematic indicators suggestingdextral shearing. The S 2 foliation is locallyaffected by late F 3 folds with hinges sub-parallel to L 2lineation or by development <strong>of</strong> S 3 crenulation cleavage.The complete <strong>structural</strong> <strong>and</strong> kinematic coherency betweenthe metagabbros <strong>and</strong> underlying granodiorite sillsuggests their common Variscan deformation history.The eastern metagabbro was affected by heterogeneousshear zones (Fig. 4d) during the Variscan deformation.The metagabbros are strongly mylonitized atthe upper <strong>and</strong> lower contacts with the hangingwall <strong>and</strong>Fig. 4. Field photographs showing main <strong>structural</strong> features <strong>of</strong> important lithologies <strong>of</strong> the Stare´ Meˇsto belt. (a) Metamorphic layeringM 1 <strong>of</strong> b<strong>and</strong>ed amphibolite complex. (b) Extensional melt-filled shear-b<strong>and</strong>s D 1 in tonalitic migmatitic gneiss <strong>of</strong> the b<strong>and</strong>ed amphibolitecomplex. (c) Coarse-grained C-O metagabbro (left photograph) mylonitized during D 2 (right photograph) from western belt. (d)Coarse-grained C-O metagabbro mylonitized by localized shear zones during D 2 from eastern belt.Ó 2005 Blackwell Publishing Ltd235


654 O. LEXA ET AL.(a)(b)(c)(d)Fig. 5. Photomicrographs <strong>of</strong> characteristic mineral assemblages <strong>and</strong> textures from the Stare´ Meˇsto belt: (a) M 1 assemblage Cpx–Amp–Pl–Qtz <strong>of</strong> layered amphibolite showing equilibrated annealed texture. (b) M 1 assemblage Grt–Bt–Sil in anatectic paragneiss. (c)Variscan dynamic M 2 recrystallization <strong>of</strong> C-O metagabbro from the western belt leading to strong mineral b<strong>and</strong>ing <strong>of</strong> amphibole <strong>and</strong>plagioclase. (d) Variscan dynamic M 2 recrystallization <strong>of</strong> amphibole <strong>and</strong> plagioclase in C-O metagabbro from the eastern belt markedby strong cataclastic grain-size reduction.footwall metapelites, respectively, while magmaticstructures are preserved in the inner part. The geometry<strong>of</strong> S 2 fabric <strong>and</strong> kinematics <strong>of</strong> deformation are identicalwith those <strong>of</strong> the western metagabbro sheet (Fig. 3).Structures <strong>of</strong> the b<strong>and</strong>ed amphibolite complexThe b<strong>and</strong>ed amphibolite complex is affected by a hightemperaturemetamorphic event characterized by arange <strong>of</strong> melt-collecting structures. The S 1 foliation ismarked by a metamorphic compositional layering bestseen in the b<strong>and</strong>ed amphibolites <strong>and</strong> in the tonaliticmigmatites (Figs 3a & 4a). In the northern part <strong>of</strong> thestudy area, S 1 dips to the NNE at shallow angles(Figs 2a & 3a). To the south, the S 1 moderately dips tothe west in the migmatitic metasediments <strong>and</strong> is generallyflat-lying in the b<strong>and</strong>ed amphibolites (Figs 2b &3b). The S 1 compositional <strong>and</strong> metamorphic layering islocally folded into rootless isoclinal F 1 folds or isaffected by asymmetrical pinch <strong>and</strong> swell boudinagewith melt collecting in some neck-zones. In the tonaliticgneisses <strong>and</strong> more rarely in the layered amphibolitesthe planar fabric is affected by extensional shearzones filled with an amphibole-bearing tonalitic melt,or the planar fabric dissipates in coarse-grained meltpatches (Fig. 4b).The Variscan D 2 deformation in the b<strong>and</strong>edamphibolite complex is marked by folding <strong>of</strong> theoriginally flat-lying Cambro-Ordovician S 1 foliation.These open to close F 2 folds strike NE–SW with subhorizontalhinges <strong>and</strong> wavelengths ranging from a fewmetres to several tens <strong>of</strong> metres (Fig. 3). Fold hingesare locally cut by subvertical, NE–SW or NW–SEtrending, conjugated shear <strong>and</strong> fracture zones filledwith granitic melts. The internal foliation within theseshear zones bears a NE–SW trending stretching L 2lineation (Fig. 3). Similar to mylonitic metagabbros,older foliations are locally affected by late F 3 foldswith hinges sub-parallel to the L 2 lineation or bydevelopment <strong>of</strong> a late S 3 crenulation cleavage subparallelto the S 2 foliation, indicating deformationcontinuity towards lower metamorphic conditions.Ó 2005 Blackwell Publishing Ltd236


CONTRASTING TEXTURAL RECORD OF TWO METAMORPHIC EVENTS 655Table 1. Mineral assemblages <strong>and</strong> textures <strong>of</strong> samples used for thermobarometry.Sample Unit Locality Rock type Texture Grt Pl Hbl Cum Cpx Qtz Bt Kfs Sil Rt Ilm Ttn Mag OpxSamples associated with D1 Cambro-Ordovician structureLAC4a LAC 66 g Grt-amphibolite Preferred shape orientation,· · · – – · – – – – · – · –annealedLAC4b LAC S149e Cpx-amphibolite Weak shape preferred orientation, – · · – · · – – – – · – · –annealedLAC4c LAC S149f Melt patches within R<strong>and</strong>om orientation – · · – – · – – – – – – – –amphiboliteTG1a LAC S24-1b Tonalitic migmatite Weak shape preferred orientation, · · · – – · · – – – – – – –annealedTG1b LAC S24-5a Tonalitic migmatite Weak shape preferred orientation, · · · – – · · – – – – – – –annealedTG1c LAC S24-6e Melt patches within R<strong>and</strong>om orientation – · · – – · · · – – – · · –migmatiteMP1 LAC Heg3 Metapelite Weak shape preferredorientation, annealed· · – – – · · · · · · – – –Samples associated with D2 Variscan structureGAHT Western S33b Grt-amphibolite Strong preferred orientation, dynamic · · · · – · – – – – · – · –gabbroic beltrecrystalliation <strong>of</strong> Pl, QtzGHT3 Western S151 Metagabbro Strong preferred orientation <strong>of</strong> Amp, – · · · · – – – – – · · – ·gabbroic beltdynamic recrystalliation <strong>of</strong> Pl, AmpGLT2 Eastern S130 Metagabbro Strong preferred orientation <strong>of</strong> Amp, – · · – – – – – – – · · – –gabbroic beltdynamic recrystalliation <strong>of</strong> Pl, AmpT3 Sill Granodiorite Dynamic recrystallization <strong>of</strong> Pl, Qtz – · · – – · · · – – · · · –PETROLOGY, METAMORPHIC TEXTURES ANDP–T ESTIMATESSample selection for petrological study <strong>and</strong> analyticalproceduresThe study samples were selected according to their<strong>structural</strong> position with respect to assumed Cambro-Ordovician <strong>and</strong> Variscan fabrics. Mineral <strong>analyses</strong>were carried out on a CAMECA SX 50 at ETH Zu¨ rich<strong>and</strong> a JEOL microprobe at the University <strong>of</strong> Mainz.Operating conditions were 15 kV acceleration voltage<strong>and</strong> beam current <strong>of</strong> 20 nA. Representative sampleswith mineral assemblages are given in Table 1 <strong>and</strong>corresponding mineral compositions are summarizedin Table 2. Mineral abbreviations used in text <strong>and</strong>tables follow Kretz (1983). Representative mineral<strong>analyses</strong> are listed in Tables 3–6. Amphibole formulaewere calculated after Holl<strong>and</strong> & Blundy (1994) <strong>and</strong>classified according to Leake et al. (1997). The geothermometers<strong>and</strong> geobarometers used for calculationsare given in Table 7. For each sample, five to 10 sets <strong>of</strong><strong>analyses</strong> were used for P–T calculations <strong>and</strong> the resultsare shown in Table 7 <strong>and</strong> Fig. 6.Metamorphic textures <strong>and</strong> mineral compositions <strong>of</strong> thegranodiorite sill <strong>and</strong> mylonitic metagabbrosThe granodiorite sill is composed <strong>of</strong> Pl + Qtz +Amp + Bt + Kfs + Ttn + Mag + Ilm + Ap. Itshows magmatic, sub-magmatic <strong>and</strong> solid-statemicrostructures marked by dynamic recrystallization<strong>of</strong> plagioclase, amphibole <strong>and</strong> quartz (Parry et al.,1997). Amphibole correspond to magnesiohornblende<strong>and</strong> tschermakite (i.e. Si p.f.u. ¼ 6.3–6.7, X Mg ¼ 0.54–0.64) <strong>and</strong> plagioclase is An 34)37 <strong>and</strong>esine.The proportions <strong>of</strong> plagioclase <strong>and</strong> amphibole in thewestern mylonitic metagabbro vary significantly <strong>and</strong>the rock locally consists <strong>of</strong> up to 90% plagioclase or90% amphibole. Minor clinopyroxene occurs in thecore <strong>of</strong> large amphibole. Titanite is an abundantaccessory mineral. Non-recrystallized amphibole isinterpreted as magmatic in origin <strong>and</strong> the compositioncorresponds to a magnesiohornblende (i.e. Si p.f.u. ¼7.00–7.25, X Mg ¼ 0.84–0.89). The magmatic plagioclaseis a An 55)62 labradorite.The mineral assemblage <strong>of</strong> highly reworked metagabbrowith b<strong>and</strong>ed mylonitic structure (Figs 4c & 5c)includes Pl + Amp ± Cum ± Cpx ± Opx ± Grt ±Ttn ± Rt ± Ilm. A granulite facies mineral assemblage,comprising saphirine <strong>and</strong> corundum, isre-equilibrated during subsequent amphibolite faciesreworking (Baratoux et al., 2005). Both amphibole <strong>and</strong>plagioclase show strong compositional variationsbetween those <strong>of</strong> old magmatic <strong>and</strong> metamorphicgrains (Table 1). Rare garnet-bearing amphibolitedeveloped through the deformation <strong>and</strong> metamorphism<strong>of</strong> a tonalitic migmatitic gneiss associatedwith metagabbro in the lower part <strong>of</strong> the gabbroicsheet. It exhibits mineral assemblage Hbl ±Cum ± Grt + Pl + Qtz ± Mag ± Ilm (Table 1)which allows the pressure during the Variscan metamorphismto be estimated.Recrystallized plagioclase exhibits an increase in theanorthite content from the core towards the rim inthe metagabbro (An 37 fi 60 ) <strong>and</strong> garnet-amphibolite(An 15 fi 27 ). Recrystallized amphibole in the metagabbrocorresponds to tschermakite, <strong>and</strong> tschermakite orferrotschermakite in the garnet-amphibolite. Garnetshows an increase in alm<strong>and</strong>ine, grossular, pyrope <strong>and</strong>X Mg (Fig. 7d), accompanied by depletion <strong>of</strong> spessar-Ó 2005 Blackwell Publishing Ltd237


656 O. LEXA ET AL.Table 2. Summarized compositions <strong>of</strong> minerals used for thermobarometry.Sample Rock type Locality Grt Pl Amp Cpx Bt IlmXMg Alm Py Grs Sps An Compositional name a XMg Si (pfu) XMg XMg XIlmSamples associated with D1 Cambro-Ordovician structureLAC4a Grt-amphibolite 66 g 0.16–0.18 62–65 12–14 11–21 4–5 31–33 Ts, Fe-Ts 0.47–0.54 6.27–6.37 – – –LAC4b Cpx-amphibolite S149c – – – – – 44–50 Mg-Hbl, Ts, Ed, Prg 0.59–0.68 6.36–6.71 0.67 b – –LAC4c Melt patche within amphibolite S149f – – – – – 38–43 Ts 0.63–0.74 6.25–6.50 – – –TG1a Tonalitic migmatite S24-1b 0.10–0.11 55 (51) 7 (8) 28 (29) 6 (13) 33–35 Fe-Prg, Hs 0.34–0.37 5.8–6.2 – 0.34–0.36 –TG1b Tonalitic migmatite S24-5a 0.11–0.12 58 (50) 8 (6) 25 (31) 9 (14) 31–34 Fe-Prg, Hs 0.25–0.40 5.8–6.2 – 0.33–0.35 –TG1c Melt patche within migmatite S24-6e – – – – – 32–35 Fe-Prg, Hs 0.34–0.38 5.95–6.25 – 0.32–0.33 –MP1 Metapelite Heg3 0.15 (0.26–0.28) b 75 (69) 15 (26) 4 (4) 5 (1) 24–26 – – – – 0.50–0.53 0.95–0.98– – –6.1–6.37.7–7.96.0–6.35.30.46–0.670.48–0.500.82–0.880.64Samples associated with D2 Variscan structureGAHT Grt-amphibolite S33b 0.15 (0.05) 70 (57) 12 (3) 15 (27) 3 (12) 15–27 (15) Fe-Ts, TsCum– – –GHT3 Metagabbro S151 – – – – – 37–60 Ts,CumGLT2 Metagabbro S130 – – – – – 48–62 (25) Mg-Hbl, Ts 0.74–0.88 6.24–6.64 – – –T3 Granodiorite – – – – – 34–37 Mg-Hbl, Ts 0.54–0.64 6.3–6.7 – – –Values in brackets represent core compositions <strong>of</strong> zoned minerals. Notes: Alm ¼ 100 · Fe 2+ /(Mg + Ca + Mn + Fe 2+ ), An ¼ 100 · Ca/(Ca + Na + K), XMg ¼ Mg/(Mg + Fe 2+ ).a Amphibole compositional name from classification <strong>of</strong> Leake et al. (1997): Prg, pargasite; Ts, tschermakite; Ed, edenite; Hs, hastingsite; Tr, tremolite; Mg-Hbl, magnesio-hornblende.b XMg ¼ Mg/(Mg + Fe tot ).tine from the core to the rim (i.e. Alm 57 fi 70Grs 2 fi 15 Py 3 fi 12 Sps 12 fi 3 ; X Mg ¼ 0.05 fi 0.15).A relic magmatic assemblage in the eastern myloniticmetagabbro is represented by Pl + Hbl ±Cpx ± Ttn ± Ilm. Magmatic plagioclase (0.5–5 mm)shows zoning marked by a rimward increase <strong>of</strong>anorthite from 50 to 60%, whereas recrystallization <strong>of</strong>the plagioclase (0.05–0.1 mm) is accompanied by adecrease in anorthite content (i.e. An 45 ). Magmaticamphibole (0.5–5 mm) with pyroxene relics in thecores is magnesiohornblende (6.8–7.0 Si p.f.u., X Mg ¼0.72–0.77). Recrystallized grains (0.02–0.1 mm) correspondto magnesiohornblende with a slightly lowercontent <strong>of</strong> Si <strong>and</strong> X Mg (6.6–6.7 Si p.f.u., X Mg ¼ 0.69–0.74) in comparison with original grains.P–T conditions <strong>of</strong> Variscan metamorphismThe pressure <strong>of</strong> crystallization <strong>of</strong> the granodiorite (T3)was calculated as 6.5–7.0 kbar using the Al-content <strong>of</strong>amphibole (Schmidt, 1992) at a temperature <strong>of</strong> 670–725 °C inferred from the Pl–Hbl thermometer <strong>of</strong>Holl<strong>and</strong> & Blundy (1994). In the western mylonitemetagabbro the temperature was estimated as770 ± 50 °C using the thermometer by Holl<strong>and</strong> &Blundy (1994) in metagabbro (GHT3), while in garnetamphibolitethe temperature was estimated as 710–740 °C (GAHT). A similar temperature <strong>of</strong> 750 ±50 °C is reported by Baratoux et al. (2005) for 40amphibole–plagioclase couples. Pressure conditionsbetween 8 <strong>and</strong> 10 kbar were calculated only in Qtzbearinggarnet-amphibolite by means <strong>of</strong> the Kohn &Spear (1990) barometer. Based on 30 amphibole–plagioclase couples from the eastern mylonite metagabbro,Baratoux et al. (2005) estimated temperatureas 650 ± 50 °C using the Pl–Hbl thermometer <strong>of</strong>Holl<strong>and</strong> & Blundy (1994). The pressure could not becalculated because <strong>of</strong> the lack <strong>of</strong> garnet.Mineral textures <strong>and</strong> mineral compositions <strong>of</strong> rocks <strong>of</strong> theb<strong>and</strong>ed amphibolite complexMineral assemblages <strong>and</strong> selected mineral compositions<strong>of</strong> all study rocks are given in Table 1 <strong>and</strong> summarizedin Table 2, respectively.B<strong>and</strong>ed amphibolites are characterized by the mineralassemblage Amp + Pl + Qtz ± Cpx ± Grt ±Mag ± Ilm. Three types <strong>of</strong> microstructures marked bystraightened grain boundaries without indications <strong>of</strong>dynamic recrystallization are distinguished: (1) Equigranularaggregates (grain size 0.05–1.0 mm) <strong>of</strong>amphibole, plagioclase <strong>and</strong> quartz (Fig. 5a); (2)amphibole-rich layers alternating with elongateaggregates (grain size 0.05–1.0 mm) <strong>of</strong> quartz <strong>and</strong>plagioclase; <strong>and</strong> (3) coarse-grained aggregates (grainsize 1–5 mm) <strong>of</strong> r<strong>and</strong>omly distributed plagioclase<strong>and</strong> quartz sometimes with amphibole occur in meltpatches. In types (1) <strong>and</strong> (2) the mineral grains arealigned <strong>and</strong> elongated parallel to the compositionalÓ 2005 Blackwell Publishing Ltd238


CONTRASTING TEXTURAL RECORD OF TWO METAMORPHIC EVENTS 657Table 3. Representative electron probe <strong>analyses</strong> <strong>of</strong> garnet.Garnet <strong>analyses</strong> recalculated to 12 oxygenEvent C-O Varis.Sample/<strong>analyses</strong> TG1a/grt23-rim TG1a/grt44-core TG1b/grt27-rim LAC4a/grt43-rim MP1/grt379-core a MP1/grt313-rim a GAHT/grt5-15-core GAHT/grt114-rimSiO 2 37.96 37.65 37.44 37.95 37.69 36.79 37.14 37.47TiO2 0.06 0.13 0.14 0.07 0.08 0.08 0.07 0.04Al 2 O 3 20.64 20.53 20.77 20.46 21.21 20.82 21.17 20.98FeO 26.47 24.70 27.05 30.02 32.60 35.12 25.97 32.46MnO 4.28 5.78 5.19 1.84 0.56 2.25 6.15 1.25MgO 1.80 1.54 1.75 3.06 6.54 3.62 0.85 3.12CaO 10.05 10.28 8.31 7.34 1.33 1.33 9.92 5.28Total 101.26 100.61 100.65 100.74 100.01 100.01 101.27 100.60Si 2.99 2.98 2.98 3.00 2.97 2.97 2.94 2.97Ti 0.00 0.08 0.01 0.00 0.00 0.00 0.00 0.00Al 1.92 1.92 1.95 1.90 1.97 1.98 1.98 1.96Fe 3+ 0.10 0.11 0.09 0.09 0.00 0.00 0.13 0.09Fe 2+ 1.65 1.53 1.71 1.89 2.15 2.37 1.59 2.07Mn 0.29 0.39 0.35 0.12 0.04 0.15 0.41 0.08Mg 0.21 0.18 0.21 0.36 0.77 0.44 0.10 0.37Ca 0.85 0.87 0.71 0.62 0.11 0.12 0.84 0.45a Fe 3+ not calculated.Table 4. Representative electron probe <strong>analyses</strong> <strong>of</strong> amphibole.Amphibole <strong>analyses</strong> recalculated after Holl<strong>and</strong> & Blundy (1994)Event C-O Varis.Sample/<strong>analyses</strong> LAC4a/amp37 LAC4b/amp16 LAC4c/amp482 TG1a/amp45 TG1b/amp12-6 TG1c/amp64 GAHT/amp115 GAHT/cum6-7 a GHT3/amp48 GLT2/amp-102 T3/amp18SiO 2 42.06 45.27 42.58 39.82 39.12 39.38 41.64 51.85 43.34 45.11 43.59TiO 2 1.59 1.47 1.75 0.91 1.05 0.94 0.53 0.05 0.20 0.61 1.26Al2O3 11.58 9.68 11.73 12.96 13.51 13.40 12.42 1.35 17.66 13.94 11.85FeO 20.83 15.88 17.00 23.31 23.52 23.46 22.17 27.34 11.03 9.82 16.03MnO 0.12 0.32 0.24 0.57 0.54 0.46 0.46 0.59 0.13 0.18 0.24MgO 7.87 11.69 10.54 5.66 5.45 5.55 7.28 13.95 12.09 13.98 10.89CaO 10.62 11.34 10.68 11.13 11.41 11.31 10.57 2.51 11.29 12.42 11.05Na2O 2.11 1.50 1.86 1.45 1.33 1.39 1.77 0.18 1.95 1.71 1.25K 2 O 0.07 0.55 0.78 1.47 1.65 1.63 0.23 0.01 0.27 0.32 1.48Total 96.85 97.70 97.16 97.28 97.58 97.52 97.07 97.83 97.95 98.09 97.64Si 6.38 6.65 6.34 6.16 6.05 6.09 6.30 7.75 6.15 6.43 6.44Ti 0.18 0.16 0.20 0.11 0.12 0.11 0.06 0.01 0.02 0.07 0.14Al 2.07 1.68 2.06 2.36 2.46 2.44 2.22 0.24 2.95 2.34 2.07Fe 3+ 0.58 0.50 0.61 0.66 0.70 0.67 0.91 0.19 0.70 0.34 0.51Fe 2+ 2.06 1.45 1.50 2.36 2.34 2.36 1.90 3.23 0.61 0.83 1.47Mn 0.02 0.04 0.03 0.08 0.07 0.06 0.06 0.08 0.02 0.02 0.03Mg 1.78 2.56 2.34 1.30 1.26 1.28 1.64 3.11 2.56 2.97 2.40Ca 1.73 1.79 1.70 1.84 1.89 1.87 1.71 0.40 1.72 1.90 1.75Na 0.62 0.43 0.54 0.44 0.40 0.42 0.52 0.05 0.54 0.47 0.36K 0.01 0.10 0.15 0.29 0.33 0.32 0.04 0.00 0.05 0.06 0.28a Cummingtonite recalculated on the basis <strong>of</strong> 23 oxygen <strong>and</strong> 15 cations + Na + K.layering. Regardless <strong>of</strong> textural type, the minerals arenot zoned. Alm<strong>and</strong>ine-rich garnet reveals flat chemicalpr<strong>of</strong>iles (Alm 62)65 Grs 11)21 Py 12)14 Sps 4)5 ; X Mg ¼ 0.16–0.18; Fig. 7a). Amphibole is magnesiohornblende ortschermakite, rarely edenite or pargasite. Plagioclase is<strong>and</strong>esine (An 30)50 ) <strong>and</strong> X Mg <strong>of</strong> rare clinopyroxeneequals 0.65–0.67.Tonalitic migmatitic gneiss is made up <strong>of</strong>Qtz + Pl + Bt + Amp ± Grt. Its structure is characterizedby alternations <strong>of</strong> leucosome, mesosome <strong>and</strong>melanosome layers several mm to 1 cm in thickness.Weak layering is defined by amphibole–plagioclaseaggregates alternating with quartz-rich domains(Fig. 4b). Plagioclase <strong>and</strong> amphibole are elongated(length 0.2–3 mm) in the leucosome <strong>and</strong> restitic layers,whereas plagioclase is isometric (up to 3 mm). Biotite(0.2–1 mm) is oriented parallel to the foliation togetherwith amphibole in melanosome. Garnet porphyroblasts(0.3 mm) occur in mesosome <strong>and</strong> melanosome. Largecoarse-grained quartz aggregates occurs in leucosome.Chemical pr<strong>of</strong>iles across garnet show slight zoning withcores being more spessartine-rich; the X Mg ratio is constant(0.10–0.12) (Fig. 7c). Amphibole is homogeneousferropargasite or hastingsite, regardless <strong>of</strong> the texturalposition <strong>and</strong> type <strong>of</strong> aggregate. X Mg in biotite rangesbetween 0.32 <strong>and</strong> 0.36, <strong>and</strong> the Ti content between 0.14<strong>and</strong> 0.22 p.f.u. The composition <strong>of</strong> plagioclase in boththe leucosome <strong>and</strong> melt patches is An 31)35 <strong>and</strong>esine.Ó 2005 Blackwell Publishing Ltd239


658 O. LEXA ET AL.Table 5. Representative electron probe <strong>analyses</strong> <strong>of</strong> plagioclase.Event C-O VarisSample/<strong>analyses</strong>LAC4b/pl13LAC4c/pl82-1LAC4a/pl50TG1a/pl46-5TG1b/pl26TG1c/pl2-8MP1/pl137GAHT/pl113 -rimGHT3/pl122-coreGHT3/pl101-rimGLT2/pl122-coreGLT2/pl101-rimGLT2/pl79-coreT3/pl19SiO2 56.92 58.57 60.07 59.94 60.11 59.82 61.64 62.51 59.02 52.36 64.73 56.05 62.00 59.35TiO 2 0.00 0.03 0.00 0.01 0.01 0.00 0.01 0.03 0.00 0.00 0.00 0.00 0.01 0.01Al2O3 27.18 26.20 25.01 25.54 25.26 25.09 24.01 23.64 25.92 30.48 22.01 27.73 23.42 26.19FeO 0.22 0.13 0.35 0.05 0.17 0.17 0.00 0.35 0.00 0.00 0.07 0.18 0.00 0.05CaO 9.74 8.53 6.60 7.09 6.99 7.31 5.28 4.95 7.17 12.18 3.27 10.09 5.26 7.02Na2O 6.30 6.73 8.16 7.65 7.73 7.44 8.76 8.77 7.63 4.50 9.70 5.93 8.66 7.59K 2 O 0.12 0.30 0.05 0.18 0.12 0.35 0.20 0.07 0.00 0.00 0.15 0.03 0.12 0.15Total 100.48 100.49 100.19 100.46 100.39 100.18 99.90 100.32 99.74 99.52 99.93 100.01 99.47 100.36Si 2.54 2.61 2.66 2.66 2.67 2.66 2.73 2.76 2.64 2.37 2.86 2.52 2.76 2.63Ti 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00Al 1.43 1.38 0.00 1.34 1.32 1.32 1.25 1.23 1.36 1.63 1.15 1.47 1.23 1.37Fe 3+ 0.01 0.00 0.01 0.00 0.01 0.01 0.00 0.00 0.00 0.00 0.00 0.01 0.00 0.00Ca 0.47 0.41 0.31 0.34 0.33 0.35 0.25 0.23 0.34 0.59 0.15 0.49 0.25 0.33Na 0.55 0.58 0.70 0.66 0.66 0.64 0.75 0.75 0.66 0.40 0.83 0.52 0.75 0.65K 0.01 0.02 0.00 0.01 0.01 0.02 0.01 0.00 0.00 0.00 0.01 0.00 0.01 0.01Table 6. Representative electron probe <strong>analyses</strong> <strong>of</strong> biotite <strong>and</strong>clinopyroxene.Bt <strong>analyses</strong> recalculated to 11 oxygen, cpx to 6 oxygenEventThe mineral assemblage <strong>of</strong> the migmatitic paragneissesis Pl + Bt + Qtz ± Grt ± Kfs ± Sil ± Rt ±Ilm. Plagioclase <strong>and</strong> quartz form large isometric grainsup to 10 mm in size. With an increasing degree <strong>of</strong>anatexis, quartz <strong>and</strong> plagioclase (0.3–5 mm) coalesceto form lenses bordered by biotite-rich rims occasionallycontaining sillimanite. Both the leucosome <strong>and</strong>melanosome aggregates contain garnet. In thenebulitic stage diffuse pockets <strong>of</strong> leucosome <strong>of</strong> differentsize developed. Garnet forms poikilitic crystals(0.1–1 cm) with inclusions <strong>of</strong> quartz <strong>and</strong> biotite. ScarceK-feldspar occurs in leucosome lenses. The distribution<strong>of</strong> elements in garnet cores is homogeneous(Alm 69 Grs 4 Py 26 Sps 1 ; X Mg ¼ 0.26–0.28) but within theouter 100 lm rim or around biotite inclusions garnet isC-OMineral cpx btSample/<strong>analyses</strong> LAC4b/cpx20 TG1a/bt25 TG1b/bt45 MP1/bt343SiO 2 50.88 33.95 35.10 36.06TiO 2 0.24 2.92 3.15 1.57Al2O3 2.35 15.71 15.76 19.78FeO 11.11 25.22 24.39 18.03MnO 0.35 0.30 0.35 0.05MgO 12.04 7.23 7.37 10.83CaO 22.14 0.02 0.07 0.00Na2O 0.58 0.04 0.05 0.22K 2 O 0.01 9.51 9.40 9.03Total 99.70 94.90 95.64 95.57Si 1.93 2.74 2.80 2.75Ti 0.01 0.18 0.19 0.09Al 0.11 1.49 1.48 1.78Fe 3+ 0.03 0.00 0.00 0.00Fe 2+ 0.32 1.70 1.63 1.15Mn 0.01 0.02 0.02 0.00Mg 0.68 0.87 0.88 1.23Ca 0.90 0.00 0.01 0.00Na 0.04 0.01 0.01 0.03K 0.00 0.98 0.96 0.88enriched with alm<strong>and</strong>ine <strong>and</strong> spessartine <strong>and</strong> depletedin pyrope (Alm 75 Grs 4 Py 26 Sps 5 ; X Mg ¼ 0.15; Fig. 7b).This zoning is attributed to continuous Fe–Mgre-equilibration between garnet <strong>and</strong> biotite duringcooling. The X Mg <strong>of</strong> matrix biotite ranges from 0.50 to0.53 <strong>and</strong> its Ti-content varies between 0.07 <strong>and</strong>0.21 p.f.u. Plagioclase is a homogeneous An 28)31oligoclase-<strong>and</strong>esine.P–T conditions <strong>of</strong> Cambro-Ordovician metamorphismThe temperature <strong>of</strong> metamorphism in amphibolites(LAC4a, LAC4b <strong>and</strong> LAC4c) was estimated with thePl–Hbl thermometer by Holl<strong>and</strong> & Blundy (1994). Thetemperature ranges <strong>of</strong> 725–825 <strong>and</strong> 730–860 °C wereinferred for amphibolites <strong>and</strong> melt patches, respectively.A comparable temperature span <strong>of</strong> 715–770 <strong>and</strong>750–770 °C was established for the tonalitic gneiss<strong>and</strong> associated melt patches, respectively (TG1a, TG1b& TG1c). The pressure in amphibolites was estimatedas 8.5–10 kbar, <strong>and</strong> in tonalitic gneisses as 9–10 kbarusing the Grt–Hbl–Pl–Qtz barometer (Kohn & Spear,1990). A similar range <strong>of</strong> 8–10 kbar was obtained fromthe aluminium content <strong>of</strong> amphibole (Schmidt, 1992)within the tonalitic gneisses. Similarly, a pressure 8–9 kbar was estimated from melt patches crystallized inthe small-scale shear zones. A pressure <strong>of</strong> 7.5 kbar wasobtained from garnet core <strong>and</strong> matrix plagioclase usingthe GASP barometer (Koziol, 1989), calculated bymeans <strong>of</strong> the ÔThermobarometryÕ s<strong>of</strong>tware (Spearet al., 1991). Pressure calculated using the GRAILbarometer (Bohlen et al., 1983) gives a range <strong>of</strong> 7.5–8 kbar for garnet cores containing rutile <strong>and</strong> ilmeniteinclusions.QUANTITATIVE TEXTURAL ANALYSISA quantitative micro<strong>structural</strong> analysis was carried outon three samples <strong>of</strong> the b<strong>and</strong>ed amphibolite complex,assumed to be exclusively the result <strong>of</strong> the Cambro-Ó 2005 Blackwell Publishing Ltd240


CONTRASTING TEXTURAL RECORD OF TWO METAMORPHIC EVENTS 659Table 7. Results <strong>of</strong> P–T calculations using the following geothermometers <strong>and</strong> geobarometers: HB, Holl<strong>and</strong> & Blundy (1994); T, Thompson (1976); FS, Ferry & Spear (1978);NW, O’Neill & Wall (1987); GN, Gasparik & Newton (1984); KS (Mg), Kohn & Spear (1990); KS (Fe), Kohn & Spear (1990); K, Koziol (1989); Bh, Bohlen et al. (1983); S,Schmidt (1992). Errors are given in table headings.Grt–Hbl–Pl–Qtz (KS) Al in Hbl (S) GASP (K) GRAIL (Bh)Pl–Hbl (HB) Grt-Bt (T) Grt-Bt (FS) P calculated at aT (°C) <strong>of</strong>Sample Rock type Combination T calculated at aP (kbar) <strong>of</strong>T (Ed-Tr) T (Ed-Ri) ± 50 °C ± 50 °C P (Mg) P (Fe) – – –±40°C ± 40 °C ± 0.5 ± 0.5 ± 0.6 Not given ± 0.5Samples associated with D1 Cambro-Ordovician structureLAC4a Grt-amphibolite Rims 10 731–770 772–823 – – 700 8.6–9.0 9.5–9.8 – – –LAC4b Cpx-amphibolite Rims 10 727–787 775–825 – – 700 – – – – –LAC4c Melt patches within amphibolite Rims 10 732–765 796–857 – – – – – – – –TG1a Tonalitic migmatite Rims 10 715–755 732–772 700–738 685–737 700 9.7–10.0 9.8–10.1 8.2–10.0 – –TG1b Tonalitic migmatite Rims 10 717–762 712–768 712–727 702–722 700 9.2–9.9 9.5–9.8 8.4–10.0 – –TG1c Melt patches within migmatite Rims 10 750–756 759–770 – – – – – 7.9–8.7 – –MP1 Metapelite Cores 7 – – 788–840 838–905 800 – – – 7.3–7.4 7.6–8.2Rims 7 – – 581–616 549–593 – – – – – –Samples associated with D2 Variscan structureGAHT Grt-amphibolite Rims 10 711–730 724–737 – – 700 7.8–8.4 9.6–10.1 – – –GHT3 Metagabbro Rims 10 716–836 711–837 – – – – – – – –T3 Tonalite Rims 6 671–691 714–724 – – – – – 6.5–7.0 – –Fig. 6. P–T diagram with blank boxes for C-O metamorphism.P–T estimates for Variscan assemblages are presented in greyboxes. Dashed lines show temperature estimates from metagabbros<strong>and</strong> from metaperidotites. Numbers correspond tosamples presented in Tables 1, 2 <strong>and</strong> 7.Ordodvician metamorphism. We investigated also sixsamples representing the Variscan granodiorite, tonalitictectonites <strong>and</strong> mylonitic metagabbros, in which,only the Variscan metamorphic history is assumed tobe preserved. Polished thin sections <strong>of</strong> the samplesparallel to lineation <strong>and</strong> perpendicular to foliation (XZsection) were prepared for optical microscopy <strong>and</strong>back-scattered electron image (BSEI) analysis. TheBSEI analysis was carried out using the Camscan S4instrument fitted with a high-resolution backscatterdetector. Line drawings <strong>of</strong> optical microscopy photographs<strong>and</strong> BSE images were digitized (Fig. 8) <strong>and</strong>processed in ESRI ArcView desktop GIS in order toobtain a map <strong>of</strong> boundaries between individual grains(Lexa, 2003). The analysis <strong>of</strong> crystal size distributions(CSD), shape preferred orientation <strong>of</strong> grains (SPO)<strong>and</strong> grain boundaries preferred orientation (GBPO), aswell as grain contact frequencies were carried out inPolyLX MATLAB TM Toolbox (Lexa, 2003).Crystal size distributionsThe theory <strong>of</strong> CSD is a well-established method inchemical engineering, metallurgy <strong>and</strong> ceramics toreveal information about nucleation, growth rates<strong>and</strong> growth times <strong>of</strong> crystals (R<strong>and</strong>olph & Larson,1971). CSD in many metamorphic <strong>and</strong> igneous rocksshow a loglinear relationship between grain size L<strong>and</strong> population density N according to the followingequationN ¼ N 0 e L=Gt ;ð1ÞÓ 2005 Blackwell Publishing Ltd241


660 O. LEXA ET AL.Fig. 7. Chemical pr<strong>of</strong>iles <strong>of</strong> garnet: (a) inlayered amphibolite showing flat patterns,(b) in sillimanite-bearing paragneiss withwell-developed flat pr<strong>of</strong>iles in core <strong>and</strong>margin affected by late diffusion, (c) intonalitic migmatitic gneiss <strong>and</strong> (d) inGrt-amphibolite within metagabbro.Numbering <strong>of</strong> samples corresponds tonumbers in Tables 1, 2 <strong>and</strong> 7. See text fordiscussion.where N 0 <strong>and</strong> Gt are constants <strong>and</strong> may be related tonucleation density <strong>and</strong> growth rate <strong>of</strong> crystal (Cashman& Ferry, 1988; Marsh, 1988).CSD plots <strong>of</strong> all samples constructed according tothe method by Peterson (1996) exhibit linear correlationsbetween logarithm <strong>of</strong> population density (i.e.number <strong>of</strong> crystals per size per volume) <strong>and</strong> crystalsize (Fig. 9b). Therefore, applying the theory <strong>of</strong>CSD, such distributions could be parameterized byzero size intercept N 0 (nucleation density) <strong>and</strong> slopeGt (growth rate multiplied by time). These twoparameters plotted in a N 0 –Gt space (Fig. 9) exhibitan inverse correlation, <strong>and</strong> the samples with acomparable metamorphic grade <strong>and</strong> deformationhistory form distinct clusters. The metagabbros(samples GLT1 & GLT2, Fig. 8b) sheared during theVariscan episode from the eastern mylonitic belt(c. 650 ± 50 °C) form a group with the highest N 0<strong>and</strong> lowest Gt values, while metagabbros (samplesGHT1 & GHT2) adjacent to the granodiorite sill(c. 750 ° ±50°C) exhibit a decrease in N 0 in conjunctionwith increasing Gt. The samples LAC1,LAC3 <strong>and</strong> TG2 (Fig. 8a,d) <strong>of</strong> b<strong>and</strong>ed amphibolites<strong>and</strong> amphibole-bearing tonalitic gneisses thatoriginated during the Cambro-Ordovician metamorphism(c. 750–850 °C) plot in the centre <strong>of</strong> the diagram(lower N 0 , higher Gt). The undeformedVariscan granodiorite (sample T1, Fig. 8c) exhibitslowest values <strong>of</strong> N 0 <strong>and</strong> highest values <strong>of</strong> Gt.Grain shapes, shape preferred orientation <strong>and</strong> grainboundary preferred orientationPlagioclase is a low-symmetry mineral characterizedby weakly elongated to round shapes under mostmetamorphic conditions. Degree <strong>of</strong> elongation <strong>of</strong>plagioclase grains is therefore an indication <strong>of</strong> intracrystallinedeformation. In contrast, amphibole ishighly elongate in low-metamorphic grade rocks, whileat high-grade ones it becomes spherical (Brodie &Rutter, 1987). Therefore, shape analysis combinedwith SPO provides information about the degree <strong>of</strong>deformation <strong>of</strong> both plagioclase <strong>and</strong> amphiboleaggregates.Both grain shapes <strong>and</strong> SPO were evaluated using thePolyLX toolbox (Lexa, 2003). The grain shapes arecharacterized by preferred orientation <strong>of</strong> the long axes<strong>and</strong> by mean axial ratio R <strong>of</strong> the best-fit ellipse on theindividual grains using the area-moments ellipse fittingmethod. To evaluate an overall single phase SPO,individual linear segments <strong>of</strong> grain boundaries <strong>of</strong>individual phases were treated as independent vectorscontributing to the bulk Scheidegger–Watson orientation-tensor.The eigenvalue analysis determines twoeigenvalues <strong>and</strong> assigns them to mutually perpendiculardirections. The ratio <strong>of</strong> eigenvalues R e is consideredas a rough estimate <strong>of</strong> the degree <strong>of</strong> SPO forthe given phase. Grain boundary preferred orientationwas assessed by a similar technique, with the bulkÓ 2005 Blackwell Publishing Ltd242


CONTRASTING TEXTURAL RECORD OF TWO METAMORPHIC EVENTS 661(a) (b)(c) (d)(d)Fig. 8. Representative digitized microstructures used in textural <strong>analyses</strong>. See text for detailed sample description.Scheidegger–Watson orientation-tensor formed fromthe individual linear segments <strong>of</strong> boundary tracesbetween the chosen phases. Resulting eigenvalues ratioR b is considered as a rough estimate <strong>of</strong> the degree <strong>of</strong>GBPO.Results <strong>of</strong> the grain shape analysis <strong>and</strong> SPO analysisare shown in Fig. 10. Both amphibole (Fig. 10b) <strong>and</strong>plagioclase (Fig. 10a) show systematic changes. Whilesamples with preserved magmatic textures (T1) <strong>and</strong>LAC samples (LAC1, TG2) exhibit very weak SPO<strong>and</strong> lowest values <strong>of</strong> axial ratios (c. 1.5), samples <strong>of</strong>mylonitic metagabbros show a significant increase inthe average axial ratio (2–3) <strong>and</strong> increase <strong>of</strong> the SPOdegree with decreasing temperature.Grain contact frequenciesGrain contact frequencies allow the statistical deviationfrom r<strong>and</strong>om spatial distribution <strong>of</strong> grainboundaries to be examined (Kretz, 1994). So far, thedegree <strong>of</strong> deviation <strong>of</strong> grain boundaries distributionfrom r<strong>and</strong>om distribution have been evaluated byplotting observed/expected ratio <strong>of</strong> like–like contacts<strong>of</strong> the two major minerals against each other. Here wepropose a diagram where v valuev ¼ Observed pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiExpected ð2ÞExpectedis plotted against the ratio <strong>of</strong> orientation-tensoreigenvalues, i.e. the degree <strong>of</strong> GBPO. The majoradvantage <strong>of</strong> this diagram is in simple visual evaluation<strong>of</strong> the degree <strong>of</strong> deviation from an expected r<strong>and</strong>omdistribution <strong>of</strong> grain contacts (v ¼ 0).Results <strong>of</strong> this analysis are presented in Fig. 10. Theamphibolite facies mylonitic metagabbros exhibitalmost a r<strong>and</strong>om distribution at higher strains (GLT2)<strong>and</strong> tend to develop more <strong>of</strong> an aggregate-type <strong>of</strong>microstructure (v values are positive for like–like <strong>and</strong>negative for unlike contacts in Fig. 10c,d) at lowerstrains (GLT1) accompanied by a significant decrease<strong>of</strong> preferred orientation <strong>of</strong> amphibole–amphibole <strong>and</strong>amphibole–plagioclase contacts. The high-grademylonitic metagabbros (GHT1 & GHT2) exhibit astrong aggregate distribution <strong>and</strong> a moderate preferredorientation <strong>of</strong> grain boundaries. Samples from theb<strong>and</strong>ed amphibolite complex (TG2 & LAC1) show atendency towards a regular distribution (negative like–like v values <strong>and</strong> positive unlike v values) <strong>and</strong> a lowdegree <strong>of</strong> GBPO. The tonalitic gneiss weakly overprintedby the Variscan deformation (LAC3) showsÓ 2005 Blackwell Publishing Ltd243


662 O. LEXA ET AL.Fig. 9. (a) Plot <strong>of</strong> crystal size distribution (CSD) parameters N 0 <strong>and</strong> Gt. Parameters are obtained from CSD plots, where straightline is fitted on plotted data. Gt represents the slope <strong>and</strong> N 0 the intercept <strong>of</strong> fitted line. Shaded ellipses from upper left to lowerright corner mark domains <strong>of</strong> LT metagabbros, HT metagabbros, b<strong>and</strong>ed amphibolites <strong>and</strong> magmatic rocks. See discussion.(b) Examples <strong>of</strong> CSD plots used for N 0 <strong>and</strong> Gt parameter estimates.transitional values <strong>of</strong> v <strong>and</strong> R b between those <strong>of</strong> theb<strong>and</strong>ed amphibolite complex <strong>and</strong> the mylonitic metagabbrosamples. The Variscan undeformed granodioritesill shows nearly a r<strong>and</strong>om distribution <strong>of</strong> graincontacts <strong>and</strong> an absence <strong>of</strong> GBPO.DISCUSSIONThe study samples <strong>of</strong> mafic lithologies from theCambro-Ordovician <strong>and</strong> Variscan structures exhibitremarkably similar mineral assemblages. In addition,the presence <strong>of</strong> garnet in metamorphosed mafic rocksindicates the medium pressure conditions for bothCambro-Ordovician <strong>and</strong> Variscan structures. Localoccurrence <strong>of</strong> orthopyroxene bearing mineral assemblagesin the b<strong>and</strong>ed amphibolites as well as in thewestern metagabbro mylonite belt (Table 1) indicatesthat both Cambro-Ordovician <strong>and</strong> Variscan metamorphicevents reached a boundary between granulite<strong>and</strong> amphibolite facies conditions.The metamorphic assemblages <strong>and</strong> P–T estimates <strong>of</strong>the Cambro-Ordovician metamorphic conditions <strong>of</strong>b<strong>and</strong>ed amphibolite complex calculated overlap withinthe likely uncertainties with those <strong>of</strong> the Variscanmylonitic metagabbros <strong>and</strong> granodiorite. Moreover, inthe southern part <strong>of</strong> the region studied, the Variscan<strong>and</strong> Cambro-Ordovician fabrics are concordant <strong>and</strong>therefore slight differences in metamorphic conditionsare indistinguishable. We suggest that in this case thest<strong>and</strong>ard methods <strong>of</strong> metamorphic petrology cannotdiscriminate the two distinct metamorphic events fromeach other.Interpretation <strong>of</strong> crystal size distributionsThe grain-size distribution in metamorphic rocks(Cashman & Ferry, 1988; Eberl et al., 1998) is primarilycontrolled by the interaction between nucleationrate which is sensitive to temperature overstepping(Cahn, 1957), <strong>and</strong> the growth rate which is anapproximately linear function <strong>of</strong> overstepping (Ridley& Thompson, 1986). A further process that influencesCSD is textural coarsening driven by a tendency todecrease the excess <strong>of</strong> interfacial energy to reach alower energy state (Voorhees, 1992). The transfer <strong>of</strong>material from smaller grains to larger ones occurs bydiffusion <strong>and</strong> is commonly expressed by the Lifshitz–Slyozov–Wagner (LSW) equation (Lifshitz & Slyozov,1961) for Ostwald ripening, or by the communicatingneighbours theory (CN) by DeH<strong>of</strong>f (1991) where adiffusion length scale is spatially dependent. This differencebetween the theories is expressed in the waythat the CSD is modified during coarsening (Higgins,1998). In the case <strong>of</strong> LSW the descriptive parametersN 0 <strong>and</strong> Gt remain constant, while the CN theoryimplies a decrease in N 0 <strong>and</strong> Gt during coarsening.Experimental studies have shown that the size <strong>of</strong>dynamically recrystallized grains is strongly dependenton recrystallization mechanisms that are controlled bystress <strong>and</strong> temperature (Twiss, 1977). Low-temperatureÓ 2005 Blackwell Publishing Ltd244


CONTRASTING TEXTURAL RECORD OF TWO METAMORPHIC EVENTS 663Fig. 10. (a, b) Plot <strong>of</strong> grains shape preferred orientation (SPO) <strong>of</strong> amphibole <strong>and</strong> plagioclase in studied samples. The results aresummarized by a boxplot-type plot <strong>of</strong> axial ratios v. eigenvalue ratios for individual phases. Individual boxes showing median, first <strong>and</strong>third quartile values. The whiskers represent statistical estimate <strong>of</strong> range <strong>of</strong> data while outliers are not shown. Vertical axis characterizesa shape <strong>of</strong> grains, while horizontal axis represents area-weighted degree <strong>of</strong> preferred orientations. (c, d) Grain boundaryfrequencies. Plots <strong>of</strong> deviations from r<strong>and</strong>om spatial distribution v. degree <strong>of</strong> preferred orientation. The value <strong>of</strong> deviation fromr<strong>and</strong>om spatial distribution is obtained by contact-frequency method (Kretz, 1969) <strong>and</strong> length-weighted degree <strong>of</strong> preferred orientationis estimated as ratio <strong>of</strong> eigenvalues <strong>of</strong> bulk matrix <strong>of</strong> inertia. Plot <strong>of</strong> like–like <strong>and</strong> unlike boundaries is separated into two panels.recrystallization mechanisms such as bulging <strong>and</strong> subgrainrotation recrystallization (Poirier & Guillope,1979) are processes dominated by heterogeneousnucleation <strong>and</strong> grain-size reduction. On the contrary,grain-boundary migration recrystallization is <strong>of</strong>tenconnected with grain coarsening (Urai et al., 1986).Hickey & Bell (1996) suggested that the syntectonicgrowth <strong>of</strong> minerals is primarily controlled by gradientsin lattice strain energy, which results from inhomogeneousdeformation involving dislocation creep.Decreasing the strain rate to temperature ratio ( _e/T)leads to a decrease in the ratio <strong>of</strong> nucleation to growthrate (N/G), <strong>and</strong> coarser grain size can develops. On thecontrary, increase <strong>of</strong> _e/T ratio leads to increasing N/G,<strong>and</strong> therefore the grain size decreases. This hypothesishas been supported by Azpiroz & Ferna´ndez (2003)who used CSD plots to analyse dynamicallyrecrystallized metabasites across a metamorphic gradient.The CSD in the mylonitic metagabbros studied arecharacterized by a decrease in N 0 <strong>and</strong> an increase inGt values with increasing metamorphic temperatureestimates, <strong>and</strong> are projected in areas <strong>of</strong> high N 0values <strong>and</strong> low Gt in the N 0 –Gt plot. If all theseresults are interpreted in terms <strong>of</strong> CSD theory, theabove-described evolutionary trend indicates adecreasing dominance <strong>of</strong> nucleation processes overgrain growth with increasing temperature. The correlation<strong>of</strong> metamorphic temperature <strong>and</strong> grain-sizedistribution could be explained as a result <strong>of</strong> texturalcoarsening using the CN theory (Higgins, 1998).However, the micro<strong>structural</strong> observations <strong>of</strong> theÓ 2005 Blackwell Publishing Ltd245


664 O. LEXA ET AL.onset <strong>of</strong> dynamic recrystallization in weaklydeformed metagabbros show that the lower-temperaturemylonites exhibit a smaller initial recrystallizationgrain size than the new recrystallized grains<strong>of</strong> the high-temperature mylonites (Baratoux et al.,2005). In addition, inspection <strong>of</strong> the N 0 –Gt plotshows that highly deformed samples (GLT2,GHT2a,b, T2) exhibit higher N 0 <strong>and</strong> lower Gt valuesrelative to weakly deformed samples (GLT1, GHT1).Based on these two observations, we suggest that thedifferences in CSD parameters are not the result <strong>of</strong>textural coarsening but are merely controlled bytemperature- <strong>and</strong> strain rate-dependent mechanisms<strong>of</strong> dynamic recrystallization. Azpiroz & Ferna´ndez(2003) reported an increase in N 0 but a constant Gt(slope) with decreasing temperature <strong>and</strong> increasing Gtvalues for constant N 0 with increasing finite strain inthe Iberian Massif metabasites. However, it is proposedthat in our study the N 0 <strong>and</strong> Gt changesimultaneously due to temperature <strong>and</strong> strainintensity variations <strong>and</strong> that the temperature changeplays a key role in the resulting CSD shape.Samples from the b<strong>and</strong>ed amphibolite complex aremarked by lower N 0 <strong>and</strong> higher Gt values in comparisonwith the mylonitic metagabbros. The CSD isdeveloped in both amphibolites <strong>and</strong> tonalitic gneiss,which indicates that it is independent on a relativeproportion <strong>of</strong> amphibole <strong>and</strong> plagioclase in both rocktypes. These data in the N 0 –Gt diagram show a continuoustrend together with the above-described samplesfrom the mylonitic metagabbros. In addition,results <strong>of</strong> Hb–Pl thermometry reveal an increase inestimated temperatures from the eastern mylonitemetagabbros, through the western mylonite metagabbrosto the b<strong>and</strong>ed amphibolite complex. Such anevolutionary trend is likely to be interpreted as theresult <strong>of</strong> a textural coarsening comparable with theresults <strong>of</strong> Cashman & Ferry (1988) reinterpreted byHiggins (1998).Interpretation <strong>of</strong> spatial distribution minerals <strong>and</strong> grainboundariesSeng (1936) <strong>and</strong> later DeVore (1959) proposed that thespatial distribution <strong>of</strong> crystals in high-grade gneisses isdominantly determined by interfacial energy. Thisconcept was adopted by Flinn (1969) who explained aprevailing number <strong>of</strong> unlike boundaries in granulitesthrough the insertion <strong>of</strong> grains <strong>of</strong> one phase betweengrains <strong>of</strong> other phases. Flinn (1969) suggested that thisfeature is a consequence <strong>of</strong> a smaller interfacial energy<strong>of</strong> unlike boundaries in comparison with like–likeboundaries. However, Ramberg (1952) suggested thatdifferences in interfacial energies are too small to drivediffusional mass transfer in high-grade rocks.Modern material science experimental studies showthat during the wetting process the low-energy (lowmisorientationangle) boundaries in one phase arepreserved while another phase preferentially precipitateson high-energy (high-misorientation angle)boundaries (e.g. Kim & Rohrer, 2004). In other words,the highest energy boundaries are progressively eliminatedfrom an inherited population by ÔinfiltrationÕ <strong>of</strong>the other phase. This is in agreement with the knownfact that in granular-polygonal aggregates the minorphase precipitates on triple points to achieve lowertotal interfacial energy (Spry, 1969; Vernon, 1974).Such a tendency was documented by Dallain et al.(1999) who showed that the predominance <strong>of</strong> unlikecontacts in polycrystalline aggregates originatedthrough wetting <strong>of</strong> grain boundaries by fluids or melts,<strong>and</strong> subsequent precipitation <strong>of</strong> other phases on like–like contacts.In contrast, the solid-state differentiation resultingfrom dynamic recrystallization leads to the development<strong>of</strong> monomineralic aggregates or b<strong>and</strong>s due tocoalescence <strong>of</strong> like phases at high strains (Schulmannet al., 1996; Kruse & Stu¨ nitz, 1999). Therefore, thelike–like contacts prevail <strong>and</strong> the so-called aggregatetypedistribution develops, which is a typical feature <strong>of</strong>high-temperature deformation <strong>of</strong> polyphase rocks suchas gabbros <strong>and</strong> granites (Dallain et al., 1999; Baratouxet al., 2005).The high-temperature mylonitic metagabbros fromthe study area exhibit high-grain SPO <strong>and</strong> GPBOassociated with the development <strong>of</strong> a strong aggregatedistribution. The lower-temperature mylonitic metagabbrosare characterized by extreme values <strong>of</strong> SPO inconjunction with an almost r<strong>and</strong>om grain distribution.We suggest that in the case <strong>of</strong> high-temperaturemylonitic metagabbros the process controlling thedevelopment <strong>of</strong> a strong aggregate distribution is solidstatedifferentiation due to a different efficiency <strong>of</strong>dislocation creep in hornblende <strong>and</strong> plagioclase. Thisprocess is likely to be accompanied by some diffusionalmass transfer responsible for preferential heterogeneousnucleation (Kruse & Stu¨ nitz, 1999) <strong>of</strong> interstitialplagioclase in coarse-grained amphibole aggregates(Baratoux et al., 2005). On the contrary, in the lowertemperaturemetagabbro mylonites, the r<strong>and</strong>om mineraldistribution <strong>and</strong> the lack <strong>of</strong> crystallographicpreferred orientation <strong>of</strong> plagioclase were interpreted tobe the result <strong>of</strong> mechanical mixing due to grainboundarysliding during granular flow.Samples from the b<strong>and</strong>ed amphibolite <strong>and</strong> tonaliticgneiss show a low SPO, a very weak elongation <strong>of</strong> bothplagioclase <strong>and</strong> amphibole, a weak GBPO connectedwith a weak dominance <strong>of</strong> unlike contacts indicating aregular to anticlustered grain distribution. Such a graindistribution <strong>and</strong> the large grain size <strong>of</strong> both phasesexclude mechanical mixing as a process explaining thistexture. We are <strong>of</strong> the opinion that the spatial distribution<strong>of</strong> plagioclase <strong>and</strong> amphibole (e.g. sampleLAC1) can result from heterogeneous nucleation <strong>of</strong>plagioclase in an amphibole aggregates. However, anoriginal grain-size distribution characteristic <strong>of</strong> nucleation<strong>and</strong> growth processes is completely obliteratedby elimination <strong>of</strong> the small grains. Very low-grainÓ 2005 Blackwell Publishing Ltd246


CONTRASTING TEXTURAL RECORD OF TWO METAMORPHIC EVENTS 665elongation, SPO <strong>and</strong> GBPO, similar CSD <strong>of</strong> bothphases <strong>and</strong> regular grain distribution indicate that theaggregates tend to achieve a state with a minimuminterfacial energy. As mentioned above, the processes<strong>of</strong> heterogeneous nucleation <strong>of</strong> minor phase lead tooccupation <strong>of</strong> the high-energy grain boundaries atearly stages <strong>of</strong> the texture development. However, inthe rocks studied it is impossible to distinguish a minorphase from the host aggregate, which indicates that theprocess <strong>of</strong> interfacial energy reduction is more advanceddue to significant diffusional mass transfer. Asthe driving differences in bulk interfacial energies aretoo small (Ramberg, 1952), the only plausible explanationis the long time-scale <strong>of</strong> the process.Different time-scales <strong>of</strong> Cambro-Ordovician <strong>and</strong> Variscanmetamorphic eventsSˇtípska´ et al. (2001) proposed a tectonic model inwhich the Cambro-Ordovician metamorphism is relatedto large-scale rifting while the Variscan metamorphicevent was connected with a thermal effect inducedby the syntectonic granodiorite sill intrusion, <strong>and</strong> wasspatially restricted to the host rocks <strong>of</strong> the intrusion.The thermal time constant s is given by the relationship:s / l 2 j ;ð3Þwhere l is the characteristic length <strong>of</strong> thermal event <strong>and</strong>j is the thermal diffusivity (Carlslaw & Jaeger, 1959).This equation indicates that duration <strong>of</strong> thermalequilibration increases with the square <strong>of</strong> the size <strong>of</strong> thethermal anomaly. Based on the proposed tectonicmodel, we assume that a relaxation <strong>of</strong> the perturbedtemperature field generated by the continental riftdiffers in time-scale by at least one order <strong>of</strong> magnitudefrom that generated by an intrusion <strong>of</strong> several km insize. In other words, the metamorphic P–T conditionsattained at a similar depth level may yield similartemperatures but the time required for metamorphicequilibration <strong>and</strong> development <strong>of</strong> specific metamorphictextures may differ substantially.ACKNOWLEDGEMENTSThis research is a part <strong>of</strong> an ongoing collaborationbetween Charles University, Prague, Mainz University,Germany, <strong>and</strong> ETH Zu¨ rich, Switzerl<strong>and</strong>. Financialsupport to P. Sˇtı´pska´ by the Charles University GrantAgency (grant no. 223/2002/B-GEO/PrF) <strong>and</strong> theCzech National Grant Agency (GA205/00/D043 <strong>and</strong>GA205/99/1195) is gratefully acknowledged. Themicroprobe work at ETH Zu¨ rich was financed by theSwiss National Foundation (ÔContinuous OrogenesisÕgrant to A.B. Thompson). Two visits <strong>of</strong> P. Sˇtı´pska´ toMainz University <strong>and</strong> a part <strong>of</strong> the microprobe workwere funded by the German Science Foundation(DFG, grant Kr 68-1, Kr 590/35). The grant no.24313005 <strong>of</strong> the Ministry <strong>of</strong> Education <strong>of</strong> the CzechRepublic provided funds for O. Lexa, P. Sˇtı´pska´ <strong>and</strong>K. Schulmann.REFERENCESAzpiroz, M. D. & Fernández, C., 2003. Characterization <strong>of</strong>tectono-metamorphic events using crystal size distribution(CSD) diagrams. A case study from the Acebuches metabasites(SW Spain). Journal <strong>of</strong> Structural Geology, 25, 935–947.Baratoux, L., Schulmann, K., Ulrich, S. & Lexa, O., 2005.Contrasting microstructures <strong>and</strong> deformation mechanisms inmetagabbro mylonites contemporaneously deformed underdifferent temperatures (c. 650 <strong>and</strong> c. 750 °C). In: DeformationMechanisms, Rheology <strong>and</strong> Tectonics: from Minerals to theLithosphere (eds Gapais, D., Brun, J. P. & Cobbold, P. R.),Geological Society Special Publications, pp. 97–125. GeologicalSociety <strong>of</strong> London, London, UK.Bohlen, S. R., Wall, V. J. & Boettcher, A. L., 1983. Experimentalinvestigations <strong>and</strong> geological applications <strong>of</strong> equilibria in thesystem FeO–TiO 2 –Al 2 O 3 –SiO 2 –H 2 O. American Mineralogist,68, 1049–1058.Brodie, K. H. & Rutter, E. H., 1987. The role <strong>of</strong> transiently finegrainedreaction-products in syntectonic metamorphism –natural <strong>and</strong> experimental examples. Canadian Journal <strong>of</strong> EarthSciences, 24, 556–564.Burg, J.-P., 1999. Ductile structures <strong>and</strong> instabilities: theirimplication for Variscan tectonics in the Ardennes. Tectonophysics,309, 1–25.Cahn, J. V., 1957. Nucleation on dislocations. Acta Metallurgica,5, 169–172.Carlslaw, H. S. & Jaeger, J. C., 1959. Conduction <strong>of</strong> Heat inSolids. Oxford University Press, Oxford.Cashman, K. V. & Ferry, J. M., 1988. Crystal size distribution(CSD) in rocks <strong>and</strong> the kinetics <strong>and</strong> dynamics <strong>of</strong> crystallization;3, Metamorphic crystallization. Contributions toMineralogy <strong>and</strong> Petrology, 99, 401–415.Dallain, C., Schulmann, K. & Ledru, P., 1999. Textural evolutionin the transition from subsolidus annealing to meltingprocess, Velay Dome, French Massif Central. Journal <strong>of</strong>Metamorphic Geology, 17, 61–74.DeH<strong>of</strong>f, R. T., 1991. A geometrically general-theory <strong>of</strong> diffusioncontrolled coarsening. Acta Metallurgica Et Materialia, 39,2349–2360.DeVore, G. W., 1959. Role <strong>of</strong> minimum interfacial free energy indetermining the macroscopic features <strong>of</strong> minerals assemblages.Journal <strong>of</strong> Geology, 67, 211–227.Eberl, D. D., Drits, V. A. & Srodon J., 1998. Deducing growthmechanisms for minerals from the shapes <strong>of</strong> crystal size distributions.American Journal <strong>of</strong> Science, 298, 499–533.Ferry, J. M. & Spear, F. S., 1978. Experimental calibration <strong>of</strong> thepartitioning <strong>of</strong> Fe <strong>and</strong> Mg between biotite <strong>and</strong> garnet.Contributions to Mineralogy <strong>and</strong> Petrology, 66, 113–117.Flinn, D., 1969. Grain contacts in crystalline rocks. Lithos, 2,361–370.Gasparik, T. & Newton, R. C., 1984. The reversed aluminacontents <strong>of</strong> orthopyroxene in equilibrium with spinel <strong>and</strong>forsterite in the system MgO–Al 2 O 3 –SiO 2 . Contributions toMineralogy <strong>and</strong> Petrology, 85, 186–196.Hickey, K. A. & Bell, T. H., 1996. Syn-deformational graingrowth: matrix coarsening during foliation development <strong>and</strong>regional metamorphism rather than by static annealing.European Journal <strong>of</strong> Mineralogy, 8, 1351–1373.Higgins, M. D., 1998. Origin <strong>of</strong> anorthosite by textural coarsening:quantitative measurements <strong>of</strong> a natural sequence <strong>of</strong>textural development. Journal <strong>of</strong> Petrology, 39, 1307–1323.Holl<strong>and</strong>, T. J. B. & Blundy, J. D., 1994. Non-ideal interactionsin calcic amphiboles <strong>and</strong> their bearing on amphibole–plagio-Ó 2005 Blackwell Publishing Ltd247


666 O. LEXA ET AL.clase thermometry. Contributions to Mineralogy <strong>and</strong> Petrology,116, 433–447.Kim, C.-S. & Rohrer, G. S., 2004. Geometric <strong>and</strong> crystallographiccharacterization <strong>of</strong> WC surfaces <strong>and</strong> grainboundaries in WC-Co composites. Interface Science, 12, 19–27.Kohn, M. J. & Spear, F. S., 1990. Two new barometers forgarnet amphibolites with applications to southeasternVermont. American Mineralogist, 75, 89–96.Koziol, A. M., 1989. Recalibration <strong>of</strong> the garnet–plagioclase–Al 2 SiO 5 –quartz (GASP) geobarometer <strong>and</strong> application tonatural parageneses. Transactions <strong>of</strong> the American GeophysicalUnion, 70, 493.Kretz, R., 1969. On the spatial distribution <strong>of</strong> crystals in rocks.Lithos, 2, 39–66.Kretz, R., 1983. Symbols for rock forming minerals. AmericanMineralogist, 68, 277–279.Kretz, R., 1994. Metamorphic Crystallization. John Wiley &Sons, Chichester.Kro¨ ner, A., O’Brien, P. J., Nemchin, A. A. & Pidgeon, R. T.,2000a. Zircon ages for high pressure granulites from SouthBohemia, Czech Republic, <strong>and</strong> their connection to Carboniferoushigh temperature processes. Contributions to Mineralogy<strong>and</strong> Petrology, 138, 127–142.Kro¨ ner, A., Sˇtı´pska´, P., Schulmann, K. & Jaeckel, P., 2000b.Chronological constraints on the pre-Variscan evolution <strong>of</strong>the northeastern margin <strong>of</strong> the Bohemian Massif, CzechRepublic. In: Orogenic Processes: Quantification <strong>and</strong> Modellingin the Variscan Belt (eds Franke, W., Haak, W., Oncken,O. & Tanner, D.), Geological Society, London, Special Publications,179, 175–197.Kruse, R. & Stu¨ nitz, H., 1999. Deformation mechanisms <strong>and</strong>phase distribution in mafic high- temperature mylonites fromthe Jotun Nappe, southern Norway. Tectonophysics, 303, 223–249.Leake, B. E., Woolley, A. R., Arps, C. E. et al., 1997. Nomenclature<strong>of</strong> amphiboles: Report <strong>of</strong> the Subcommittee onAmphiboles <strong>of</strong> the International Mineralogical AssociationCommission on new minerals <strong>and</strong> mineral names. MineralogicalMagazine, 61, 295–321.Lexa, O., 2003. Numerical Approaches in Structural <strong>and</strong> Micro<strong>structural</strong>Analyses. Unpublished PhD Thesis, Charles University,Prague.Lifshitz, I. M. & Slyozov, V. V., 1961. The kinetics <strong>of</strong> precipitationfrom supersaturated solid solutions. Journal <strong>of</strong>Physics <strong>and</strong> Chemistry <strong>of</strong> Solids, 19, 35–50.Marsh, B. D., 1988. Crystal size distribution (CSD) in rocks <strong>and</strong>the kinetics <strong>and</strong> dynamics <strong>of</strong> crystallization; 1, Theory.Contributions to Mineralogy <strong>and</strong> Petrology, 99, 277–291.O’Neill, H. S. C. & Wall, V. J., 1987. The olivine–orthopyroxene–spineloxygen geobarometer, the Nickel precipitationcurve, <strong>and</strong> the oxygen fugacity <strong>of</strong> the Earth’s upper mantle.Journal <strong>of</strong> Petrology, 28, 1169–1191.Parry, M., Sˇtı´pska´, P., Schulmann, K., Hrouda, F., Jezˇek, J. &Kro¨ ner, A., 1997. Tonalite sill emplacement at an obliqueplate boundary: northeastern margin <strong>of</strong> the Bohemian Massif.Tectonophysics, 280, 61–81.Peterson, T. D., 1996. A refined technique for measuring crystalsize distributions in thin section. Contributions to Mineralogy<strong>and</strong> Petrology, 124, 395–405.Poirier, J. P. & Guillope, M., 1979. Deformation inducedrecrystallisation <strong>of</strong> minerals. Bulletin de Mineralogie, 102,67–74.Powell, R. & Holl<strong>and</strong>, T. J. B., 1988. An internally consistentdataset with uncertainties <strong>and</strong> correlations: 3. Applications togeobarometry, worked examples <strong>and</strong> a computer program.Journal <strong>of</strong> Metamorphic Geology, 6, 173–204.Ramberg, H., 1952. The Origin <strong>of</strong> Metamorphic <strong>and</strong> MetasomaticRocks. University <strong>of</strong> Chicago Press, Chicago, IL.R<strong>and</strong>olph, A. D. & Larson, M. A., 1971. Theory <strong>of</strong> ParticleProcesses. Academic Press, New York.Ridley, J. & Thompson, A. B., 1986. The role <strong>of</strong> mineral kineticsin the development <strong>of</strong> metamorphic microtextures. In: Fluid–Rock Interactions during Metamorphism (eds Walther, J. V. &Wood, B. J.), pp. 154–193. Springer-Verlag, New York.Schmidt, M. W., 1992. Amphibole composition in tonalite as afunction <strong>of</strong> pressure: an experimental calibration <strong>of</strong> the Al-inhornblendebarometer. Contributions to Mineralogy <strong>and</strong>Petrology, 110, 304–310.Schulmann, K., Mlcoch, B. & Melka, R., 1996. High-temperaturemicrostructures <strong>and</strong> rheology <strong>of</strong> deformed granite, Erzgebirge,Bohemian Massif. Journal <strong>of</strong> Structural Geology, 18,719–733.Seng, H., 1936. Die Migmatitfrage und der Mechanismusparakristalliner Pra¨ gung. Geologische Rundschau, 27, 471–492.Spear, F. S., Peacock, S. M., Kohn, M. J., Florence, F. P. &Menard, T., 1991. Computer programs for petrologic P–T–tpath calculations. American Mineralogist, 76, 2009–2012.Spry, A. H., 1969. Metamorphic Textures. Pergamon Press,Oxford.Sˇtı´pska´, P., Schulmann, K., Thompson, A. B., Jezˇek, J. &Kro¨ ner, A., 2001. Thermo-mechanical role <strong>of</strong> a Cambro-Ordovician paleorift during the Variscan collision: the NEmargin <strong>of</strong> the Bohemian Massif. Tectonophysics, 332, 239–253.Thompson, A. B., 1976. Mineral reactions in pelitic rocks: II.Calculation <strong>of</strong> some P–T–X (Fe–Mg) phase relations. AmericanJournal <strong>of</strong> Science, 276, 425–454.Turner, J. T. & Weiss, L. E., 1963. Structural Analysis <strong>of</strong>Metamorphic Tectonictes. McGraw-Hill, San Francisco, CA.Twiss, R. J., 1977. Theory <strong>and</strong> applicability <strong>of</strong> a recrystallizedgrain size paleopiezometer. Pure <strong>and</strong> Applied Geophysics, 115,227–244.Urai, J. L., Means, W. D. & Lister, G. S., 1986. Dynamicrecrystallization <strong>of</strong> minerals. AGU Geophysical Monograph,36, 161–199.Vernon, R. H., 1974. Controls <strong>of</strong> mylonitic compositionallayering during non-cataclastic ductile deformation. GeologicalMagazine, 111, 121–123.Voorhees, P. V., 1992. Ostwald ripening <strong>of</strong> two-phase mixtures.Annual Review <strong>of</strong> Materials Science, 22, 197–215.Received 3 June 2005; revision accepted 17 July 2005.Ó 2005 Blackwell Publishing Ltd248


Contrasting microstructures <strong>and</strong> deformation mechanismsin metagabbro mylonites contemporaneously deformed underdifferent temperatures (c. 650 ~ <strong>and</strong> c. 750 ~L. BARATOUX 1'2'5, K. SCHULMANN 3, S. ULRICH t'4 & O. LEXA t~Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University,Albertov 6, 12843, Prague, Czech Republic(e-mail: Ika @natur.cuni.cz)2UMR 5570, ENS <strong>and</strong> Lyon 1 University, 2 rue RaphaYl Dubois,69622, Villeurbanne Cedex, France3Universitd Louis Pasteur, EOST, UMR 7517, 1 Rue Blessig,Strasbourg, 67084, France4Geophysical Institute, Czech Academy <strong>of</strong> Sciences, Bo(n{ 11/1401,14131 Praha 4, Czech Republic5Present address." Czech Geological Survey, Klarov 3, Praha 1,11821, Czech RepublicAbstract: Deformation mechanisms <strong>of</strong> amphibole <strong>and</strong> plagioclase were investigated in twometagabbroic sheets (the eastern <strong>and</strong> western metagabbros from the Stars M~sto belt,eastern Bohemian Massif), using petrology, quantitative micro<strong>structural</strong> <strong>and</strong> electron backscattereddiffraction methods. After the gabbroic pyroxene was replaced by amphibole, bothgabbroic bodies became progressively deformed. The eastern metagabbros were deformedunder temperature <strong>of</strong> c. 650 ~ <strong>and</strong> the western metagabbros under c. 750 ~ Subgrainrotation <strong>and</strong> dislocation creep, characterized by strong crystallographic <strong>and</strong> shape preferredorientations, operated in plagioclase <strong>of</strong> the eastern belt during the early stages <strong>of</strong> deformation.Subsequent r<strong>and</strong>omizing <strong>of</strong> plagioclase crystallographic preferred orientation isinterpreted to be due to grain boundary sliding in the mylonitic stage. Large (50-150 ixm) grain sizes during the mylonitic stages are interpreted to be due to low strainrates. Amphibole is stronger <strong>and</strong> deforms cataclastically, leading to important grain sizereduction when the bulk rock strength drops substantially. In the western belt, plagioclasedeformed by dislocation creep accompanied by grain boundary migration (possibly chemicallyinduced) while heterogeneous nucleation <strong>and</strong> syndeformational grain growth inconjunction with dislocation creep were typical for amphiboles.Micro<strong>structural</strong> <strong>and</strong> rheological behaviour <strong>of</strong>natural polyphase rocks is a complex problem,which has been studied in detail mostly inquartzo-feldpathic rocks (e.g. Gapais 1989;H<strong>and</strong>y 1990). In these rocks, feldspars are generallyconsidered as strong minerals while quartzrepresents the weak phase. H<strong>and</strong>y (1994) proposeda scheme for rocks containing minerals <strong>of</strong> contrastingrheology with two end-members: load-beatingframework (LBF) <strong>and</strong> interconnected weak layers(IWL), based on the assumption that at least onemineral (generally the weaker one) is deformedby the mechanism <strong>of</strong> dislocation creep. Aspointed out by Jordan (1988) the LBF is notstable with increasing strain, resulting in mechanicallyinduced compositional layering. The strength<strong>of</strong> a two-phase material depends on the proportion<strong>of</strong> the weak mineral <strong>and</strong> on the rheological contrastbetween the two phases (Jordan 1988; H<strong>and</strong>y1990, 1994). However, if the deformation mechanismswitches from crystal plasticity to some grainsize sensitive process, the strength <strong>of</strong> the bulk rockdrops rapidly with respect to the preceding modeldue to strain s<strong>of</strong>tening (Knipe 1989). It was alsoshown that generally stronger feldspars maybecome even weaker than quartz if they deformby diffusional creep (e.g. Voll 1976; Simpson1985; Gapais 1989; Martelat et al. 1999).Equivalent comparative micro<strong>structural</strong> study<strong>of</strong> plagioclase <strong>and</strong> hornblende in naturally deformedmetabasic rocks as well as experimentalstudies <strong>of</strong> the rheology <strong>of</strong> amphibole-plagioclaseFrom: GAPA1S, D., BRUN, J. P. & COBBOLD, P. R. (eds) 2005. Deformation Mechanisms, Rheology <strong>and</strong>Tectonics: from Minerals to the Lithosphere. Geological Society, London, Special Publications, 243, 97-125.0305-8719/05/$15.00 ~') The Geological Society <strong>of</strong> London 2005.249


98 L. BARATOUX ETAL.bearing rocks are rare (e.g. Hacker & Christie1990; Wilks & Carter 1990). However, metabasitesare considered to constitute a major part<strong>of</strong> the lower continental crust (Rutter & Brodie1992) <strong>and</strong> the study <strong>of</strong> deformation microstructures<strong>and</strong> mechanisms is therefore crucial forunderst<strong>and</strong>ing the deformation behaviour <strong>and</strong>rheology <strong>of</strong> mafic tectonites. Plagioclase rheologyis believed to control the strength <strong>of</strong> the lowercrust in several models (e.g. Carter & Tsenn1987; Ord& Hobbs 1989). Hornblende isassumed to be a relatively strong phase <strong>and</strong> tobehave passively during deformation unless itforms a load-supporting framework (Brodie &Rutter 1985).Many studies concerned with deformationmechanisms <strong>of</strong> plagioclase <strong>and</strong> hornblendehave been published to date. Experiments suggestthat plagioclase is deformed by dislocation creepunder lower crustal conditions (e.g. Tullis &Yund 1987; Ji & Mainprice 1990; Kruse et al.2001) with commonly active (010)[001] principalslip system (e.g. Olsen & Kohlstedt 1985;Kruhl 1987). Kruhl (1987) also observed a(001) [ 100] slip system in naturally deformed plagioclases<strong>and</strong> Stiinitz et al. (2003) have shownthat slip on (001) <strong>and</strong> {111} in (110) directionare similarly active in experimentally deformedAn60 crystals. Besides these examples, otherless common slip systems were suggested byMarshall & McLaren (1977a, b), Montardi &Mainprice (1987), <strong>and</strong> Olsen & Kohlstedt(1984, 1985). In addition, grain size sensitiveprocesses in plagioclases have been referred toin some studies (Boullier & Gu6guen 1975;Lapworth et al. 2002). Hornblende generallyundergoes brittle deformation under low temperatureconditions (e.g. Brodie & Rutter 1985;Nyman et al. 1992; Lafrance & Vernon 1993).Many studies document crystal plastic deformation<strong>of</strong> hornblende with a dominant (100)[001] slip system from experiments <strong>and</strong> naturalrocks (e.g. Rooney et al. 1970; Cumbest et al.1989a, b; Hacker & Christie 1990). Volume diffusionrates are, however, slow in amphiboles(e.g. Freer 1981) implying that dislocationclimb is limited even at geological strain rates.Syndeformational chemical reactions betweenhornblende <strong>and</strong> plagioclase are common in metabasicrocks (Brodie 1981; Brodie & Rutter 1985)<strong>and</strong> involve deformation mechanisms such aschemically induced grain boundary migration(CIGM) (Cumbest et al. 1989a), on nucleation<strong>of</strong> new plagioclase (Rosenberg & Stfinitz 2003)or hornblende at plagioclase grain boundaries(Kruse & Sttinitz 1999).Metagabbros from the Star6 M6sto belt (easternmargin <strong>of</strong> the Bohemian Massif) representan example <strong>of</strong> a two-phase metabasic systemdeformed at different temperatures <strong>and</strong> straingradients. The aim <strong>of</strong> this study is to show twotypes <strong>of</strong> progressive evolution <strong>of</strong> deformationmicrostructures <strong>and</strong> textures <strong>of</strong> dynamicallyrecrystallized plagioclase-hornblende bearingmetagabbros at amphibolite <strong>and</strong> upper amphibolitefacies conditions with increasing bulk strain.Methods such as quantitative textural analysis,crystallographic preferred orientations (CPO)<strong>and</strong> study <strong>of</strong> mineral chemistry were used toconstrain deformation mechanisms for bothminerals. Finally, deformation mechanisms <strong>of</strong>rheologically contrasting plagioclase <strong>and</strong> hornblendeare correlated <strong>and</strong> the mechanical behaviour<strong>of</strong> mafic rocks deformed under lowercrustal metamorphic conditions is discussed.Geological settingThe Star6 M6sto (SM) belt in the eastern margin<strong>of</strong> the Bohemian Massif separates the highgrade gneisses <strong>of</strong> thickened continental crust <strong>of</strong>the Lugian domain in the west from a Neo-Proterozoic continental margin in the east(Fig. l a). The SM domain was thinned duringCambro-Ordovician rifting <strong>and</strong> underwent granulitefacies metamorphism (Stfpskfi et al. 2001).Subsequent Variscan tectonics resulted in NE-SW trending structures dipping at relativelyhigh angles to the west (Figs 1 a <strong>and</strong> b).Strongly deformed 'western' metagabbros <strong>of</strong>Cambro-Ordovician protolith ages occur at thetop <strong>of</strong> the SM belt (Fig. 1) (Kr6ner et aL 2000),The metagabbros were pervasively affected bya ductile shear zone along which was emplacedsyntectonically a Carboniferous tonalite silldated at 340 Ma (Stfpskfi et aL 2004). A carboniferousmetamorphism <strong>of</strong> adjacent metagabbrosreached higher amphibolite facies conditionsbecause <strong>of</strong> the strong heat input from the tonalitesill. A leptyno-amphibolite complex <strong>of</strong> Cambro-Ordovician age comprising a sequence <strong>of</strong> alternatingamphibolites <strong>and</strong> tonalitic gneissesoccurs in the footwall <strong>of</strong> the tonalite sill. Thiscomplex is underlined by the 'eastern' metagabbrosheet, which is supposed to form part <strong>of</strong> thesame lower crust as the western (upper) metagabbrosheet (Stfpskfi et aL 2001). Amphibolitefacies Carboniferous metamorphic conditions <strong>of</strong>the eastern (lower) metagabbro are also attributedto the heat input <strong>of</strong> a more distant hangingwalltonalite intrusion.Rocks <strong>of</strong> Cambro-Ordovician age suffered twodeformations: D~ <strong>of</strong> Cambro-Ordovician age <strong>and</strong>D2 <strong>of</strong> Carboniferous age. The latter one prevailsin most lithologies <strong>of</strong> the SM belt, marked by apenetrative west-dipping $2 foliation bearing a250


TEXTURES OF NATURALLY DEFORMED METAGABBROS 99Fig. 1. (a) Location <strong>of</strong> the studied area in the frame <strong>of</strong> the European Variscides. Geological map <strong>of</strong> the Star6 M~stobelt is based on unpublished geological maps at 1:25 000 provided by courtesy <strong>of</strong> the Czech Geological Survey(Dr. M. Opletal, author). Important thrusts <strong>and</strong> faults as well as location <strong>of</strong> samples <strong>and</strong> a cross-section A-A' (Fig. lb)are indicated. (b) Schematic cross-section based on field observations shows major structures, lithology <strong>and</strong> majortectonic boundaries. (Vertical axis not to scale.)subhorizontal or gently inclined N-S trendingmineral lineation. The eastern metagabbros inthe hangingwall <strong>of</strong> the high-grade rocks <strong>of</strong> theSilesian domain were affected by localized D2shear zones that developed under amphibolitefacies conditions attained during the peak <strong>of</strong>the Carboniferous metamorphism. Identical geometry<strong>of</strong> structures <strong>and</strong> kinematics suggest thatthe eastern <strong>and</strong> western metagabbro belts weredeformed under a dextral transpressive shear251


100 L. BARATOUX ETAL.regime but under different thermal conditions(Stfpsk~i et al. 2001).Methods <strong>and</strong> techniquesMineral chemistryMinerals were analysed with a Cameca SX 100electron microprobe equipped with four WDSspectrometers at Blaize Pascal University,Clermont-Ferr<strong>and</strong>, France. Operating conditionswere 15kV, 10hA beam current, 2-5p~mbeam diameter, 20 s counting time, <strong>and</strong> naturalmineral st<strong>and</strong>ards. Some hornblende <strong>and</strong> plagioclasewere analysed using a CamScan $4scanning electron microscope <strong>and</strong> attachedLink EDX microanalytical system, at CharlesUniversity, Prague, Czech Republic. Operatingconditions were 20 kV, 1.8 nA beam current,1-3 p,m beam diameter, 120 s counting time,<strong>and</strong> mineral st<strong>and</strong>ards Structure Probe Instruments(SPI). Ca maps from plagioclase weremade at Claude Bernard University in Lyon,with operating conditions 15 kV, 15 nA beamcurrent <strong>and</strong> spatial resolution <strong>of</strong> 512 x 512pixels.<strong>Quantitative</strong> micro<strong>structural</strong> analysis<strong>Quantitative</strong> micro<strong>structural</strong> analysis <strong>of</strong> grainboundaries was carried out on traced <strong>and</strong> digitizedoutlines <strong>of</strong> grains in ESRI ArcView 3.2Desktop GIS environment. The map <strong>of</strong> grainboundaries was generated using ArcView extensionPoly (Lexa 2003). The resulting polygonshave been treated by MATLAB TM PolyLXToolbox (http://petrol.natur.cuni.cz/~ ondro;Lexa 2003), in which grain shape <strong>and</strong> grainboundary preferred orientations (SPO <strong>and</strong>GBPO, respectively) were analysed using themoments <strong>of</strong> inertia ellipse fitting <strong>and</strong> eigenanalysis<strong>of</strong> bulk orientation tensor techniques (Lexa,2003; modified SURFOR technique by Panozzo(1983) for GBPO). Their degree is expressed asthe eigenvalue ratio (r = E1/E2) <strong>of</strong> the weightedorientation tensor <strong>of</strong> grain shapes or boundaries.The orientation is defined by V~ <strong>and</strong> V2 eigenvectors.The grain sizes <strong>of</strong> the minerals werecalculated in terms <strong>of</strong> Ferret diameters <strong>of</strong> grainsection without stereological corrections.Crystallographic preferred orientationAmphibole <strong>and</strong> plagioclase crystallographic preferredorientations (CPO) were collected using ascanning electron microscope CamScan $4 inPrague <strong>and</strong> a JEOL JSM 5600 in Montpellierequipped with Channel5 electron backscatterdiffraction (EBSD) system from HKL Technology(Prior et al. 1999). Thin sections werepolished using 0.25 ~m diamond paste. Toremove all surface damage <strong>and</strong> minimize reliefbetween minerals, sections were chemicallypolished using a colloidal silica suspension. Allthin sections were carbon-coated. The coatingreduces the quality <strong>of</strong> the electron backscatterdiffraction patterns (EBSP) so that automaticindexing mode <strong>of</strong> the EBSP system could notbe used. Most data were therefore collectedmanually. Operating conditions were 20 kV inPrague <strong>and</strong> 15 kV in Montpellier, 5.6 nA beamcurrent, working distance 39 ram, <strong>and</strong> 2-5 ~mbeam diameter. For each measurement, threeEuler angles (vl, 05, v2) characterizing thelattice orientation as well as the nature <strong>of</strong> themineral were determined <strong>and</strong> stored. Practiceshows that plagioclase diffraction patterns donot change significantly from albite up to atleast An65 (Lapworth et al. 2002). Therefore,the Anorthite48 database was used for indexingplagioclase. Pole figures <strong>and</strong> inverse polefigures were projected using the s<strong>of</strong>tware developedby D. Mainprice (ftp://ftp.dstu.univmontp2.fr/pub/TPHY/david/pc).Projections<strong>of</strong> crystallographic axes ([a], [b] <strong>and</strong> [c]) oroptical indicatrix (o~, /3, <strong>and</strong> 3/) are generallyused for plagioclase (e.g. Jensen & Starkey1985; Olsen & Kohlstedt 1985; Ji & Mainprice1988, 1990; Prior & Wheeler 1999).The degree <strong>of</strong> CPO has been quantified usingorientation tensor <strong>of</strong> crystallographic planes<strong>and</strong> directions (Mainprice, ftp://ftp.dstu.univmontp2.fr/pub.TPHY/david/pc).The intensity<strong>of</strong> CPO is given by the I parameter (Lisle 1985):3I = 15/2 x Z(Ei - 1/3) 2i=1where E~, E2, <strong>and</strong> E3 represent the eigenvalues <strong>of</strong>the preferred orientation <strong>of</strong> poles to planes <strong>and</strong>directions <strong>of</strong> amphibole <strong>and</strong> plagioclase grainsplotted in pole figures. The values <strong>of</strong> I rangebetween 0 (no preferred orientation) <strong>and</strong> 5 (allfabric elements perfectly parallel one to another).MicrostructuresThe metagabbros from the eastern (lower) beltare affected by localized shear zones (Fig. lb)<strong>and</strong> all stages from non-deformed rock, protomylonite,mylonite, <strong>and</strong> up to highly strainedultramylonite are present (Figs 2a, b, c <strong>and</strong> d).In the western metagabbro (upper) belt, theprotolith stage is not preserved <strong>and</strong> only twodeformational stages can be distinguished252


TEXTURES OF NATURALLY DEFORMED METAGABBROS101Eastern belt (-650 ~Western belt (-750 ~Fig. 2. Macro photographs <strong>of</strong> typical metagabbro structures from the eastern (a-d) <strong>and</strong> the western (e, t3 belts:(a) non-deformed metagabbro (E0); (b) protomylonite (El); (c) mylonite (E2); (d) ultramylonite (E3); (e) augenmylonite (W1); (f) b<strong>and</strong>ed mylonite (W2).(Figs 2e <strong>and</strong> f): mylonites with hornblendeporphyroclasts related most probably to theCambro-Ordovician metamorphic event <strong>and</strong>completely recrystallized mylonites characterizedby a monomineral layering.Metagabbros <strong>of</strong> the eastern belt consist <strong>of</strong> plagioclase(40-60%), amphibole (40-60%), relictpyroxene in the cores <strong>of</strong> some large amphibolegrains (less than 1%), <strong>and</strong> titanite (less than1%). No substantial change in modal proportions253


102 L. BARATOUX ETAL.is observed with increasing deformation. Modalproportions <strong>of</strong> amphibole <strong>and</strong> plagioclase in thewestern metagabbros are more variable thanthose in the eastern belt. Amphibole proportionmay vary between 20 <strong>and</strong> 80%, most likely dueto original magmatic compositional variations.For the purpose <strong>of</strong> our study, samples composed<strong>of</strong> c. 50% <strong>of</strong> amphibole <strong>and</strong> c. 50% <strong>of</strong> plagioclasewere chosen.Deformation <strong>of</strong> the eastern (lower)metagabbro sheetNon-deformed metagabbro (EO). At low strain,plagioclase <strong>and</strong> hornblende exhibit euhedral r<strong>and</strong>omlyoriented crystals <strong>of</strong> 0.5-3 mm in size.Tapering mechanical twins <strong>and</strong> local undulatoryextinction occur in plagioclase. There is no evidencefor any kind <strong>of</strong> dynamic recrystallizationor crystallization <strong>of</strong> new grains. Undulatoryextinction locally attests to some bending <strong>of</strong>amphibole grains. Amphibole porphyroclastsshow r<strong>and</strong>om spatial distribution <strong>and</strong> they aregenerally not interconnected.Protomylonite (El). At higher strains, about20-25% <strong>of</strong> the total volume <strong>of</strong> plagioclasegrains but only 8-10% <strong>of</strong> amphibole showevidence <strong>of</strong> strain, suggesting that the deformationwas accommodated mostly by plagioclaserecrystallization <strong>and</strong> associated grain-sizereduction. Plagioclase grains show polysynthetictwins according to albite <strong>and</strong> pericline laws(Tullis 1983) <strong>and</strong> patchy undulatory extinction.Large plagioclase porphyroclasts <strong>of</strong> 2-5 mmare cut by brittle fractures (Fig. 3a) reducingthe grain size to 0.5-1 mm. The fractures arefilled with small twin-free recrystallized grains<strong>of</strong> 0.02-0.1 mm. Two recrystallization mechanismshave been identified: bulging <strong>and</strong> subgrainrotation recrystallization (Fig. 3b), as proposedby Poirier & Guillop6 (1979) or Fitz Gerald &StiJnitz (1993), leading to core-mantle structures.Core <strong>and</strong> mantle structures were observed withan intermediate zone <strong>of</strong> subgrains <strong>and</strong> newgrains <strong>of</strong> similar size developed along porphyroclastboundaries. Boundaries between the neoblastsbecome progressively straight, meeting attriple junctions <strong>of</strong> 120 ~ .Large porphyroclasts <strong>of</strong> hornblende revealstrong internal deformation such as kinking,bending leading to sweeping undulatory extinction<strong>and</strong> (100) twinning. The twin planes arenot regular <strong>and</strong> they locally form finger-likestructures. Brittle fractures transecting largegrains are <strong>of</strong>ten present. R<strong>and</strong>omly orientedporphyroclasts <strong>of</strong> 2-5mm in size rotateFig. 3. Drawing <strong>and</strong> micrographs (XPL) <strong>of</strong> the eastern metagabbro protomylonite. (a) Initial stage <strong>of</strong> plagioclasedeformation characterized by brittle fractures cross-cutting large porphyroclasts. Digitized drawing was used for thequantitative textural analysis. Arrows mark three clasts derived by fracturing <strong>of</strong> one porphyroclast <strong>and</strong> showcorresponding grains in the digitized drawing <strong>and</strong> micrograph. Plagioclase is white, hornblende is light grey, <strong>and</strong>opaque minerals are black. (b) New grains develop preferentially along fractures by a mechanism <strong>of</strong> subgrain rotation(SR). Note that the neoblasts are twin-free. (P) refers to porphyroclast.254


TEXTURES OF NATURALLY DEFORMED METAGABBROS 103progressively their cleavage planes { 110} parallelto the foliation. Small needle-like grains (0.02-0.05 x 0.1-0.3 mm in size), arranged parallelto the mylonitic foliation, are present in highstraindomains.Mylonite (E2). Within the eastern metagabbros,plagioclase porphyroclasts are recrystallized intoelongate aggregates or b<strong>and</strong>s wrapped aroundhornblende clasts, <strong>and</strong> porphyroclasts representonly 5-10% <strong>of</strong> the total plagioclase volume.Elongate needle-like hornblende grains arearranged into aggregates or b<strong>and</strong>s parallel to themain mylonitic foliation (Fig. 4a). Isolated hornblendeneedles scattered within plagioclase-richdomains are inclined at an angle <strong>of</strong> 20-30 ~ withrespect to the main fabric, indicating noncoaxialdeformation (Figs 4a <strong>and</strong> b). Plagioclasegrains attain 0.1-0.25 mm in size. Increase ingrain size with respect to the protomylonitestage <strong>and</strong> common optical zoning <strong>of</strong> recrystallizedgrains due to increase in anorthite contentsuggest syndeformational growth. Matrix grainsare <strong>of</strong>ten twin-free; the twinned grains representFig. 4. Digitized drawings <strong>and</strong> micrographs <strong>of</strong> the eastern mylonite. (a) Mylonitic foliation with alternating layers <strong>of</strong>amphibole <strong>and</strong> plagioclase aggregates (PPL). (b) Plagioclase <strong>of</strong> subequant shapes <strong>and</strong> grain size typical for themylonitic stage (E2-pll). (c) High-strain plagioclase domain (E2-p12) adjacent to an amphibole porphyroclast ischaracterized by elongate shapes <strong>and</strong> strong SPO (XPL). (d) Amphibole porphyroclast (E2) cross-cut by microshearzones filled by small needle-like grains (XPL). Undulatory extinction documents strong internal deformation. Arrowsmark corresponding grains in the digitized drawings <strong>and</strong> micrographs.255


104 L. BARATOUX ETAL.only 10-20% <strong>of</strong> the total plagioclase volume.Plagioclases generally exhibit subequant shapeswith straight grain boundaries meeting at 120 ~triple joints (Fig 4b). However, in the vicinity<strong>of</strong> amphibole porphyroclasts, the plagioclasegrains are more elongate with higher aspectratios up to 2.5 (Fig. 4c).Hornblende grains are arranged in anastomosinginterconnected b<strong>and</strong>s defining the myloniticfoliation. Most <strong>of</strong> the hornblende matrix grains,0.1-0.3 mm in length, are elongate in XZ sectionsattaining high aspect ratios (up to 6) whilethey are lozenge-shaped in the YZ sectionswith aspect ratios <strong>of</strong> 1.5-3. A total <strong>of</strong> 10-20%<strong>of</strong> the hornblende is present in a form <strong>of</strong>locked-up porphyroclasts <strong>of</strong> 0.2-2 mm in sizeshowing undulatory or patchy extinction characteristic<strong>of</strong> strong internal strain. These porphyroclastsare elongate parallel to the foliation.Porphyroclasts are <strong>of</strong>ten transected by microshearzones (Fig. 4d) <strong>and</strong> small needle-shapedgrains consequently develop within these zonesby activation <strong>of</strong> the weak cleavage system{110}.Ultramylonite (E3). The ultramylonitic metagabbros<strong>of</strong> the eastern sheet are marked by veryuniform mineral b<strong>and</strong>ing <strong>of</strong> fine-grained hornblende<strong>and</strong> plagioclase layers ranging between 1 <strong>and</strong>5 mm in width (Fig. 5a). Asymmetric fabric <strong>of</strong>hornblende <strong>and</strong> plagioclase is no longer present.Plagioclase forms continuous b<strong>and</strong>s includingabundant interstitial grains <strong>of</strong> amphiboles(Fig. 5b). In contrast, amphibole-rich b<strong>and</strong>s arealmost monomineral. Plagioclase is completelyrecrystallized, varying in size between 0.05 <strong>and</strong>0.3mm. The grains have smooth roundedshapes locally elongate parallel to the foliation.Rounded grains <strong>of</strong> 0.5 mm in size, interpretedas relicts <strong>of</strong> original porphyroclasts, locallyoccur in the fine-grained matrix. Plagioclasegrains are affected by strong retrograde sericitization<strong>and</strong> albitization.Interstitial <strong>and</strong> matrix grains <strong>of</strong> hornblendereach 0.03-0.1 x 0.1-0.3 mm in size. Lockedupporphyroclasts <strong>of</strong> amphibole (0.5-1 mm insize) occur parallel to the foliation. However,they are less common than in the mylonite(E2). All amphibole grains are characterized byvery strong SPO (Fig. 5c).The mylonitic foliation is cross-cut by extensionalveins filled by a mixture <strong>of</strong> epidote <strong>and</strong>amphibole. These veins show that fluid activitywas high after the main mylonitic episode. Albitization<strong>and</strong> sericitization <strong>of</strong> plagioclase in somesamples suggest that increased amount <strong>of</strong> waterwas present in these rocks also during late retrogression,which is most likely not related to themain process <strong>of</strong> mylonitization.Deformation <strong>of</strong> the western (upper)metagabbro sheetAugen mylonite (W1). In the western metagabbrobelt, the deformation was pervasive sothat initial protomylonite stages <strong>of</strong> deformationFig. 5. Drawings <strong>and</strong> micrograph <strong>of</strong> the eastern ultramylonite. (a) Plagioclase <strong>and</strong> amphibole b<strong>and</strong>ing (PPL).Plagioclase is darker than amphibole due to strong sericitization. (b) Plagioclase b<strong>and</strong> in ultramylonite (E3)with abundant interstitial hornblende. Small grain size <strong>and</strong> subequant shapes are typical for these plagioclases.(c) Amphibole layer from ultramylonite (E3) is characterized by very fine grain size <strong>and</strong> strong SPO.256


TEXTURES OF NATURALLY DEFORMED METAGABBROS 105are absent. The least deformed rock representedby augen mylonite is marked by the presence<strong>of</strong> rare 2-3 mm large porphyroclasts <strong>of</strong> plagioclase(Fig. 6a) <strong>and</strong> by 1-2ram wide <strong>and</strong>4-5 mm long porphyroclasts <strong>of</strong> hornblendewithin the recrystallized matrix (Fig. 6b). Thisrock type is preserved only in the southern part<strong>of</strong> the metagabbro belt. Plagioclase porphyroclastsrepresent only 5-10% <strong>of</strong> the total plagioclasevolume <strong>and</strong> show deformational twins<strong>and</strong> undulatory extinction. Recrystallized plagioclasegrains show optical zoning.Modal proportion <strong>of</strong> amphibole porphyroclastscannot be estimated as they differ fromthe neoblasts only in their chemical composition,making them optically indiscernable. Amphiboleporphyroclasts exhibit patchy or sweeping undulatoryextinction but deformation twins have notbeen observed. The local occurrence <strong>of</strong> twins isassumed to be growth-related, the twin planesbeing always perfectly straight. Prismatic porphyroclastsattain higher aspect ratios thanmore rounded neoblasts. All rocks in thewestern belt are affected by amphibolite faciesre-equilibration partly destroying the granulitefacies peak metamorphic assemblages.Fig. 6. Drawings <strong>and</strong> micrographs <strong>of</strong> the western augenmylonite. (a) Digitized drawing <strong>of</strong> plagioclase (W1)used for the quantitative textural analysis. Largeporphyroclasts are surrounded by a mantle <strong>of</strong> finegrainedmatrix grains. (b) Hornblende (W 1) showsvariable grain size <strong>and</strong> strong SPO. Plagioclase is white,hornblende is light grey, <strong>and</strong> opaque minerals are black.B<strong>and</strong>ed mylonite (W2). Distinct monomineralb<strong>and</strong>s <strong>of</strong> 1-10 mm in width are typical <strong>of</strong> thesehighly deformed mylonites. Small hornblendegrains <strong>of</strong> irregular shape are common withinthe plagioclase layers, but plagioclase is rarelyincluded within amphibole b<strong>and</strong>s. Both mineralsare completely recrystallized. Hornblende is insome places replaced by later prisrnatic cummingtonite,which grows either parallel orobliquely to the mylonitic foliation <strong>and</strong> islocally included within the plagioclase layers.Two types <strong>of</strong> plagioclase microstructuresoccur within one plagioclase b<strong>and</strong>. The firsttype is characterized by grain sizes rangingfrom 0.3 to 1 mm in diameter, subequant orslightly elongate, but always irregular, grainshapes with serrated boundaries (Fig. 7a), <strong>and</strong>growth-related optical zoning. Mechanicaltwins cross-cut the optical zoning in somecases (Fig. 7b), suggesting that at least some <strong>of</strong>them formed later. The second type is characterizedby narrow (c. 0.5 mm) high-strain zonestrending parallel or slightly oblique to the layering(Fig. 7c). Here, the grain size is reduced to0.05-0.2 ram. Plagioclase grains attain higheraspect ratios (up to 2) <strong>and</strong> exhibit undulatoryextinction <strong>and</strong> subgrain boundaries elongate parallelto the high-strain zone. The contact betweenthese two deformational domains is either sharpor transitional.Hornblendes <strong>of</strong> 0.1 - 1 mm in size have straightgrain boundaries locally meeting at high anglessuggesting a high degree <strong>of</strong> textural equilibration(Vernon 1976) (Fig. 7d). No undulatory extinctionoccurs <strong>and</strong> twinning is very rare. Wherepresent, the twin planes are always very straight,indicating that twinning is growth-related ratherthan deformational. Aspect ratios vary between1.5 <strong>and</strong> 2.5, but some grains with aspect ratioclose to 1 occur. Grains with high aspect ratiosare not always parallel to the mylonitic foliation.Bulk rock chemistry, mineral chemistry<strong>and</strong> zoningBulk rock chemistryTable 1 presents bulk rock chemical <strong>analyses</strong>studied by X-ray fluorescence at ClaudeBernard University, Lyon, France. The eastern<strong>and</strong> western metagabbro sheets are regarded tobe a part <strong>of</strong> the same lower crustal unit prior totheir involvement into the Variscan tectonics(Stfpskfi et al. 2001). However, their chemistryis not identical <strong>and</strong> bulk rock <strong>analyses</strong>(Table 1) show that western metagabbros arecharacterized by higher contents in A1 <strong>and</strong> Ca<strong>and</strong> lower contents in Na <strong>and</strong> Fe than their257


106 L. BARATOUX ET AL.Fig. 7. Drawings <strong>and</strong> micrographs <strong>of</strong> the western b<strong>and</strong>ed mylonite. (a) Lobate boundaries document grain boundarymigration (GBM) in coarse-grained plagioclases <strong>of</strong> b<strong>and</strong>ed mylonite (XPL). (b) Mechanical twins (MT) transect thegrowth-related zoning (GZ) in b<strong>and</strong>ed mylonites documenting a later deformation phase (XPL). (c) Digitized drawing<strong>of</strong> two plagioclase domains (W2-pll <strong>and</strong> W2-p12) used for quantitative textural analysis. (d) Hornblendes (W2) arecharacterized by straight grain boundaries meeting occasionally at triple points at variable angles. Many small grainsindicating high nucleation rate are present. SPO is rather low. Plagioclase is white, hornblende is light grey, pyroxeneis dark grey, <strong>and</strong> opaque minerals are black in all drawings.eastern counterpart. The increase in Na <strong>and</strong>depletion in Ca in both metagabbro belts isrelated either to mylonitization or to primaryvariations due to postmagmatic processes.Trace elements show very similar trends forthe eastern <strong>and</strong> western metagabbros, suggestingthat they may originate from the same source.Metagabbros from both belts display a slightlynegative Nb anomaly normalized to MORB(Sun & McDonough 1989). Augen mylonitesfrom the western belt are depleted in Zr <strong>and</strong> Ticompared to eastern mylonites. Contents <strong>of</strong> Rb,Ba, <strong>and</strong> K are strongly variable in both belts,which is most probably related to postmagmaticalterations <strong>and</strong>/or mobility <strong>of</strong> these elementsduring metamorphism.Mineral chemistry <strong>and</strong> zoningSyndeformational chemical reactions arecommon in metabasic rocks (Brodie 1981;Brodie & Rutter 1985). Variations in plagioclasecomposition are shown in Figures 8 <strong>and</strong> 9.Amphibole compositions according to the classification<strong>of</strong> Leake et al. (1997) are plotted inFigure 10. Representative <strong>analyses</strong> <strong>of</strong> mineralcompositions are listed in Tables 2 <strong>and</strong> 3.Mineral abbreviations are according to Kretz(1983).It is likely that the studied metagabbros experienceda metamorphic event prior to the maindeformation <strong>and</strong> metamorphism studied in thiswork. This is supported by existence <strong>of</strong> incompletelyamphibolized pyroxenes in undeformedrocks <strong>of</strong> low-strain domains. In addition, thecompositions <strong>of</strong> cores <strong>of</strong> large amphibole porphyroclasts(actinolites to actinolitic magnesiohornblendes)deviate from those <strong>of</strong> recrystallizedmagnesio-hornblendes in the eastern belt. Suchcompositions <strong>of</strong> primary amphiboles may indicateearly greenschist to amphibolite facies conditionspreceding the main higher gradeCarboniferous deformation.The peak metamorphic assemblages within theeastern metagabbro consist <strong>of</strong> P1 + Hbl + Qtz +Ttn _+ Ilm __%Mag. The composition <strong>of</strong> plagioclaseporphyroclasts in protomylonite (El)varies between Anso <strong>and</strong> An6o (Fig. 8a). Smallrecrystallized grains show a composition similarto the mother host grain corresponding toAnso-6o. More albitic compositions (An4o_45)occur along grain boundaries or triple junctions<strong>of</strong> recrystallized grains (Fig. 9a). The geometry<strong>and</strong> sharp gradient in mineral zoning crosscuttingseveral grains is attributed to post-peak258


TEXTURES OF NATURALLY DEFORMED METAGABBROS107Table 1. Representative <strong>analyses</strong> <strong>of</strong> bulk rock compositionsEastern beltWestern beltProtolith (E0) Mylonite (El) Ultramyl. (E3) Augen myl. (Wl) B<strong>and</strong>ed myl. (W2) B<strong>and</strong>ed myl. (W2)SiO2 48.61 49.20 52.69 46.23 50.20 48.92TiO2 0.71 0.82 0.94 0.10 0.35 0.37A1203 17.76 19.74 18.04 21.57 21.43 24.20FeO T 6.14 6.32 7.33 4.86 4.10 3.84MnO 0.12 0.10 0.13 0.10 0.08 0.06MgO 9.31 6.56 5.81 7.91 6.20 4.76CaO 12.47 9.22 8.60 13.15 12.08 11.58Na20 2.64 3.29 4.99 0.82 2.98 3.55K20 0.16 1.52 0.10 1.60 0.51 0.25P205 0.04 0.03 0.03 0.01 0.02 0.07LOI 0.88 1.91 0.57 2.76 1.21 1.65H20- 0.06 0.12 0.05 0.19 0.09 0.10Total 98.90 98.83 99.28 99.30 99.25 99.35Ba 32.1 149.0 52.0 374.3 52.3 54.6Rb 1.6 55.1 4.9 74.7 26.9 11.9Sr 409.4 742.0 226.9 219.1 346.3 397.4Zr 33.0 34.7 39.3 7.8 10.7 29.6Nb 1.3 1.3 1.7 0.6 1.3 1.2Y 12.2 9.4 20.2 3.3 7.2 6.4V 185.9 160.1 203.9 115.0 131.9 92.0Cr 550.5 214.7 222.4 136.0 622.9 269.9Ni 122.8 60.6 58.8 48.9 87.3 90.9CO 36.8 35.4 29.2 32.0 26.6 25.4Sc 42.3 24.9 30.4 41.3 33.6 16.8Cu 8.0 3.6 19.1 2.9 20.4 33.3Pb 0.8 4.5 2.5 6.0 2.4 2.3Ga 13.8 15.0 12.2 12.8 16.9 17.1XMg* 0.730 0.649 0.586 0.744 0.729 0.688Major elements are given in wt%, trace elements are given in ppm. *XMg = Mg/(Fe 2+ + Mg).a) 12- Porphyroclasts Eastern belt b) Western beltPorphyroclaststrogressionII IN!I IIIII8 i . . . . i i i , i , i i , i i , i ig o I] I] , I ,I III~_ Recrystallized grains ~ protomylonite (El) Recrystallized grains6 I ~ mylonite(E2) I r~ ~ Augen mylonite (W1)1Late retrogression [Z] ultramylonite (E3)I---1 B<strong>and</strong>ed rnylonite (W2)l2HI I, 111 t h10 20 30 40 50 60 70 80 90 10 20 30 40 50 60 70 80 90% An % AnFig. 8. Composition <strong>of</strong> plagioclase in (a) the eastern <strong>and</strong> (b) the western belts. Porphyroclasts are depicted separatelyfrom matrix grains. Deformation grade increases from dark grey to white colour.259


108 L. BARATOUX ET AL.Fig. 9. BSE images <strong>of</strong> plagioclase <strong>and</strong> amphibole<strong>and</strong> representative chemical maps <strong>of</strong> plagioclasecompositional variations. (a) Plagioclase inprotomylonite from the eastern belt (El) suffers fluidrelatedchemical variations modifying the compositionat triple points <strong>and</strong> grain boundaries. (b) Growth-relatedchemical zoning in mylonite from the eastern belt (E2).(c) Growth-related zoning in coarse-grained plagioclase<strong>of</strong> b<strong>and</strong>ed mylonite from the western belt (W2).fluid-phase activity along grain boundaries(McCaig & Knipe 1990). Recrystallized matrixplagioclases in the mylonite (E2) have <strong>and</strong>esiticcompositions (An3o-5o) <strong>and</strong> show increases inanorthite content towards their rims (Fig. 9b),which is commonly ascribed to increasingmetamorphic conditions. Rare rounded porphyroclast,as well as matrix grains in ultramylonites(E3), underwent strong <strong>and</strong> irregularly developedlate albitization. Matrix grains correspond toAn3o_43 although oligoclase <strong>and</strong> albite compositions(An6_ 15) are also common along fractures,cleavages <strong>and</strong> grain boundaries. These irregularcompositional variations are attributed to late retrogressionunrelated to the main deformationprocess.Amphibole porphyroclasts in protomylonites(El) are magnesio-hornblendes with variablecontent in Si but constant XMg (Fig. 10a).Amphiboles with more actinolitic compositionreplacing former magmatic pyroxenes can befound in the cores <strong>of</strong> some grains. Small amphibolegrains show similar compositions comparedto the porphyroclasts, some <strong>of</strong> them being moretschermakitic. The amphiboles in mylonites(E2) are magnesio-hornblendes with slightlylower Si contents than those in the protomylonites.High Si/low A1 domains documentingincomplete transformation <strong>of</strong> pyroxenes intohornblendes are present in the cores <strong>of</strong> large porphyroclastsas well. Locked-up grains in theultramylonites (E3) vary between magnesiohornblende<strong>and</strong> tschermakitic composition.Small new grains have slightly lower A1 contentscompared to the less deformed stages.In the western belt, the peak metamorphicassemblages correspond to P1 § Hbl _ Cpx _+Opx +_ Grt ___ Ttn + Ilm. A granulite faciesmineral assemblage, comprising also sapphirine<strong>and</strong> corundum, is not stable <strong>and</strong> re-equilibratedduring subsequent upper amphibolite faciesreworking. In the south, granulite facies assemblagesare absent <strong>and</strong> probably completelyre-equilibrated. In the north, a late metamorphicphase is documented by the presence <strong>of</strong> prismaticorthoamphibole (gedrite).Plagioclase porphyroclast compositions arevery similar to those <strong>of</strong> the eastern belt <strong>and</strong> varybetween An4o <strong>and</strong> An65. Recrystallized matrixgrains in the augen mylonite (W1) show locallyincreasing anorthite content up to An6o_9o comparedto the porphyroclasts, indicating thatchemical reactions took place during recrystallization.Sharp limited domains <strong>of</strong> An4o_45 compositionalong grain boundaries are also common,suggesting late fluid activity (Knipe & McCaig1994). Recrystallized plagioclase grains inb<strong>and</strong>ed mylonites (W2) show stronger variabilityin compositions (An35-9o). Growth-relatedzoning due to increasing metamorphic grade isdocumented in Figure 9c. The cores have morealbitic composition (An34) compared to the rims(An6o_65) <strong>and</strong> are attributed to the early metamorphicstage. Modifications <strong>of</strong> plagioclasecomposition characteristic <strong>of</strong> fluid-related compositionalchanges are also developed in theb<strong>and</strong>ed mylonite.Amphibole porphyroclast compositions (Fig.10b) were analysed in two different thin sections<strong>of</strong> augen mylonite (W1) <strong>and</strong> the difference intheir mineral chemistry may be explained bylocal bulk rock chemical variations. The amphiboleporphyroclasts correspond to magnesiohornblendeswith similar compositions to thosein the eastern belt. Porphyroclast rim compositionsshow a decrease in Si content withrespect to the cores. Neoblasts are marked260


1.0--0.9__Mg~(Mg2++Fe2+)9 Core <strong>of</strong> porphyroctast9 Rim <strong>of</strong> porphyroclastb, Core <strong>of</strong> neoblasto Rim <strong>of</strong> neoblastTremoliteActinolileTEXTURES OF NATURALLY DEFORMED METAGABBROS 109I1,00.9TremoliteAclinoliteMagnesio-homblendefoo:oI Core <strong>of</strong> o oyroclast~Magnesio-hornblende ! TschermakiteCore <strong>of</strong> A i oporphyroclast [....Tschermakite9 9 A9 A,~ Ultramylonite (E3)i Protomylonite (El)I]-schermakiteMylonite (E2)irrotschermakite} . - - - -{ 5,75!iI~rrotschermakite5.75O.C8.00Ferro-actinolite7.50Ferro-hombtendeFerrotschermakite i6.50 5,751.0 i0.9iTremoliteActinoliteMagnesio-hornblende ~, ~.~ Tschermakiteo o ,.1~, ~ oo ~1.0I 0.9_Ng~ _(Mg2++Fe2+)[:008.004TremoliteActino}iteFerro*actinoliteMagnesio-hornblendeTschermakitePorphyroctast 1 ~ . . _ ~oO~ooPorphyroclast 2 ~ o o.... 0,9o-- -Augen mylonite (Wl)Ferro-homblendeFerrotschermakite7.50 )6.50 5.75~.50B<strong>and</strong>edmylonite (W2)Ferrotschermakite5,75Fig. 10. Composition <strong>of</strong> amphiboles in (a) the eastern <strong>and</strong> (b) the western belts. (Classification <strong>of</strong> Leake et al. (1997)was used, (Na + K)A < 0.5.)by sharp compositional differences expressedby decreasing Si <strong>and</strong> XMg contents. In theb<strong>and</strong>ed mylonite (W2), larger grains showhigher Si content than small neoblasts (verylow Si content), both being <strong>of</strong> tschermakiticcomposition.Temperature (T) was calculated using thehornblende- plagioclase thermometer byHoll<strong>and</strong> & Blundy (1994; thermometer B).Estimated temperatures are based on 30 <strong>and</strong> 40amphibole-plagioclase couples in the eastern<strong>and</strong> western belt, respectively. In the eastern(lower) sheet, calculated T was estimated tobe 650 4-50~ while T in the west was750 + 50 ~ Deformation is interpreted to havetaken place under or close to these metamorphicconditions. Pressure (P) could not be estimated asgarnet is absent in the mineral assemblage <strong>of</strong> themetagabbros. However, P <strong>of</strong> 9 kbar was estimatedat the lower contact between the metagabbrosheet <strong>and</strong> tonalitic sill, where garnetshave been formed (St/pskfi et al. 2001).261


110 L. BARATOUX ETAL.r~t'-,IMdMdddd~dd ~ d ~ d d d d d d ~r~


TEXTURES OF NATURALLY DEFORMED METAGABBROS 111e4Mo('4,,.0 ~ I'~ ~0 oo e4 oo ~1- P--. e~h ~"~ ',4~ ~ ',~ t"q ~0 ',,0 0", tt3 c4 0", ,--~9~-~.~~84-~dd dd,~,~d~ ~dd,~ddd4~dddII-F+d4-o.1...o0~8r~~ .~ . ~ .qq ~ . q ~ . ~ .~263


112 L. BARATOUX ET AL.<strong>Quantitative</strong> textural analysisGrain size distributionThe average grain sizes are presented as medianvalues <strong>of</strong> the Ferret diameter <strong>and</strong> the grain sizespread is expressed by the difference <strong>of</strong> third<strong>and</strong> first quartile (Q3- Ql). The statisticalvalues <strong>of</strong> the quantitative micro<strong>structural</strong> <strong>and</strong>grain size <strong>analyses</strong> are summarized in Table 4.Plagioclase. Plagioclase in XZ section <strong>of</strong> theprotomylonitic stage in the eastern metagabbrobelt is characterized by an average grain size <strong>of</strong>45 ixm (El) (Fig. l la). Matrix grains in mylonitesare marked by an increase <strong>of</strong> the averagegrain size, reaching 74 pore in less (E2-pll) <strong>and</strong>58 Ixm in more (E2-p12) strained domains.Grain coarsening in the mylonite is consistentwith the syndeformational growth <strong>of</strong> grains originatedfrom subgrain rotation in the protomyloniticstage. This is supported by a greater amount<strong>of</strong> larger grains in grain size frequency histograms.The grain size is reduced to ~45 Ixm inthe ultramylonite (E3).Plagioclase grains in the augen mylonite <strong>of</strong>the western metagabbro belt (W1) are 43 Ixmon average, which is the statistical valueTable 4, Statistical values <strong>of</strong> the quantitative textural analysisEastern beltWestern beltProtomyl. Mylonite Mylonite Ultramyl. Augen myl. B<strong>and</strong>ed myl. B<strong>and</strong>ed myl.pl amp, pll p12 amp, pl amp, pl amp, pll p12GBPO Pl-pl 1.23 1.30 1.58 1.44 1.39 1.24 1.59Eigenvalue Amp-amp 2.50 3.53 2.26 1.96r = el/e2 Amp-plt 1.41 1.94 2.60 2.18 1.55 1.50Eigenvector V1 Pl-pl 7 24 23 17 3 - 3 5Orientation (~ Amp-amp 11 2 4 - 2Amp-pl * 16 12 18 7 -5 1SPO P1 1.45 1.46 1.73 1.85 1.86 1.38 1.64Eigenvalue Amp 1.83 4.40 2.89 2.21r = el/ea Amp t 1.81 2.67 3.43 3.26 1.96 1.76Eigenvector V1 P1 - 14 21 22 8 1 2 8Orientation (~ Amp 16 2 8 - 2Amp t 18 9 21 8 - 1 9Aspect ratio P1 1.59 1.56 1.82 1.84 1.68 1.50 1.72(median) Amp 4.46 3.79 2.36 2.12Amp ~ 2.67 2.59 3.02 3.00 2.28 1.84Grain size-Ferret diameterMedian (Ixm) P1 45 74 58 45 43 92Amp 49 41 119 94Amp t 49 39 43 33 59 50Q1 (~m) P1 33 52 43 30 30 57Amp 36 30 62 61Amp t 38 29 31 24 34 35Q3 (~m) P1 63 103 80 72 63 147Amp 68 62 211 142Amp t 64 52 60 44 102 74Q3 - QJ (~m) P1 30 51 37 42 32 90Amp 33 32 150 81Amp t 26 23 29 20 68 39Skewness "+ PI 1.44 - 0.34 - 0.08 - 0.09 1.15 0.01Amp 0.62 0.25 0.02 0.36Kurtosis * P1 7.30 2.86 2.77 2.53 6.99 2.98Amp 4.86 3.12 2.43 2.74R 2 P1 0.9151 0.9909 0.9979 0.9951 0.9442 0.9977correlation Amp 0.9768 0.9942 0.9899 0.9883coef.*Positive values are oriented anticlockwise with respect to the horizontal line.tin plagioclase domains.*The R 2 values, skewness close to zero, <strong>and</strong> kurtosis close to 3 signify well-fitted lognormal distribution.264725010656


' ~'-----2~--~----TEXTURES OF NATURALLY DEFORMED METAGABBROS 113b)Amphibole~.. ~ Amphibole in plagioclase domainsAmphibole,~ Amphibole in plagioclase domains\ A0 50 100 150Grain size - Ferret diameter (pm)9 Protomylonite- E1"~ A Mylonite- E2-pllLU z~ Mylonite- E2-pl2/x Ultramylonite- E39 ~ 9 Augen mylonite- Wlr o B<strong>and</strong>ed mylonite- W2-pllo B<strong>and</strong>ed mylonite- W2-p12Plagioclaseo50200 181614~8'Plagioclase ~ East ' "2Em Protomylonite - E1I. ~ Mylonit~- E2ttl [~ Ultramylonite- E36420fill L.. Plagioclase - West~21]il-grained plagioclasedlllllllllllllllilLlUL J~nn n~l][]nlIJ]n~Amphibole - Eastt 3~I ~ Mylonite- E2b9 11~ I I Ultramylonite- E3llilJLAmphibole - Westm Augen mylonite W1- [~ B<strong>and</strong>ed mylonite- W250 100 150 200 250 3000 50 100 150 200 250 300 350 400Grain size - Ferret diameter (~m)Grain size - Ferret diameter (p.m)Fig. 11. Grain size evolution in the eastern (triangles) <strong>and</strong> western (circles) metagabbros. Degree <strong>of</strong> defomlationincreases from dark to white colour. (a) Calculated medians <strong>and</strong> st<strong>and</strong>ard deviations. (b) Histograms showing thecharacteristic frequencies <strong>of</strong> grain size for each deformational stage.representing the size <strong>of</strong> the recrystallized matrixgrains. Plagioclase grain size is substantiallyhigher in the coarse-grained domain (W2-pll)<strong>of</strong> the b<strong>and</strong>ed mylonite (92 ~zm). The grain sizeis reduced to 72 ~m on average in the highstraindomain (W2-p12).The grain size distributions (Fig. 1 lb) in protomylonites<strong>of</strong> the eastern belt (El) show slightlybimodal distribution due to presence <strong>of</strong> largeporphyroclasts surrounded by small recrystallizedgrains. A higher amount <strong>of</strong> large grainscharacterizes the western b<strong>and</strong>ed mylonite(W2) compared to the augen mylonite (W1)<strong>and</strong> eastern mylonites (El-E3). Lognormal distributionscorrelate with intensity <strong>of</strong> deformationin both belts.Amphibole. The amphibole grain size is 49 p~min the eastern metagabbro mylonite (E2) <strong>and</strong>41 ~m in the ultramylonite (E3). However, thegrain size spread is always lower than that <strong>of</strong>plagioclase in the same rock.The size <strong>of</strong> amphibole grains included withinplagioclase domains decreases systematicallywith increasing deformation, from 49 ~m in theleast deformed sample (El) to 39 ~m <strong>and</strong>43 ~m in the mylonite (E2-pll <strong>and</strong> E2-p12,respectively) <strong>and</strong> 33 p~m in the ultramylonite(E3). The grain sizes are always slightly lowerthan those <strong>of</strong> amphibole matrix grains in thesame thin section.Amphibole grain size in the western metagabbrobelt is highest in the sample including porphyroclasts(W1), reaching 119 Fm on average.Very high variance is typical <strong>of</strong> this mylonite,which is due to a relatively low number <strong>of</strong> newgrains <strong>and</strong> a high number <strong>of</strong> large porphyroclasts.In the b<strong>and</strong>ed mylonites (W2), grain sizedecreases to 94 Fm on average.In the eastern belt, amphibole grain size distributions(Fig. 11 b) are similar to those <strong>of</strong> plagioclase.They are characterized by well-developedlognormal distribution in both mylonite <strong>and</strong>ultramylonite stages. In contrast, the distribution<strong>of</strong> amphibole grain size in the western belt exhibitsweaker fits <strong>of</strong> lognormal distribution withincreasing deformation, which is attributed to ahigher number <strong>of</strong> large grains <strong>of</strong> variable size(Table 4).Shape preferred orientationPIagioclase. Recrystallized plagioclase grainsin the protomylonite <strong>of</strong> the eastern belt (El)<strong>and</strong> mylonite (E2-pll) have aspect ratio (AR)<strong>of</strong> 1.59 <strong>and</strong> 1.56, respectively (Fig. 12a). Bothaspect ratio (AR = 1.82) <strong>and</strong> SPO increase inthe analysed high-strain plagioclase aggregateadjacent to the rigid amphibole porphyroclastin mylonite (E2-pl2). The ultramylonite (E3)exhibits further strengthening <strong>of</strong> the SPO <strong>and</strong>similar high aspect ratios (AR = 1.84).265


114 L. BARATOUX ETAL.o ._,


TEXTURES OF NATURALLY DEFORMED METAGABBROS 115maximum eigenvector V1 is only slightlyinclined with respect to the foliation, suggestingminor component <strong>of</strong> non-coaxial deformation. Inthe western belt, ramp_am p weakens from 2.26 inthe augen mylonite (W1) to 1.96 in the b<strong>and</strong>edmylonite (W2), being still stronger than rpl-plin similar rock types. The eigenvector V~ is parallelto the foliation in both types <strong>of</strong> mylonites.Crystallographic preferred orientationFive thin sections, covering all the deformationstages, were analysed using the EBSD technique.In the eastern mylonite (E2) <strong>and</strong> the westernb<strong>and</strong>ed mylonite (W2), two plagioclase domainswere investigated.Besides pole figures <strong>of</strong> [100] directions <strong>and</strong>poles to (010) <strong>and</strong> (001) planes, inverse polefigures, calculated parallel to stretching lineation<strong>and</strong> pole to foliation, are used in the eastern belt inorder to identify orientation maxima, which couldcorrespond to less common slip systems reportedin the literature. The inverse pole figures may beemployed because the orientations <strong>of</strong> lineation<strong>and</strong> foliation follow strong SPO <strong>and</strong> CPO <strong>of</strong>amphibole. For the description <strong>of</strong> inverse polefigures, only the slip systems previouslydescribed in the literature will be mentioned(for summary see e.g. Kruse et al. 2001 ; Sttinitzet al. 2003). The I parameters representing theintensity <strong>of</strong> CPO are shown in Table 5.PlagioclaseWithin the protomylonite <strong>of</strong> the eastern belt (El),the plagioclase CPO is not r<strong>and</strong>om (Figs 13 <strong>and</strong>14a). In the inverse pole figure, the lineations aredistributed subparallel to the horizontal planewith a maximum around [i 10]. Poles to foliationare arranged in a N-S girdle with maximaaround (001) <strong>and</strong> a* (100). The (001) [1 i0] slipsystem (Olsen & Kohlstedt 1984, 1985) couldTable 5. Intensity I <strong>of</strong> crystallographic preferredorientationE1 1.51tE2-pl 1E2-p12E3 2.906Wl 1.422W2-pll 0.948W2-p12HornblendePlagioclase(100) [001] {110} [1001 (010) (001)1.926 0.897 0.116 0.666 0.3190.095 0.109 0.1000.108 0.347 0.1711.591 1.454 0.083 0.033 0.1781.854 0.846 0.111 0.074 0.0201.606 0.653 0.457 0.552 0.9480.590 0.299 0.741thus be possibly active. The orientation <strong>of</strong> porphyroclastsis different from that <strong>of</strong> recrystallizedgrains, suggesting that the recrystallization wasnot host-controlled (e.g. Ji & Mainprice 1990;Kruse et al. 2001). The mylonite (E2-pll <strong>and</strong>E2-p12) shows weak CPO in both plagioclasedomains. A weak maximum <strong>of</strong> lineations is situatedaround [110] in the E2-pll _domain, <strong>and</strong> someclusters <strong>of</strong> lineation around [1_12] directions <strong>and</strong>poles to foliation close to (201) can be observedin the high-strain domain in sample E2-p12. Theseorie_ntationsmay indicate the possible activation <strong>of</strong>(201)1/21112] (Marshall & McLaren 1977a, b) inthe high-strain fine-grained domain adjacent to thehornblende porphyroclast (Fig. 4d). The ultramyloniteCPO (E3) displays maximum for polesto foliation clustering near the (021) <strong>and</strong> maximumfor lineations near the [112] direction. SuchCPO would suggest activation <strong>of</strong> (021)1/21112]slip system (Olsen & Kohlstedt 1984; Montardi& Mainprice 1987). The I parameters attain lowervalues in more deformed stages for [100], (010),<strong>and</strong> (001), respectively.The CPO <strong>of</strong> the augen mylonite in the westernbelt (W1) (Fig. 15) is very weak <strong>and</strong> no slipsystems could be extracted even by examiningthe inverse pole figures, which are not shownhere. Within b<strong>and</strong>ed mylonite (W2-pll <strong>and</strong> W2-p12), both coarse-grained (pll) <strong>and</strong> fine-grained(p12) plagioclase domains exhibit strong maxima<strong>of</strong> [100] directions close to the lineation. Polesto (010) <strong>and</strong> (001) planes are distributed close tothe pole <strong>of</strong> the foliation, which is defined by compositionallayering. The domain W2-p12 representslate shear zones marked by importantgrain refinement (Fig. 7c). The 1 parameters arehigher in the b<strong>and</strong>ed mylonite (W2) comparedto the augen mylonite (W1), indicating strengthening<strong>of</strong> plagioclase CPO. It is likely that thislate shearing was not entirely coaxial with respectto the main flow represented by coarse-grainedaggregate fabrics (W2-pl 1). Therefore, the difference<strong>of</strong> (010) <strong>and</strong> (001) maxima in pole figurescould result from slightly oblique orientation <strong>of</strong>late shear plane with respect to dominant compositionallayering. The shear zone (W2-p12) ismarked by an inclination <strong>of</strong> [100] maximum <strong>of</strong>about 10 ~ with respect to the lineation, whichcould be attributed to non-coaxial deformation (Jiet al. 1988). Activity <strong>of</strong> slip systems (001)[100](Marshall & McLaren 1977a, b) <strong>and</strong> (010)[100](Montardi & Mainprice 1987) may be inferredfrom such orientation patterns.AmphiboleSmall recrystallized hornblendes adjacent totheir host grain in the eastern protomylonite267


116 L. BARATOUX ETAL.Plagioclase - eastern belt[1001 (010) (001).,7, ...... .. . . . . . ~,,. . :~':. .' ..~_s kQ...'........,....._~...:/r ....' .:.-,~.~ :- -.~ '...: ,'9 MD = 3.36 9 MD = 6.24 9 MD = 3.63.e-...........,, ~ .......-.-" :? ,::..,...9 MD = 3.28 9 MD = 2,83 9 MD = 2.54.,..,"E --9 MD = 3.33 9 MD = 3.99 9 MD = 3.33-.; ~&~-41 . --.,.."" :'. ".,. f" """~.-. ):: ..........., ,,~ ..Lower hemisphere 9 MD = 2.71 9 MD = 2.67 9 MD = 2.63Fig. 13. Point <strong>and</strong> contour pole figures <strong>of</strong> plagioclase CPO in the eastern belt (lower hemisphere, equal areaprojections). Diagrams are contom'ed as multiples (0.5, 1.0, 1.5, 2.0, 2.5 .... x ) <strong>of</strong> uniform distribution. MD ismaximum density <strong>of</strong> data in the contour diagrams. Foliation is represented by the horizontal line, lineation istrending E-W.(El) show progressive reorientation <strong>of</strong> the crystallographicplanes (100) towards the foliationdirection <strong>and</strong> c-axes [001] towards the lineation(Fig. t6a). Cleavage planes { 110} cluster parallelto the foliation as well. The asymmetry <strong>of</strong> theCPO fabric is consistent with the dextral shearzone cross-cutting the porphyroclast under theangle <strong>of</strong> ~20 ~ with respect to the mylonitic foliation.The CPO <strong>of</strong> hornblende in the ultramylonite(E3) displays a strong maximum <strong>of</strong> polesto the (100) <strong>and</strong> {110} planes (Fig. 16b) clusteringnormal to the foliation (I--2.9 for (100)).The c-axes trend in the direction <strong>of</strong> lineation.Such CPO patterns suggest that the (100)[001]slip system could be activated. Inclination <strong>of</strong>data with respect to the mylonitic foliation dueto dextral shear is present. Similar fabric asymmetry,interpreted as the result <strong>of</strong> a non-coaxialstrain, has been described in eclogite faciesglaucophanites (Zucali et al. 2002).Hornblende CPO in the augen mylonite (W 1)from the western belt form a cluster <strong>of</strong> (100)planes subparallel to the mylonitic foliation<strong>and</strong> c-axes parallel to the lineation. However, afew grains occur with a different orientation(Fig. 16c). In the b<strong>and</strong>ed mylonite (W2), the(100) planes <strong>and</strong> c-axes reveal similar orientationas in the augen mylonite but the maxima are lesspronounced (Fig. 16d). The I parameters areslightly lower in the b<strong>and</strong>ed mylonite (W2) comparedto the augen mylonite (W1). This CPOpattern is consistent with activation <strong>of</strong> the(100)[001] slip system. No asymmetry withrespect to the shear plane occurs.Discussion <strong>and</strong> conclusionsRecrystallization <strong>and</strong> deformationmechanisms <strong>of</strong> plagioclaseThe early deformation stage in plagioclase in theeastern amphibolite facies metagabbro belt ismarked by undulatory extinction, twinning, <strong>and</strong>fracturing, which are typical for low temperature268


TEXTURES OF NATURALLY DEFORMED METAGABBROS 117Lineationsa) upper hemisphere a-coo) a-(~oo)Protomylonite {.-'..:: "_~.'.~.~E1Plagioclase - eastern belt=219Poles to foliationsa*(100)a*(100)~ M D = 3.91 9D = 3.60, 9a*(100) a'(100) a*(100) a*(100)N=160 N:'.~-...:.Mylonite , .....E2-pl 1 ~!-'".:=o .;MyloniteE2-p12D=3,53 9D=2.839a*(lOO) a*(lO0) a*(lOO) a*(lOO)= 160, ", o-N=I ~ 0oUltramyloniteE3MD = 2,95 9D = 3.20 9a'(~00) a'(10o) a'0 00) a'(100)N = 200D=4, 9 D=4.b) ~ IPF Reference frame 12~~0Directions [uvw] _ Poles to planes (hkl) /o/o21O1 ~' 1~1~10,~~ O l O O l O ~ o i - oFig. 14. (a) Inverse pole figures <strong>of</strong> plagioclase CPO in the eastern belt (upper hemisphere, equal area projections).Projections <strong>of</strong> lineations are presented as point <strong>and</strong> contour diagrams in the left, projections <strong>of</strong> poles to foliations are inthe fight side <strong>of</strong> the figure. N is the number <strong>of</strong> plotted points. Diagrams are contoured as multiples (0.5, 1.0, 1.5, 2.0,2.5 .... • ) <strong>of</strong> uniform distribution. MD is maximum density <strong>of</strong> data in the contour diagrams. Large circle <strong>and</strong> square inthe protomylonite represent two porpbyroclasts. (b) The crystallographic reference frame.<strong>and</strong>/or high differential stress (Tullis & Yund1987, 1992; Rosenberg & Sttinitz 2003). Smallgrains concentrated in these fractures exhibittypical features <strong>of</strong> dynamic recrystallizationsuch as bulging <strong>and</strong> subgrain rotation. However,some <strong>of</strong> them might have developed bycrushing <strong>of</strong> the host grain or by heterogeneousnucleation (StiJnitz et al. 2003; Rosenberg &Sttinitz 2003). Twin-free grains adjacent totwinned host grains (see Fig. 3b), compositionalresemblance <strong>of</strong> recrystallized grains to hostgrains, <strong>and</strong> strong CPO (Lister et al. 1978) mayindicate a recrystallization mechanism by subgrainrotation. This recrystallization mechanismis also supported by small average grain sizes,small aspect ratios <strong>and</strong> weak SPO <strong>of</strong> newgrains. A fairly small degree <strong>of</strong> GBPO for plagioclase-plagioclaseboundaries is in agreementwith subequant shapes <strong>of</strong> equidimensional newgrains, thus favouring the hypothesis <strong>of</strong> subgrain269


118 L. BARATOUX ETAL.Plagioclase - western belt(010) (001)N= 1509 MD = 3.64 9 MD = 3.93 9 MD = 3.169 MD = 5.23 9 MD = 4.39 9 MD = 5.41Lower hemisphere 9 MD = 6.00 9 MD = 4.38 9 MD = 4.77Fig. 15. Point <strong>and</strong> contour pole figures <strong>of</strong> plagioclase CPO in the western belt (lower hemisphere, equal areaprojections). Diagrams are contoured as multiples (0.5, 1.0, 1.5, 2.0, 2.5 .... • ) <strong>of</strong> uniform distribution. MD ismaximum density <strong>of</strong> data in the contour diagrams. Foliation is represented by the horizontal line, lineation is trendingE-W.rotation recrystallization mechanisms. The CPOis strong <strong>and</strong> indicates a possible activity <strong>of</strong>the (001)[110] slip system.The mylonitic stage is characterized by subequantshapes <strong>of</strong> recrystallized plagioclasegrains (low aspect ratios) with straight boundaries,weak shape preferred orientation <strong>and</strong>slightly higher degree <strong>of</strong> GBPO compared tothe protomylonite. Recrystallized grains showincreasing grain sizes compared to protomyloniticstage. Compositional zoning in plagioclaseindicates syndeformational growth <strong>of</strong> matrixgrains (Sodre Borges & White 1980). The plagioclaseCPO in the mylonite is fairly weak. Theseobservations suggest that grain boundary sliding(Boullier & Gu6guen 1975; Lapworth et al.2002), accompanied by some component <strong>of</strong>crystal plastic deformation in plagioclase, wasthe dominant deformation mechanism. Mixing<strong>of</strong> amphibole <strong>and</strong> ptagioclase grains could beanother argument for the presence <strong>of</strong> grainboundary sliding. Rosenberg & Sttinitz (2003)suggested for the syntectonically cooled <strong>and</strong>deformed Bergell tonalite that mixing <strong>of</strong> plagioclase<strong>and</strong> biotite in a fine-grained matrix togetherwith weakening <strong>of</strong> CPO implied a mechanism <strong>of</strong>diffusion-accommodated grain boundary sliding.A switch <strong>of</strong> deformation mechanism from dislocationcreep to grain size sensitive (GSS) flowdue to strain s<strong>of</strong>tening <strong>and</strong> grain size reduction iswell known from quartzo-feldspathic rocks (e.g.Walker et al. 1990; Tullis & Yund 1991). Verysmall grain size is required for activation <strong>of</strong>GSS deformation: 'less than 10 ~m' byBoullier & Gu~guen (1975), 2-16 &m by Tullis& Yund (1991), <strong>and</strong> 3-30 p~m by Stfinitz &Fitz Gerald (1993). However, some studiessuggest grain boundary sliding deformation forgrain sizes <strong>of</strong> 100-150 I-~m for plagioclase(Jensen & Starkey 1985), 24-41 I~m for mixedplagioclase-hornblende layers (Kruse & Stfinitz1999), <strong>and</strong> 100-250 p,m for plagioclase(Lapworth et al. 2002). In our samples withweak or r<strong>and</strong>om CPO, the grain size variesbetween 30 <strong>and</strong> 150 txm. The occurrence <strong>of</strong>grain-boundary diffusion creep for such grainsize may be explained by very low strain rates(Fliervoet et al. 1999; Lapworth et al. 2002).Fine-grained plagioclases from high-straindomains adjacent to amphibole porphyroclasts<strong>and</strong> matrix plagioclase grains in ultramylonitesshow a decrease in average grain size coupledwith increased aspect ratios, strong SPO <strong>and</strong>GBPO. The CPO is strong for the ultramylon_itesample, suggesting activation <strong>of</strong> a (021) 1/2 [ 112]slip system. These features may indicate thatat very high strains the grain size sensitiveflow becomes less important in plagioclase.270


TEXTURES OF NATURALLY DEFORMED METAGABBROS119[a](lOO)Hornblende[c][001]Lower [a]{110} hemisphere (100)[c][001] {110}a) Protomylonite - E 1 b) Ultramyionite - E3- a,~,~ -. .~- | 9 . .." 9N = 209N=989 MD = 7,69 9 MD = 9.57 9 MD = 4.619 MD = 15.98 9 MD = 12.08 9 MD = 5.55c) Augen mylonite - Wl d) B<strong>and</strong>ed mylonite - W2N= 105 N= 107m MD = 7.93 9 MD = 9,94 =MD = 5.11 =MD = 5.15 = MD = 9,46 9 MD = 4.91Fig. 16. Point <strong>and</strong> contour pole figures <strong>of</strong> amphibole CPO in (a <strong>and</strong> b) the eastern belt <strong>and</strong> (c <strong>and</strong> d) the western belt(lower hemisphere, equal area projections). Contours are counted as multiples (1, 2, 3, 4 .... • ) <strong>of</strong> uniform distribution.MD is maximum density <strong>of</strong> data in the contour diagrams. Foliation is represented by the horizontal line, lineationis trending E-W. The CPO <strong>of</strong> porphyroclast is marked by a circle.A possible switch from diffusion-dominated flowin the mylonite to dislocation creep in the ultramylonitemay be due to a decrease in temperature<strong>of</strong> deformation as described from felsic granulitesby Martelat et al. (1999). The switch fromthe GSS to the GSI regime for decreased grainsize is possible <strong>and</strong> implies an increase <strong>of</strong>strain rate <strong>and</strong>/or a decrease <strong>of</strong> temperature inthe shear zone (H<strong>and</strong>y 1990). A similar transitionfrom diffusion-dominated creep in plagioclaseaggregates in augen mylonites to dislocationcreep in b<strong>and</strong>ed ultramylonites was reportedby Schulmann et al. (1996) from deformedmetagranitoids.Dominant (010)[001] slip was identified inplagioclase from both CPO patterns <strong>and</strong>/orTEM observations (e.g. Olsen & Kohlstedt1984, 1985; Montardi & Mainprice 1987; Ji &Mainprice 1988, 1990). However, the observedCPO patterns for amphibolite facies metagabbrosare not compatible with activation <strong>of</strong> the classicslip system. Various other slip systems, whichmay be activated in plagioclase also, weredescribed by Marshall & McLaren (1977a, b),Olsen & Kohlstedt (1984, 1985), <strong>and</strong> Montardi& Mainprice (1987). In addition, St(initz et al.(2003) concluded that during experimentaldeformation <strong>of</strong> plagioclase several slip systems271


120 L. BARATOUX ET AL.are contemporaneously activated. The presence<strong>of</strong> several maxima in inverse pole figures <strong>of</strong>our samples may indicate that several slipsystems probably operated simultaneously <strong>and</strong>their unequivocal identification from CPOpatterns is therefore difficult.In the western augen mylonites, plagioclaserecrystallizes dynamically by subgrain rotation.Aggregates <strong>of</strong> plagioclase matrix grains showmicrostrnctures indicative for dislocation creep,such as subgrain formation, strong SPO, highaspect ratios <strong>and</strong> relatively strong GBPO <strong>of</strong>plagioclase boundaries. Large differences inchemical compositions between porphyroclast<strong>and</strong> matrix may imply that some <strong>of</strong> the grains originatedby heterogeneous nucleation (Rosenberg& Sttinitz 2003). Surprisingly, the CPO <strong>of</strong>recrystallized grains is very weak, which maysuggest that some GSS process took place atthis stage <strong>of</strong> deformation.In the b<strong>and</strong>ed mylonites, the plagioclasematrix grains show strongly migrated boundaries,<strong>and</strong> lower SPO, aspect ratios <strong>and</strong> GBPOcompared to the augen mylonite. Plagioclase ischaracterized by a relatively high average grainsize <strong>and</strong> a common growth-related, chemicallyzoned composition. The plagioclase cores areovergrown by a new plagioclase with a morecalcic composition at upper amphibolite to granulitefacies conditions. A similar mechanism was<strong>numerical</strong>ly modelled by Jessell et al. (2003)showing that some original grain cores aredirectly in contact with rims <strong>of</strong> new grains dueto preferential growth <strong>of</strong> grains at the expense<strong>of</strong> others. This sharp compositional differencemust necessarily introduce a strong chemicalpotential between these two grains. Serratedboundaries indicate a mechanism <strong>of</strong> grain boundarymigration. A process called chemicallyinduced grain boundary migration (CIGM) wasintroduced by Hillert & Purdy (1978) <strong>and</strong> inthe case <strong>of</strong> naturally deformed amphibolesdescribed by Cumbest et al. (1989a). Such aprocess requires a strong chemical potential <strong>and</strong>is driven by a reduction in the free chemicalenergy associated with the chemical change(Yund & Tullis 1991). However, Yund &Tullis (1991) observed important compositionalchanges in experimentally deformed plagioclase,which were not assumed to be the driving forcefor associated dynamic recrystallization. TheCPO in b<strong>and</strong>ed mylonites is relatively strong<strong>and</strong> (001)[100] <strong>and</strong> (010)[100] slip may beinferred from the pole figures. Martelat et al.(1999) reported a strong CPO in plagioclasedeformed by high-temperature diffusion creepin coarse-grained felsic granulites. We propose,however, that a difference in chemical potentialcould have been one <strong>of</strong> the main driving forcesfor grain boundary migration in the b<strong>and</strong>edmylonites.Fine-grained shear zones, cross-cutting thecoarse-grained plagioclase b<strong>and</strong>s, developedlater at somewhat lower temperatures. Thisis supported by the mechanical twins crosscuttingchemically zoned plagioclases fromthese shear zones. Higher aspect ratio, strongerSPO <strong>and</strong> GBPO, <strong>and</strong> strong CPO suggest dominantactivity <strong>of</strong> dislocation creep with activation<strong>of</strong> (001)[100] slip in these domains.Deformation mechanisms <strong>of</strong> amphiboleMicro<strong>structural</strong> evidence for cataclastic flow waspresented by Nyman et al. (1992) from amphibolitesthat were previously interpreted to havebeen deformed plastically. Amphiboles are relativelyresistant to crystal plastic deformation evenat high metamorphic grades (Brodie & Rutter1985). The dominant slip system (100)[001] inamphibole was inferred from both experimental<strong>and</strong> natural studies (Dollinger & Blacic 1975;Rooney et al. 1975; Cumbest et al. 1989b;Hacker & Christie 1990; Skrotzki 1990). The(hk0)[001] slip system was determined byBiermann & Van Roermund (1983) <strong>and</strong>Morrison-Smith (1976). The (010)[100] <strong>and</strong>(001)[100] slip systems were observed byMorrison-Smith (1976) in experimental studiesbut not in nature. All these studies concludedthat the only possible active Burgers vectoroperating in amphiboles WaSoin [001], which isalso the shortest one (5.299 A). [100] <strong>and</strong> [010]Burgers vectors (9.885 A <strong>and</strong> 18.169 A, respectively)were thought to be too long to be activated(Rooney et al. 1970).Hornblende microstructures from the easternmetagabbro belt (protomylonite <strong>and</strong> mylonite)exhibit twinning on (100) planes (Rooney et al.1975), kinking, <strong>and</strong> strong undulatory extinctionrelated to bending/twins <strong>of</strong> the crystalline lattice.Microshear zones transect large porphyroclastsinto bookshelf-like segments <strong>and</strong> these segmentssubsequently rotate with their c-axes parallel tothe shear direction. This is consistent with polefigures showing CPOs <strong>of</strong> small grains that aredifferent from the host porphyroclast. In YZ sections,the grains have lozenge shapes, with theirgrain boundaries parallel to the {110} cleavage.The SPO <strong>and</strong> GBPO are relatively weak forthese strongly elongate grains when comparedto grains from more advanced stages <strong>of</strong> deformation.This is interpreted to be the result <strong>of</strong>development <strong>of</strong> new grains along intracrystallinemicroshear zones <strong>and</strong> along margins <strong>of</strong>porphyroclasts.272


TEXTURES OF NATURALLY DEFORMED METAGABBROS 121Some observations are consistent with brittledeformation <strong>of</strong> amphibole. For instance, hornblendehost grains are <strong>of</strong>ten rimmed by 'subgrains'forming core-mantle-like structures.Grain boundaries between new grains <strong>and</strong> porphyroclastsare very sharp, indicating that newgrains formed by fracturing rather than by subgrainrotation (Nyman et al. 1992). Interstitialplagioclase decorate grain boundaries betweensmall recrystallized grains demonstrating opening<strong>of</strong> spaces between amphiboles. However, theoperating deformation mechanism in hornblendecannot be unequivocally determined fromoptical observations <strong>and</strong> EBSD data alone. Thestrong CPO patterns may be interpreted by(100)[001] slip with dislocation creep being themajor operative deformation mechanism, bytranslation gliding along {110} planes indicatingdominant micr<strong>of</strong>racturing or by mechanicalrotation <strong>of</strong> anisotropic grains. The strong asymmetry<strong>of</strong> CPO reflects the development <strong>of</strong>antithetic microshear zones operating duringbookshelf-like rotations associated with dextralshear.Hornblende crystals in ultramylonites arecharacterized by higher aspect ratios, moreuniform grain sizes, <strong>and</strong> stronger SPO <strong>and</strong>GBPOs compared to amphiboles from protomylonites <strong>and</strong> mylonites. Strong CPO <strong>of</strong> (100)<strong>and</strong> { 110} cleavage planes subparallel to the foliation<strong>and</strong> c-axes subparallel to the lineation areconsistent with activation <strong>of</strong> the (100)[001 ] <strong>and</strong>{110}[001] slip systems. As mentioned above, astrong CPO <strong>of</strong> amphibole may also be producedby rigid body rotation.The most important observation in the westernaugen mylonitic hornblendes is that the originalporphyroclasts <strong>and</strong> new grains have differentchemical compositions (magnesio-hornblende<strong>and</strong> tschermakite with higher Na <strong>and</strong> K content,respectively). Chemical differences between old<strong>and</strong> new grains suggest that new grains originatedby heterogeneous nucleation under highermetamorphic conditions (e.g. Binns 1965; Laird& Albee 1981), which may be chemically <strong>and</strong>/or deformationally induced (Berger & Sttinitz1996). The porphyroclasts show high aspectratios <strong>and</strong> SPO, GBPO, <strong>and</strong> CPO maximaoriented parallel to the mylonitic foliation. Arigid body rotation <strong>of</strong> large porphyroclasts inthe weaker plagioclase matrix is the most likelymechanism to explain the strong preferred orientation<strong>of</strong> these grains. New grains have significantlylower aspect ratios, strong SPO, CPO<strong>and</strong> high degrees <strong>of</strong> GBPO <strong>of</strong> like-like boundaries.The orientation <strong>of</strong> CPO maxima is consistentwith the possible activation <strong>of</strong> (100)[001]slip.In the b<strong>and</strong>ed mylonites, primary hornblendeporphyroclasts are missing; however, slightchemical zoning within large <strong>and</strong> smaller newgrains indicates their syndeformational growth.Amphiboles, in monomineral hornblende b<strong>and</strong>s,show straight boundaries decorated with smallamounts <strong>of</strong> tiny interstitial plagioclase. Amphibolegrains show no signs <strong>of</strong> internal deformation<strong>and</strong> have straight to lobate grain boundaries,which is indicative <strong>of</strong> high degrees <strong>of</strong> texturalequilibration (Brodie & Rutter 1985). The grainsizes <strong>and</strong> the degree <strong>of</strong> SPO <strong>and</strong> GBPO decreasewith respect to less deformed rocks (augen mylonite).However, the CPO is strong <strong>and</strong> verysimilar to that from less deformed augen mylonite,indicating possible activation <strong>of</strong> (100) [001 ]slip. Hornblende grains, which occur in plagioclaseaggregates, are marked by significantlysmaller grain sizes compared to host matrix crystals(Fig. 7c). They are <strong>of</strong>ten located at highenergysurfaces or unstable triple points, <strong>and</strong>show globular shapes typical for a very differentstructure <strong>of</strong> host <strong>and</strong> inclusion grain. In someplaces, where the minor phase becomes abundant,it is forced to occupy less favourableplanar interfaces (Vernon 1976). All these criteriasuggest that these grains developed byheterogeneous nucleation, namely in a highsurface energy plagioclase aggregate (Spry1969; Dallain et al. 1999). The mechanism <strong>of</strong>syndeformational growth is more pronouncedin hornblende layers. This is supported byshapes <strong>of</strong> grain size distribution histograms thatshow important numbers <strong>of</strong> large grains. Possiblemechanisms to explain strong CPO in the studiedrocks is a nucleation <strong>and</strong> growth <strong>of</strong> new hornblendegrains in a non-lithostatic stress field aswas proposed for pyroxenes by Helmstaedtet al. (1972) or Mauler et al. (2001).Implications for the theology <strong>of</strong> atwo-phase amphibole-plagioclase systemExperimental deformation <strong>of</strong> amphibolite <strong>and</strong>granulite facies natural samples <strong>of</strong> metabasites(Wilks & Carter 1990) reveals that the rheology<strong>of</strong> such polyphase systems depends on temperature,rock composition, deformation mechanisms,water content, <strong>and</strong> to a minor extent,pressure. The relationship between these factors<strong>and</strong> differential stress <strong>and</strong> strain is, however,complex <strong>and</strong> Wilks & Carter (1990) concludedthat it is necessary to estimate the contribution<strong>of</strong> each phase to the creep rate <strong>of</strong> the bulk rockin order to establish a convenient flow law.Rosenberg & Sttinitz (2003) proposed thatinterconnected monomineralic networks <strong>of</strong>273


122 L. BARATOUX ET AL.recrystallized plagioclase in middle or deepcrustal metabasic rocks are rare <strong>and</strong> restrictedto rocks with very high plagioclase modal abundancessuch as anorthosites (Ji et al. 1988).However, as mentioned by other authors (e.g.Brodie & Rutter 1985), b<strong>and</strong>ed metabasic mylonitesare common, which is confirmed by thiswork. We show that plagioclase may easilyform interconnected weak layer networks inhornblende gabbros under upper amphibolitefacies conditions.Our micro<strong>structural</strong> study <strong>of</strong> an amphibolitefacies metagabbro (650 _+ 50 ~ shows thatthe load-bearing framework structure (H<strong>and</strong>y1990, 1994) is restricted to the lowest deformationintensities. The non-deformed metagabbroshows coarse-grained ophitic structure composed<strong>of</strong> r<strong>and</strong>omly distributed hornblende <strong>and</strong> plagioclase,where amphibole grains are only locallyin contact. With ongoing strain, the deformationis mostly concentrated in the plagioclase, whichis at the transient region between brittle <strong>and</strong>plastic behaviour, leading to development <strong>of</strong> afine-grained matrix (Tullis & Yund 1987).Amphibole grains showing high internal strain<strong>and</strong> local fracturing behave as rigid bodies surroundedby interconnected layers <strong>of</strong> plagioclasegrains. Such microstructures may be interpretedas an interconnected weak layer structure(IWL) with a high viscosity contrast betweenrigid clasts <strong>of</strong> amphibole <strong>and</strong> weaker, finegrainedplagioclase layers (H<strong>and</strong>y et al. 1999).In the amphibolite facies metagabbroic mylohires,the monomineralic hornblende layers areobserved, while plagioclase-rich layers showalmost perfect mixing with hornblende. Thegrain size <strong>of</strong> plagioclase <strong>and</strong> amphibole in theplagioclase-rich matrix areas is fairly similar,the latter showing slightly higher elongation. Thephase distribution <strong>of</strong> plagioclase-hornblendemixture, the absence <strong>of</strong> CPO in plagioclase,<strong>and</strong> its aspect ratio suggest that the dominantmechanism is granular flow (grain boundarysliding). Based on this interpretation, wesuggest that the phase mixing is probably amechanical process. The microstructures <strong>and</strong>CPO <strong>of</strong> amphibole forming monomineraliclayers indicate either dislocation creep or cataclasticflow. The absence <strong>of</strong> boudinage <strong>and</strong> progressivemixing <strong>of</strong> plagioclase <strong>and</strong> amphibolesuggest that the diffusion-dominated flowprocess operating in plagioclase aggregates ismechanically as efficient as dislocation or cataclasticflow in the hornblende layers. The finalstructure resembles the interconnected weaklayer structure with low viscosity contrast(H<strong>and</strong>y 1994). A switch in deformation mechanismfrom dislocation creep towards a grain sizesensitive process is thought to be responsiblefor the convergence <strong>of</strong> mechanical properties <strong>of</strong>amphibole <strong>and</strong> plagioclase in the mylonite,resulting in a drop <strong>of</strong> bulk rock strength(Etheridge & Wilkie 1979; Kirby 1985; Rutter &Brodie 1988).Two deformation stages were observed in theupper amphibolite facies (750 -t- 50 ~ metagabbros:(1) augen mylonite with locally preservedporphyroclasts <strong>of</strong> both plagioclase <strong>and</strong> hornblende;<strong>and</strong> (2) b<strong>and</strong>ed mylonites with completelyrecrystallized amphibole <strong>and</strong> plagioclase,each arranged in monomineralic layers. Theinitial stages <strong>of</strong> deformation are characterizedby tectonic grain size reduction <strong>of</strong> plagioclasewhile hornblendes represent strong objects floatingin the weak plagioclase matrix. The deformation<strong>of</strong> the metagabbro is interpreted to haveoccurred via dislocation creep accompanied bydiffusion mass transfer mechanisms, responsiblefor moderate mixing <strong>of</strong> plagioclase <strong>and</strong> amphibole.The b<strong>and</strong>ed fabric, only developed at highstrains, can be defined as an interconnectedweak layer structure with low viscosity contrast(H<strong>and</strong>y et al. 1999). The layered structureshows that the strengths <strong>of</strong> amphibole monornineralicaggregates <strong>and</strong> plagioclase-rich b<strong>and</strong>sare similar, suggesting convergence <strong>of</strong> rheologies<strong>of</strong> both minerals at high strains (Jordan1988; H<strong>and</strong>y 1994).In conclusion, amphibolite <strong>and</strong> upper amphibolitefacies metagabbroic mylonites are characterizedby layered low-viscosity IWL structures.This indicates that at high strains this b<strong>and</strong>edstructure in metagabbros forms a so-calledsteady-state foliation (Means 1990). Mechanicalmixing <strong>of</strong> phases is more important in lowertemperature (eastern belt) than in higher temperature(western belt) amphibolite facies mylonites.The bulk strength <strong>of</strong> amphibolite <strong>and</strong>upper amphibolite facies mylonitic metagabbrosis controlled by an equal contribution <strong>of</strong> bothrock-forming minerals showing contrasting butequally efficient deformation mechanisms.We are indebted to D. Mainprice for his help with theEBSD <strong>analyses</strong>. Fruitful discussions with F. Holub, D. J.Prior, H. Stiinitz <strong>and</strong> J. Wheeler are gratefully acknowledged.We thank P. T~)cov~, J. Haloda, P. Gr<strong>and</strong>jean <strong>and</strong>P. Capiez for the help with microprobe <strong>and</strong> bulk rock <strong>analyses</strong>.K. Brodie, H. Van Roermund <strong>and</strong> D. Gapais arethanked for thorough reviews, which improved significantlythe original manuscript. The project was fundedby grants <strong>of</strong> Czech National Grant Agency No. 42-201-204 to K.S. <strong>and</strong> 42-201-318 to P. Stipskfi, by Czech GeologicalService assignment No. 6327 to P. Mixa, <strong>and</strong> by aPhD financial support attributed by the French governmentto L.B.274


TEXTURES OF NATURALLY DEFORMED METAGABBROS 123ReferencesBERGER, A. & STI~INITZ, H. 1996. Deformation mechanisms<strong>and</strong> reaction <strong>of</strong> hornblende: examples fromthe Bergell tonalite (central Alps). Tectonophysics,257, 149-174.BIERMANN, C. & VAN ROERMUND, H. L. M. 1983.Defect structures in naturally deformed clinoamphiboles;a TEM study. Tectonophysics, 95,267-278.BINNS, R. A. 1965. The mineralogy <strong>of</strong> metamorphosedbasic rocks from the Williama Complex, BrokenHill district, New South Wales, Part I, Hornblendes.Mineralogical Magazine, 35, 306-326.BOULLtER, A. M. & GUEGUEN, Y. 1975. SP-mylonites;origin <strong>of</strong> some mylonites by superplastic flow.Contributions to Mineralogy <strong>and</strong> Petrology, 50,93-104.BROD~E, K. H. 1981. Variation in amphibole <strong>and</strong>plagioclase composition with deformation. Tectonophysics,78, 389-402.BRODIE, K. H. & RUTTER, E. H. 1985. On the relationshipbetween deformation <strong>and</strong> metamorphism, withspecial reference to the behavior <strong>of</strong> basic rocks. In:THOMPSON, A. B. & RUBIE, D. C. (eds) MetamorphicReactions; Kinetics, Textures, <strong>and</strong> Deformation.Springer-Verlag, New York, MY, UnitedStates, 138-179.CARVER, N. L. & TSENN, M. C. 1987. Flow properties<strong>of</strong> continental lithosphere. Tectonophysics, 136,27-63.CUMBEST, R. J., DRURY, M. R., VAN ROERMUND,H. L. M. & S1MPSON, C. 1989a. Dynamic recrystallization<strong>and</strong> chemical evolution <strong>of</strong> clinoamphibolefrom Senja, Norway. Contributions to Mineralogy<strong>and</strong> Petrology, 101, 339-349.CUMBEST, R. J., Van ROERMUND, H. L. M., DRURY,M. R. & SIMPSON, C 1989b. Burgers vector determinationin clinoamphibole by computer simulation.American Mineralogist, 74, 586-592.DALLAIN, C., SCHULMANN, K. & LEDRU, P. 1999.Textural evolution in the transition from subsolidusannealing to melting process, Velay Dome, FrenchMassif Central. Journal <strong>of</strong> Metamorphic Geology,17, 61-74.DOLLINGER, G. & BLACIC, J. D. 1975. Deformationmechanisms in experimentally <strong>and</strong> naturally deformedamphiboles. Earth <strong>and</strong> Planetary ScienceLetters, 26, 409-416.ETHERIDGE, M. A. & WILKIE, J. C. 1979. Grainsizereduction, grain boundary sliding <strong>and</strong> theflow strength <strong>of</strong> mylonites. Tectonophysics, 58,159-178.FITZ GERALD, J. D. & STI)NITZ, H. 1993. Deformation<strong>of</strong> granitoids at low metamorphic grade; I, Reactions<strong>and</strong> grain size reduction. Tectonophysics,221, 269-297.FLIERVOET, T. F., DRURY, M. R. & CHOPRA, P. N.1999. Crystallographic preferred orientations <strong>and</strong>misorientations in some olivine rocks deformedby diffusion or dislocation creep. Tectonophysics,303, 1-27.FREER, R. 1981. Diffusion in silicate minerals <strong>and</strong>glasses: a data digest <strong>and</strong> guide to the literature.Contributions to Mineralogy <strong>and</strong> Petrology, 76,440-454.GAPAIS, D. 1989. Shear structures within deformedgranites; mechanical <strong>and</strong> thermal indicators.Geology, 17, 1144-1147.HACKER, B. R. & CHRISTIE, J. M. 1990. Brittle/ductile<strong>and</strong> plastic/cataclastic transitions in experimentallydeformed <strong>and</strong> metamorphosed amphibolite.In: DUBA, A. G., DURHAM, W. B., HANDIN, J. W.& WANG HERBERT, F. (eds) The Brittle-DuctileTransition in Rocks. American GeophysicalUnion, Washington DC, Geophysical Monograph,56, 127-147.HANDY, M. R. 1990. The solid-state flow <strong>of</strong> polymineralicrocks. Journal <strong>of</strong> Geophysical Research,B, 95, 8647 - 8661.HANDY, M. R. 1994. Flow laws for rocks containingtwo non-linear viscous phases; a phenomenologicalapproach. Journal <strong>of</strong> Structural Geology, 16,287-301.HANDY, M. R_, WISSING, S. B. & STREIT, L. E. 1999.Frictional-viscous flow in mylonite with variedbimineralic composition <strong>and</strong> its effect on lithosphericstrength. Tectonophysics, 303, 175-191.HELMSTAEDT, H., ANDERSON, O. L. & GAVASCI, A. T.1972. Petr<strong>of</strong>abric studies <strong>of</strong> eclogite, Spinel-Websterite, <strong>and</strong> Spinel-Lherzolite xenoliths fromkimberlite-bearing breccia pipes in SoutheasternUtah <strong>and</strong> Northeastern Arizona. Journal <strong>of</strong>Geophysical Research, 77, 4350-4365.HILLERT, M. & PURDY, G. R. 1978. Chemicallyinduced grain boundary migration. Acta Metallica,26, 333-340.HOLLAND, T. J. B. & BLUNDY, J. 1994. Non-idealinteractions in calcic amphiboles <strong>and</strong> theirbearing on amphibole-plagioclase thermometry.Contributions to Mineralogy <strong>and</strong> Petrology, 116,433 -447.JENSEN, L. N. & STARKLY, J. 1985. Plagioclase micr<strong>of</strong>abricsin a ductile shear zone from the JotunNappe, Norway. Journal <strong>of</strong> Structural Geology,7, 527-539.J~SSELL, M. W., KOSTEN~:O, O. & JAMTVEIT, B. 2003.The preservation potential <strong>of</strong> microstructuresduring static grain growth. Journal <strong>of</strong> MetamorphicGeology, 21, 481-491.JI, S. & MAINPRICE, D. 1988. Natural deformationfabrics <strong>of</strong> plagioclase; implications for slip systems<strong>and</strong> seismic anisotropy. Tectonophysics,147, 145-163.JI, S. & MAINPRICE, D. 1990. Recrystallization <strong>and</strong>fabric development in plagioclase. Journal <strong>of</strong>Geology, 98, 65-79.JI, S., MAINPRICE, D. & BOUDIER, F. 1988. Sense <strong>of</strong>shear in high-temperature movement zones fromthe fabric asymmetry <strong>of</strong> plagioclase feldspars.Journal <strong>of</strong> Structural Geology, 10, 73-81.JORDAN, P. G. 1988. The rheology <strong>of</strong> polymineralicrocks - an approach. Geologische Rundschau,77, 285-294.KIRBY, S. H. 1985. Rock mechanics observations pertinentto the theology <strong>of</strong> the continental lithosphere<strong>and</strong> the localization <strong>of</strong> strain along shear zones.Tectonophysics, 119, 1-27.275


124 L. BARATOUX ETAL.KNIPE, R. J. 1989. Deformation mechanisms, recognitionfrom natural tectonites. Journal <strong>of</strong> StructuralGeology, 11, 127-146.KNIPE, R. J. & MCCAIG, A. M. 1994. Micro<strong>structural</strong><strong>and</strong> microchemical consequences <strong>of</strong> fluid flow indeforming rocks. In: PARNELL, J. (ed) Ge<strong>of</strong>luids;Origin, Migration <strong>and</strong> Evolution <strong>of</strong> Fluids in SedimentaryBasins. Geological Society, London,Special Publications, 78, 99-111.KRETZ, R. 1983. Symbols for rock-forming minerals.American Mineralogist, 68, 277-279.KRONER, A., STIPSKA, P., SCHULMANN, K. &JAECKEL, P. 2000. Chronological constraints onthe pre-Variscan evolution <strong>of</strong> the northeasternmargin <strong>of</strong> the Bohemian Massif, Czech Republic.In" FRANKE, W., HAAK, V., ONCKEN, O. &TANNER, D. (eds) Orogenic Processes; Quantification<strong>and</strong> Modelling in the Variscan Belt.Geological Society, London, Special Publications,179, 175-197.KRUHL, J. H. 1987. Preferred lattice orientation <strong>of</strong>plagioclase from amphibolite <strong>and</strong> greenschistfacies rocks near the Insubric line (WesternAlps). Tectonophysics, 135, 233-242.KRUSE, R. & STf0NITZ, H. 1999. Deformation mechanisms<strong>and</strong> phase distribution in mafic hightemperaturemylonites from the Jotun Nappe,southern Norway. Tectonophysics, 303, 223-249.KRUSE, R., STONITZ, H. & KUNZE, K. 2001. Dynamicrecrystallization processes in plagioclase porphyroclasts.Journal <strong>of</strong> Structural Geology, 23,1781-1802.LAFRANCE, B. & VERNON, R. H. 1993. Mass transfer<strong>and</strong> micr<strong>of</strong>racturing in gabbroic mylonites <strong>of</strong>the Guadalupe igneous complex, California. In:BOLAND, J. N. & FITZ GERALD, J. D. (eds)Defects <strong>and</strong> Processes in the Solid State;Geoscience Applications; The McLaren Volume.Developments in Petrology, 14, 151-167.LAIRD, J. & ALBEE, A. L. 1981. Pressure, temperature,<strong>and</strong> time indicators in mafic schists: their applicationto reconstructing the polymetamorphichistory <strong>of</strong> Vermont. American Journal <strong>of</strong> Science,281, 127-175.LAPWORTH, T., WHEELER, J. & PRIOR, D. J. 2002.The deformation <strong>of</strong> plagioclase investigated usingelectron backscatter diffraction crystallographicpreferred orientation data. Journal <strong>of</strong> StructuralGeology, 24, 387-399.LEAKE, B. E., WOOLLEY, A. R. et aI. 1997. Nomenclature<strong>of</strong> amphiboles; report <strong>of</strong> the subcommittee onamphiboles <strong>of</strong> the International MineralogicalAssociation, Commission on New Minerals <strong>and</strong>Mineral Names. The Canadian Mineralogist, 35,219-246.LEXA, O. 2003. Numerical approach in <strong>structural</strong><strong>and</strong> micro<strong>structural</strong> <strong>analyses</strong>. PhD thesis, CharlesUniversity, Prague.LISLE, R. J. 1985. The use <strong>of</strong> the orientation tensor forthe description <strong>of</strong> statistical fabrics. Journal <strong>of</strong>Structural Geology, 7, 115-117.LISTER, G. S., PATERSON, M. S. & HOBBS, B. E. 1978.The simulation <strong>of</strong> fabric development in plasticdeformation <strong>and</strong> its application to quartzite; themodel. Tectonophysics, 45, 107-158.MARSHALL, D. B. & MCLAREN, A. C. 1977a. Deformationmechanisms in experimentally deformedplagioclase feldspars. Physics <strong>and</strong> Chemistry <strong>of</strong>Minerals, 1,351-370.MARSHALL, D. B. & MCLAREN, A. C. 1977b. Thedirect observation <strong>and</strong> analysis <strong>of</strong> dislocations inexperimentally deformed plagioclase feldspars.Journal <strong>of</strong> Material Science, 12, 893-903.MARTELAT, J. E., SCHULMANN, K., LARDEAUX, J. M.,NICOLLET, C. & CARDON, H. 1999. Granulite micr<strong>of</strong>abrics<strong>and</strong> deformation mechanisms in southernMadagascar. Journal <strong>of</strong> Structural Geology, 21,671-687.MAULER, A., GODARD, G. & KUNZE, K. 2001. Crystallographicfabrics <strong>of</strong> omphacite, rutile <strong>and</strong> quartzin Vendee eclogites (Armorican Massif, France);consequences for deformation mechanisms <strong>and</strong>regimes. Tectonophysics, 342, 81 - 112.McCAI6, A. M. & KNIPE, R. J. 1990. Mass-transportmechanisms in deforming rocks; recognitionusing micro<strong>structural</strong> <strong>and</strong> microchemical criteria.Geology, 18, 824-827.MEANS, W. D. 1990. Kinematics, stress, deformation<strong>and</strong> material behavior. Journal <strong>of</strong> StructuralGeology, 12, 953-971.MONTARDI, Y. & MAINPRICE, D. 1987. A transmissionelectron microscopy study <strong>of</strong> the natural plasticdeformation <strong>of</strong> calcic plagioclases (An 68-70).Bulletin de Mineralogie, 110, 1-14.MORRISON-SMITH, D. J. 1976. Transmission electronmicroscopy <strong>of</strong> experimentally deformed hornblende.American Mineralogist, 61, 272-280.NYMAN, M. W., LAW, R. D. & SMELIK, E. A. 1992.Cataclastic deformation mechanism for the development<strong>of</strong> core-mantle structures in amphibole.Geology, 20, 455-458.OLSEN, T. S. & KOHLSTEDT, D. L. 1984. Analysis <strong>of</strong>dislocations in some naturally deformed plagioclasefeldspars. Physics <strong>and</strong> Chemistry <strong>of</strong> Minerals,11, 153-160.OLSEN, T. S. & KOHLSTEDT, D. L. 1985. Naturaldeformation <strong>and</strong> recrystallization <strong>of</strong> some intermediateplagioclase feldspars. Tectonophysics,111, 107-131.ORD, A. & HOBBS, B. E. 1989. The strength <strong>of</strong> the continentalcrust, detachment zones <strong>and</strong> the development<strong>of</strong> plastic instabilities. Tectonophysics, 158,269-289.PANOZZO, R. H. 1983. Two-dimensional analysis <strong>of</strong>shape-fabric using projections <strong>of</strong> digitized lines ina plane. Tectonophysics, 95, 279-294.POIRIER, J. P. & GUILLOPt~, M. 1979. Deformationinduced recrystallization <strong>of</strong> minerals. Bulletin deMindralogie, 102, 67-74.PRIOR, D. J. & WHEELER, J. 1999. Feldspar fabrics in agreenschist facies albite-rich mylonite from electronbackscatter diffraction. Tectonophysics, 303,29 -49.PRIOR, D. J., BOYLE, A. P. et al. 1999. The application<strong>of</strong> electron backscatter diffraction <strong>and</strong> orientationcontrast imaging in the SEM to textural problemsin rocks. American Mineralogist, 84, 1741-1759.276


TEXTURES OF NATURALLY DEFORMED METAGABBROS 125ROONEY, T. P., RIECKER, R. E. & Ross, M. 1970.Deformation twins in hornblende. Science, 169,173-175.ROONEY, T. P., RIECKER, R. E. & GAVASCI, A. T.1975. Hornblende deformation features. Geology,3, 364-366.ROSENBERG, C. L. & STONITZ, H. 2003. Deformation<strong>and</strong> recrystallization <strong>of</strong> plagioclase along atemperature gradient: An example from theBergell tonalite. Journal <strong>of</strong> Structural Geology,25, 389-408.RUTTER, E. H. & BRODIE, K. H. 1988. Experimental'syntectonic' dehydration <strong>of</strong> serpentinite underconditions <strong>of</strong> controlled pore water pressure.Journal <strong>of</strong> Geophysical Research, B, 93, 4907-4932.RUTTER, E. H. 8,: BRODIE, K. H. 1992. Rheology <strong>of</strong> thelower crust. In: FOUNTAIN, D. M., ARCULUS, R. &KAY, R. W. (eds) Continental Lower Crust. Developmentsin Geotectonics, Elsevier, Amsterdam, 23,201-267.SCHULMANN, K., ML(~OCH, B. 8,: MELKA, R. 1996.High-temperature microstructures <strong>and</strong> rheology <strong>of</strong>deformed granite, Erzgebirge, Bohemian Massif.Journal <strong>of</strong> Structural Geology, 18, 719-733.SODRE-BORGES, P. & WHITE, S. H. 1980. Micro<strong>structural</strong><strong>and</strong> chemical studies <strong>of</strong> sheared anorthasites,Roneval, South Harris. Journal <strong>of</strong> StructuralGeology, 2, 273-280.SIMPSON, C. 1985. Deformation <strong>of</strong> granitic rocksacross the brittle ductile transition. Journal <strong>of</strong>Structural Geology, 7, 503-511.SKROTZKI, W. 1990. Microstructure in hornblende <strong>of</strong> amylonitic amphibolite. In: Knipe ROBERT, J. &RUTTER, E. H. (eds) Deformation Mechanisms,Rheology <strong>and</strong> Tectonics. Geological Society,London, Special Publications, 54, 321-325.SPRY, A. 1969. Metamorphic Textures. PergamonPress, Oxford.STiPSKA, P., SCHULMANN, K., THOMPSON, A. B.,JE~EK, J. & KRONER, A. 2001. Thermo-mechanicalrole <strong>of</strong> a Cambro-Ordovician paleorift during theVariscan collision; the NE margin <strong>of</strong> the BohemianMassif. Tectonophysics, 332, 239-253.STfPSKA, P., SCHULMANN, K. & KRONER, A. 2004.Vertical extrusion <strong>and</strong> middle crustal spreading <strong>of</strong>omphacite granulite: a model <strong>of</strong> synconvergentexhumation (Bohemian Massif, Czech Republic).Journal <strong>of</strong> Metamorphic Geology, 22, 179-198.ST15NITZ, H. & F/TZ GERALD, J. D. 1993. Deformation<strong>of</strong> granitoids at low metamorphic grade; II, Granularflow in albite-rich mylonites. Tectonophysics,221, 299-324.STUNITZ, H. FITZ GERALD, J. D. & TULLIS, J. 2003.Dislocation generation, slip systems, <strong>and</strong> dynamicrecrystallization in experimentally deformedplagioclase single crystals. Tectonophysics, 372,215-233.SUN, S. S. & MCDONOUGH, W. F. 1989. Chemical <strong>and</strong>isotopic systematics <strong>of</strong> oceanic basalts; implicationsfor mantle composition <strong>and</strong> processes. In:SAUNDERS, A. D. & NORRY, M. J. (eds) Magmatismin the Ocean Basins. Geological Society,London, Special Publications, 42, 313-345.TULLIS, J. 1983. Deformation <strong>of</strong> feldspars. In:RmBE, P. H. (ed) Feldspar Mineralogy. Reviewsin Mineralogy. Mineralogical Society <strong>of</strong> America,Washington, DC, 2, 297-323.TULLIS, J. & YUND, R. A. 1987. Transition fromcataclastic flow to dislocation creep <strong>of</strong> feldspar;mechanisms <strong>and</strong> microstructures. Geology, 15,606-609.TULLIS, J. & YUND, R. A. 1991. Diffusion creepin feldspar aggregates; experimental evidence.Journal <strong>of</strong> Structural Geology, 13, 987-1000.TULLIS, J. & YUND, R. 1992. The brittle-ductile transitionin feldspar aggregates; an experimentalstudy. In: EVANS, B. & WONG TENG, F. (eds)Fault Mechanics <strong>and</strong> Transport Properties <strong>of</strong>Rocks; A Festschrift in Honor <strong>of</strong> W. F. Brace. AcademicPress, San Diego, United States, 89-117.VERNON, R. H. 1976. Metamorphic Processes. Murby,London, Wiley, New York.VOLL, G. 1976. Recrystallization <strong>of</strong> quartz, biotite <strong>and</strong>feldspars from Erstfeld to the Leventina Nappe,Swiss Alps, <strong>and</strong> its geological significance. SchweizerischeMineralogische und PetrographischeMitteilungen, 56, 641-647.WALKER, A. N., RUTTER, E. H. & BRODIE, K. H. 1990.Experimental study <strong>of</strong> grain-size sensitive flow <strong>of</strong>synthetic, hot-pressed calcite rocks. In: KNIPE, R. J.& RUTTER, E. H. (eds) Deformation Mechanisms,Rheology <strong>and</strong> Tectonics. Geological Society,London, Special Publications, 54, 259-284.WILKS, K. R. • CARTER, N. L. 1990. Rheology <strong>of</strong>some continental lower crustal rocks. Tectonophysics,182, 57-77.YUND, R. & TULLIS, J. 1991. Compositional changes<strong>of</strong> minerals associated with dynamic recrystallization.Contributions to Mineralogy <strong>and</strong> Petrology,108, 346-355.ZUCALI, M., CHATEIGNER, D., DUGNANI, M.,LETTEROTTI, L. & OULADDIAF, B. 2002. <strong>Quantitative</strong>texture analysis <strong>of</strong> glaucophanite deformedunder eclogite facies conditions (Sesia-Lanzozone, Western Alps): comparison between X-ray<strong>and</strong> neutron diffraction analysis. In: DE MEER, S.,DRURY, M. R., DE BRESSER, J. H. P. &PENNOCK, G. M. (eds) Deformation Mechanisms,Rheology <strong>and</strong> Tectonics: Current Status <strong>and</strong>Future Perspectives. Geological Society, London,Special Publications, 200, 219-238.277


ClickHereforFullArticleJOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, B10210, doi:10.1029/2006JB004820, 2007Extreme ductility <strong>of</strong> feldspar aggregates—Melt-enhanced grainboundary sliding <strong>and</strong> creep failure: Rheological implications forfelsic lower crustProkop Závada, 1,4 Karel Schulmann, 2 Jiří Konopásek, 3 Stanislav Ulrich, 1,4<strong>and</strong> Ondrej Lexa 1,2Received 25 October 2006; revised 17 July 2007; accepted 13 August 2007; published 27 October 2007.[1] High-grade orthogneisses from granulite-bearing lower crustal unit show extremefinite strains <strong>of</strong> both K-feldspar <strong>and</strong> plagioclase with respect to weakly deformedquartz aggregates. K-feldspar aggregate in the most intensely deformed sampleshows interstitial grains <strong>of</strong> quartz <strong>and</strong> albite, which also mark some intragranular fractureswithin K-feldspar grains. Both interstitial grains <strong>and</strong> fractures are oriented mostlyperpendicular to the sample stretching lineation. Quartz <strong>and</strong> albite grains withinK-feldspar b<strong>and</strong>s are interpreted as crystallized from interstitial melt <strong>and</strong> the petrologystudy shows that the melt was produced by a metamorphic reaction inplagioclase-mica b<strong>and</strong>s. Thermodynamic Perple_X modeling shows that melt volumeincrease was negligible <strong>and</strong> melt amount was too small to generate considerablemelt overpressure for calculated PT conditions. It is therefore suggested that dilation <strong>of</strong>K-feldspar aggregates <strong>and</strong> fracturing <strong>of</strong> its grains represent a final creep failure state,which resulted from the cavitation process accompanying grain boundary slidingcontrolled diffusion creep. The consequence <strong>of</strong> cavitation-driven dilation <strong>of</strong> K-feldsparaggregates is the local underpressure resulting in infiltration <strong>of</strong> melt from plagioclaseb<strong>and</strong>s. Analogy with metallurgy experiments shows that the cavitation process,exclusively developed in cryptoperthitic K-feldspar, can be attributed to its lower puritycompared to more pure plagioclase. Contrasting rheological behavior <strong>of</strong> feldsparswith respect to quartz prior to fracturing is attributed to different deformationmechanisms. Feldspars appear weaker due to grain boundary sliding accommodated bycoupled melt-enhanced diffusion creep along grain boundaries <strong>and</strong> dislocation creepwithin grains, in contrast to quartz deforming via grain boundary migrationaccommodated dislocation creep.Citation: Závada, P., K. Schulmann, J. Konopásek, S. Ulrich, <strong>and</strong> O. Lexa (2007), Extreme ductility <strong>of</strong> feldspar aggregates—Meltenhancedgrain boundary sliding <strong>and</strong> creep failure: Rheological implications for felsic lower crust, J. Geophys. Res., 112, B10210,doi:10.1029/2006JB004820.1. Introduction[2] It has been experimentally demonstrated that smallamount <strong>of</strong> interstitial melt increases creep rates <strong>of</strong> deformingrocks <strong>and</strong> can induce switch <strong>of</strong> deformation mechanisms[Cooper <strong>and</strong> Kohlstedt, 1984; Dell’Angelo et al., 1987;Dell’Angelo <strong>and</strong> Tullis, 1988]. Dell’Angelo et al. [1987]have shown transition from dislocation creep to meltenhanceddiffusion creep in fine-grained granitic aggregates1 Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University,Prague, Czech Republic.2 Centre de Geochimie de la Surface, UMR 7516, Université LouisPasteur, Strasbourg, France.3 Czech Geological Survey, Prague, Czech Republic.4 Geophysical Institute, Czech Academy <strong>of</strong> Sciences, Prague, CzechRepublic.Copyright 2007 by the American Geophysical Union.0148-0227/07/2006JB004820$09.00at small volume fractions <strong>of</strong> melt F = 0.01–0.03 contemporaneouslywith strength drop below the limit <strong>of</strong> detection.Rosenberg <strong>and</strong> H<strong>and</strong>y [2005] have shown that the ‘‘meltconnectivity transition’’ marked by melt fraction F = 0.07 iscritical in mineral aggregate strength drop at experimentalconditions. These results imply that small amount <strong>of</strong> meltcan be responsible for considerable weakening <strong>of</strong> crustalrocks, which can explain, e.g., the development <strong>of</strong> largescaleshear zones bounding regions <strong>of</strong> rapid uplift [Hollister<strong>and</strong> Crawford, 1986].[3] Melt bearing microstructures in deformed rocks canbe also used for reconstruction <strong>of</strong> grain-scale melt migrationpathways, because deformation <strong>and</strong> melt extraction arecoupled [Brown <strong>and</strong> Rushmer, 1997; Rosenberg <strong>and</strong> H<strong>and</strong>y,2000; Rosenberg, 2001]. A typical feature <strong>of</strong> deformationexperiments at low melt volumes is the preferential distribution<strong>of</strong> melt along grain boundaries oriented at low angleto principal compressive stress direction [Dell’Angelo et al.,B102102791<strong>of</strong>15


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 1. Simplified geological map <strong>of</strong> the studied area around the Eger river valley, NE <strong>of</strong> Stráž nadOhří town [Hradecký, 2002] (scale 1:25000) <strong>and</strong> the position <strong>of</strong> the studied area in the Bohemian Massif[Franke, 2000] <strong>and</strong> European Variscides. AM, Armorican massif; MC, Massif Central; BM, Bohemianmassif.1987; Daines <strong>and</strong> Kohlstedt, 1997; Gleason et al., 1999;Rosenberg <strong>and</strong> H<strong>and</strong>y, 2000]. This phenomenon can beexplained either by dilation <strong>of</strong> granular aggregates due tomelt overpressure [Renner et al., 2000], when hydrostaticmelt pressure overcomes the confining pressure, or bypassive melt migration into the incipient voids during grainboundary sliding (GBS) [Čadek, 1988].[4] However, studies <strong>of</strong> natural microstructures showingevidence <strong>of</strong> large finite strains <strong>and</strong> interstitial quenchedmelt, which would be indicative for pronounced meltenhancedweakening <strong>and</strong> grain-scale melt migration arepractically absent. This is due to two reasons: (1) Naturalmelt-bearing microstructures equilibrated at high-grade conditionsare likely to be overprinted by subsequent static ordynamic recrystallization [Rosenberg <strong>and</strong> Riller, 2000;Rosenberg, 2001]. (2) High finite strains, slow creep rates<strong>and</strong> high homologous temperatures necessary to introducehigh-temperature melt-enhanced deformation mechanisms(e.g., diffusion creep or dislocation climb accommodatedGBS [Boullier <strong>and</strong> Gueguen, 1975]) cannot be attainedusing conventional experimental facilities. There are threeapproaches that may help to underst<strong>and</strong> the rheologicalbehavior <strong>of</strong> minerals deformed at high-grade conditions inthe presence <strong>of</strong> melt <strong>and</strong> their influence on rheology <strong>of</strong>polyphase rocks: (1) Investigation <strong>of</strong> high-grade rocks,where micro<strong>structural</strong> record was preserved due to rapidcooling rates, (2) comparison <strong>of</strong> high-grade rock microstructureswith results <strong>of</strong> various experiments on creepproperties <strong>of</strong> metals, alloys <strong>and</strong> steels, <strong>and</strong> (3) application<strong>of</strong> more efficient rock deformation laboratory facilities (e.g.,Paterson rig apparatus).[5] In this work, we characterize deformation microstructuresin a strongly deformed high-grade b<strong>and</strong>ed orthogneiss,where melting was concurrent with deformation. Meltbearing microstructures have been preserved due to rapidcooling that effectively increased quenching <strong>of</strong> melt instudied rocks [Zulauf et al., 2002]. Deformation mechanisms<strong>of</strong> constituent mineral phases are investigated usingquantitative micro<strong>structural</strong> analysis. The mechanism responsiblefor preferred distribution <strong>of</strong> interstitial melt insample with maximum strain intensity is critically evaluatedusing modeling <strong>of</strong> melt volume produced by a metamorphicreaction <strong>and</strong> molar volume change <strong>of</strong> the system. The effect<strong>of</strong> mobile melt phase on deformation mechanisms <strong>and</strong> bulkrock rheology is discussed. Finally, we consider the apparentstrength relationship between quartz <strong>and</strong> feldspars.2. Geological Setting[6] The rocks studied belong to a lower crustal hightemperaturemetamorphic unit that was rapidly exhumedalong the major Variscan tectonic boundary in the BohemianMassif called the North Bohemian Shear Zone (NBSZ <strong>of</strong>Zulauf et al. [2002]). The trend <strong>and</strong> spatial limits <strong>of</strong> theNBSZ are not defined in the field, because it is mostlycovered by Neogene volcanics <strong>and</strong> sedimentary sequences(Figure 1). The sampled unit forms the uppermost thrustsheet within a crustal nappe stack that displays typicalmetamorphic inversion <strong>and</strong> is called the ‘‘Upper CrystallineNappe’’ (UCN) [Konopásek <strong>and</strong> Schulmann, 2005]. TheUCN in the studied area consists <strong>of</strong> b<strong>and</strong>ed orthogneisses,high-pressure granulites <strong>and</strong> migmatites (Figure 1). PTcalculations <strong>of</strong> metamorphic conditions for the partiallymolten orthogneisses in this study reveal pressures <strong>of</strong>900 MPa <strong>and</strong> temperatures 700 ± 20°C. In contrast,the granulites exhibit peak metamorphic conditions <strong>of</strong>1600 MPa <strong>and</strong> 800°C [Kotková etal., 1996] indicating thatthe granulites <strong>and</strong> orthogneisses have been juxtaposed atmiddle crustal levels during differential exhumation. Furthermore,very fast cooling rates <strong>of</strong> 50°C +25°C/ 17°CMa 1 were deduced [Zulauf et al., 2002] from the U-Pbdating <strong>of</strong> formation <strong>of</strong> zircons <strong>and</strong> monazites (342 ± 1 Ma)<strong>and</strong> 40 Ar- 39 Ar cooling ages <strong>of</strong> muscovites <strong>and</strong> biotites(341 ± 4 Ma) from the orthogneisses.2<strong>of</strong>15280


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 2. (a) Traced <strong>and</strong> digitized mineral phase boundaries in XZ <strong>and</strong> YZ rock sections <strong>of</strong> themylonite type 3 sample (B109) <strong>and</strong> XZ section <strong>of</strong> a metagranite type 1 sample (B108) for comparison(mineral abbreviations after Kretz [1983], Mat means matrix <strong>and</strong> represents indiscernible mineral phases).(b) Flinn diagram <strong>of</strong> the phase aggregate’s shapes. R xy <strong>and</strong> R yz designate X/Y <strong>and</strong> Y/Z ratios <strong>of</strong> the strainellipsoid axes, respectively [Flinn, 1962].[7] Intensive north-south Variscan compression resultedin the development <strong>of</strong> vertical foliations in both the UCN<strong>and</strong> adjacent middle crustal rocks (LCN, Lower CrystallineNappe) [Konopásek et al., 2001] <strong>and</strong> also vertical stretchinglineations in the UCN. Zulauf et al. [2002] proposed thatthis vertical fabric is related to the rapid vertical motion <strong>of</strong>lower crustal rocks along the NBSZ <strong>and</strong> to intense deformation<strong>of</strong> gneisses <strong>and</strong> granulites. The UCN orthogneissshows significantly higher finite strains <strong>of</strong> feldspars comparedto relatively weakly deformed quartz suggestingextreme ductility <strong>of</strong> feldspars. In this work we focus onthe evaluation <strong>of</strong> the competency relationship betweenfeldspars <strong>and</strong> quartz, which can be <strong>of</strong> great importance forthe rheology <strong>of</strong> felsic lower crust.3. Finite Strain Analysis[8] In order to evaluate the influence <strong>of</strong> individual mineralson the rheology <strong>of</strong> UCN orthogneiss, the specimenshave been sampled according to the degree <strong>of</strong> finite strain.Plates cut along the XZ <strong>and</strong> YZ planes <strong>of</strong> the rock fabricellipsoid (XZ, parallel to lineation <strong>and</strong> perpendicular t<strong>of</strong>oliation; YZ, perpendicular to lineation <strong>and</strong> foliation) werecolored [Gabriel <strong>and</strong> Cox, 1929] in order to distinguish K-feldspar from plagioclase aggregates. The boundaries <strong>of</strong>mineral phases in both XZ <strong>and</strong> YZ sections were thenmanually traced <strong>and</strong> digitized from photographs in ArcViewGIS environment (Figure 2a). The phase aggregate objectswere then statistically evaluated using the PolyLX Matlabtoolbox [Lexa et al., 2005] (http://petrol.natur.cuni.cz/ondro/polylx:home). Because both feldspars in b<strong>and</strong>edorthogneiss samples formed apparent infinite prolate-shapedaggregates, their strain was calculated using a method basedon normalization <strong>of</strong> their volume by the mean volume <strong>of</strong> thephase aggregates in sample with low finite strain intensity(where the phase objects are closed, sample B108). It isassumed that initially nonb<strong>and</strong>ed feldspar aggregates coalescedtogether during the deformation. Quartz aggregatesin the XZ section appeared to retain similar cross-sectionalarea as in the sample with lowest strain intensity. Wetherefore suggest that quartz aggregates did not undergoextensive pinching by other phases during deformation,which would decrease their cross-sectional area <strong>and</strong> finitestrain estimate. Bulk deformation was calculated from allphases together as object area weighted geometrical mean.[9] Results <strong>of</strong> shape analysis are demonstrated in theFlinn diagram [Flinn, 1962] (Figure 2b). Sample B108shows the lowest degree <strong>of</strong> deformation <strong>and</strong> represents a‘‘metagranite’’ type 1. The ‘‘b<strong>and</strong>ed orthogneisses’’ type 2(samples B13, B109-2) show rod-shaped aggregates <strong>of</strong> bothfeldspars <strong>and</strong> plane strain to slightly prolate symmetry <strong>of</strong>quartz. This type shows bulk intensities around D = 5 (D =(R 2 xy +R 2 yz ) 1/2 [Ramsay <strong>and</strong> Huber, 1983], D <strong>of</strong> K-feldsparclose to 10 <strong>and</strong> <strong>of</strong> plagioclase around 6. Maximum bulkintensity D = 16 was attained in sample B109 with K-feldspar <strong>and</strong> plagioclase reaching D = 26 <strong>and</strong> D = 50,respectively. This sample represents the final member <strong>of</strong>micro<strong>structural</strong> evolution <strong>and</strong> will be designated the‘‘mylonite’’ type 3 sample. All stages <strong>of</strong> deformation showhigher finite strains in K-feldspar <strong>and</strong> plagioclase aggregatescompared to quartz.4. Deformation Microstructures[10] Microstructures <strong>and</strong> mineral textures <strong>of</strong> all rock‘‘types’’ were studied in detail in order to evaluate theevolution <strong>of</strong> deformation mechanisms with respect to increasingbulk rock strain. <strong>Quantitative</strong> micro<strong>structural</strong> analysiswas carried out using manually digitized grainboundaries [see Lexa et al., 2005] from several sequentialmicrophotographs <strong>of</strong> colored XZ thin sections. Texturalanalysis comprises quantification <strong>of</strong> the grain size <strong>and</strong> grainsize spread, grain shape (axial ratio) <strong>and</strong> shape preferredorientation (SPO), expressed as the ratio <strong>of</strong> eigenvalues <strong>of</strong>3<strong>of</strong>15281


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Table 1. Micro<strong>structural</strong> Analysis Data From Digitized ThinSections a Ferret, mm Axial RatioKfs Pl Qtz Kfs Pl QtzB108 0.32 0.25 0.47 1.40 1.44 1.51B13 0.30 0.26 0.42 1.48 1.42 1.49B109-2 0.26 0.17 0.47 1.42 1.45 1.65B109 0.12 0.09 0.90 1.43 1.49 1.58Q 3 –Q 1 , mmSPOKfs Pl Qtz Kfs Pl QtzB108 0.21 0.16 0.35 1.20 1.18 1.14B13 0.22 0.18 0.23 1.30 1.21 1.08B109-2 0.20 0.15 0.25 1.17 1.07 1.12B109 0.07 0.05 0.93 1.18 1.13 1.45a The grain size represented by the Ferret diameter is the diameter <strong>of</strong>circles with the same area as the measured grain. SPO is the eigenvalueratio <strong>of</strong> the shape preferred orientation tensor. Q 3 –Q 1 designates grain sizespread quantified as the difference between third <strong>and</strong> first quartile.the shape preferred orientation tensor [Lexa et al., 2005].These are important parameters when determining thedeformation mechanisms in deformed rocks [Kruse et al.,2001; Schmid et al., 1999; Ulrich et al., 2002].[11] Crystallographic preferred orientation (CPO) <strong>of</strong> quartz,plagioclase <strong>and</strong> K-feldspar was determined using the electronbackscattered diffraction method (EBSD using the CHAN-NEL5 s<strong>of</strong>tware) in manual mode. Each thin section wascalibrated by a silicone monocrystal <strong>and</strong> only measurementswith mean angular deviation below 1 were accepted <strong>and</strong> saved.The output data were plotted in pole figures (ftp://www.gm.univ-montp2.fr/mainprice//CareWare_Unicef_Programs/) <strong>and</strong>presented on lower hemisphere stereographic projections inwhich the trace <strong>of</strong> foliation is oriented in E-W direction. In case<strong>of</strong> K-feldspar <strong>and</strong> plagioclase, pole figures <strong>of</strong> published slipplanes <strong>and</strong> slip directions [Tullis, 1983;Kruse et al., 2001]were plotted in order to select <strong>and</strong> present possible operativeslip systems.4.1. Micro<strong>structural</strong> Analysis[12] Microscopic observations, grain size <strong>and</strong> shape characteristicsare described below for three different typesmarked by increasing degree <strong>of</strong> strain intensity. The type1 (metagranite) <strong>and</strong> type 2 (b<strong>and</strong>ed orthogneiss) revealcompletely recrystallized equigranular mineral aggregateswith similar shapes <strong>of</strong> mineral grains without relics <strong>of</strong>original magmatic porphyroclasts. Quartz aggregates consist<strong>of</strong> large grains with highly lobate mutual quartz boundaries<strong>and</strong> numerous ‘‘left-over grains’’ [Jessell, 1987]. Grainboundaries with neighboring feldspars as well as feldsparinclusions in quartz exhibit cuspate-lobate phase boundaries,typical for diffusional mass transfer along interphaseboundaries in high-temperature quartzo-feldspathic rocks[Gower <strong>and</strong> Simpson, 1992; Martelat et al., 1999]. Monomineralicfeldspar ribbons are composed <strong>of</strong> equidimensionalrecrystallized grains marked by straight grain boundariesshowing 120° triple point junctions. Average grain size,expressed as geometrical mean, ranges between 0.26 <strong>and</strong>0.32 mm for K-feldspar <strong>and</strong> between 0.17 <strong>and</strong> 0.26 mm forplagioclase. Grain size distribution <strong>of</strong> these minerals ismarked by low grain size spread (K-feldspar 0.2–0.21mm <strong>and</strong> plagioclase 0.15–0.18 mm) quantified as thedifference between third <strong>and</strong> first quartile (Table 1). Quartzexhibits average grain size ranging between 0.42 <strong>and</strong> 0.47mm <strong>and</strong> significantly larger grain size spread (Q 3 –Q 1 =0.23–0.35 mm) in comparison with feldspars. Axial ratio <strong>of</strong>all mineral phases is very low in conjunction with low SPO(see Table 1).[13] The microstructure <strong>of</strong> type 3 sample shows alternatingb<strong>and</strong>s <strong>of</strong> feldspars <strong>and</strong> quartz. Elongate quartz aggregates<strong>and</strong> one-grain thick ribbons consist <strong>of</strong> large quartzgrains with gently curved to straight mutual boundariesperpendicular to ribbon boundaries. Both feldspars formmonomineralic b<strong>and</strong>s with polygonal mosaic <strong>of</strong> feldspargrains with a triple junction network similar to microstructure<strong>of</strong> types 1 <strong>and</strong> 2. Grain size <strong>of</strong> both minerals hasconsiderably decreased to 0.12 mm <strong>and</strong> 0.9 mm for K-feldspar <strong>and</strong> plagioclase in comparison with type 1 <strong>and</strong> 2samples. The grain size spread also decreased (from averageQ 3 –Q 1 = 0.21–0.07 mm for K-feldspar <strong>and</strong> from averageQ 3 –Q 1 = 0.16 mm to 0.05 mm for plagioclase) <strong>and</strong> the axialratio together with SPO remain low (Table 1). In contrast,quartz exhibits significant increase in grain size (from0.47 mm in type 2 to 0.9 mm in type 3), grain size spread(from average 0.28 in types 1 <strong>and</strong> 2 to 0.93 in type 3) <strong>and</strong>SPO (Table 1).4.2. Crystallographic Preferred Orientations[14] The crystallographic preferred orientation <strong>of</strong> grains<strong>of</strong> all mineral phases was measured in two samples; amoderately deformed type 2 b<strong>and</strong>ed orthogneiss <strong>and</strong> ahighly strained type 3 mylonite using the EBSD method.Only quartz texture in sample B13 (type 2) was measuredusing the U stage.[15] K-feldspar <strong>and</strong> plagioclase in type 2 sample exhibitweak CPO that do not show any coincidence with publishedslip systems [Kruse et al., 2001; Tullis, 1983] (Figure 3).The only exception can be seen in case <strong>of</strong> K-feldspar thatshows maxima compatible with the activity <strong>of</strong> [110](001)slip system [Willaime et al., 1979]. C axes pattern <strong>of</strong> quartzis characterized by incomplete crossed girdles with twomaxima distributed along the periphery <strong>of</strong> the diagramhaving mutual distance <strong>of</strong> 90° <strong>and</strong> weak central submaximum.This type <strong>of</strong> quartz CPO corresponds to type I crossgirdle pattern <strong>of</strong> Lister <strong>and</strong> Hobbs [1980] that is typical forcombination <strong>of</strong> basal hai <strong>and</strong> prism hai slip systems active athigh-temperature plane strain deformation [Lister <strong>and</strong>Dornsiepen, 1982; Morgan <strong>and</strong> Law, 2004]. The form <strong>of</strong>quartz c axes pattern is consistent with the shape <strong>of</strong>measured plane strain ellipsoid <strong>of</strong> quartz aggregates(Figure 2b) [Passchier <strong>and</strong> Trouw, 1996].[16] K-feldspar in the type 3 sample shows crystallographicpreferred orientation marked by [100] directionsforming weak maximum close to the lineation (Figure 4).This slip direction operates on (010) planes [Willaime et al.,1979] that exhibit weak maximum in the YZ plane at anangle <strong>of</strong> 30° with respect to Y axis <strong>of</strong> the specimen’scoordinate system. Inspection <strong>of</strong> K-feldspar CPO patternrevealed that [001](120) slip system can be also active[Willaime <strong>and</strong> G<strong>and</strong>ais, 1977]. Plagioclase grains showweak CPO <strong>and</strong> reveal some activity <strong>of</strong> three slip directions,namely, [201], [100] <strong>and</strong> [101] on the (010) slip planereported by Marshall <strong>and</strong> McLaren [1977] <strong>and</strong> Olsen <strong>and</strong>Kohlstedt [1984]. Quartz CPO is characterized by two broadvertical parallel girdles <strong>of</strong> c axes indicating again combined4<strong>of</strong>15282


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 3. CPO data for K-feldspar, plagioclase <strong>and</strong> quartz from the type 2 (b<strong>and</strong>ed orthogneiss) sample(B13). Contoured at multiples <strong>of</strong> uniform distribution, lower hemisphere, equal-area projection. X <strong>and</strong> Zdesignate direction <strong>of</strong> lineation <strong>and</strong> pole to the foliation, respectively. Mineral abbreviations are afterKretz [1983].Figure 4. CPO data for K-feldspar, plagioclase, <strong>and</strong> quartz from the type 3 (mylonite) sample (B109).Contoured at multiples <strong>of</strong> uniform distribution, lower hemisphere, equal-area projection. X <strong>and</strong> Zdesignate direction <strong>of</strong> lineation <strong>and</strong> pole to the foliation, respectively. Mineral abbreviations are afterKretz [1983].5<strong>of</strong>15283


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210activity <strong>of</strong> basal hai <strong>and</strong> prism hai slip systems [Lister <strong>and</strong>Hobbs, 1980]. This CPO pattern is typical <strong>of</strong> constrictionaltype <strong>of</strong> deformation that is consistent with prolate symmetry<strong>of</strong> quartz aggregates from this sample [Passchier <strong>and</strong>Trouw, 1996].Figure 5. BSE image <strong>of</strong> the type 3 (mylonite) b<strong>and</strong>edmicrostructure. Note the circular inclusions <strong>of</strong> K-feldspar inquartz. Scale bar is 500 mm. Inset shows plagioclasecomposition (Or, orthoclase component; Ab, albite component).Rhombs indicate b<strong>and</strong>-forming plagioclase <strong>of</strong>oligoclase composition. Triangles indicate interstitial albitein K-feldspar aggregates or oligoclase rims. Mineralabbreviations are after Kretz [1983].5. Melt Topology[17] Investigation <strong>of</strong> melt topology is important for underst<strong>and</strong>ingthe possible influence <strong>of</strong> melt phase on operativedeformation mechanisms <strong>of</strong> its host aggregates. Melttopology also reflects grain-scale melt migration pathways[Marchildon <strong>and</strong> Brown, 2001; Rosenberg, 2001]. Sawyer[2001] reviewed criteria for recognition <strong>of</strong> former presence<strong>of</strong> melt on grain scale in deformed rocks. A typical feature<strong>of</strong> some rapidly quenched melting experiments is thedevelopment <strong>of</strong> melt pools with cuspate margins [Jurewicz<strong>and</strong> Watson, 1984] or thin melt films along crystal faces[Daines <strong>and</strong> Kohlstedt, 1997]. Melt phase crystallized asalbite, quartz or K-feldspar grains at triple point junctions orat crystal faces in residual aggregates have been commonlyreported from natural examples <strong>and</strong> interpreted in terms <strong>of</strong>melt topology in partially molten granite [e.g., Brown et al.,1999; Rosenberg <strong>and</strong> Riller, 2000; Marchildon <strong>and</strong> Brown,2001; Rosenberg, 2001].[18] In our study, polygonal mosaic <strong>of</strong> K-feldspar grainscontains numerous interstitial quartz <strong>and</strong> albite grains up to50 mm wide (Figures 5 <strong>and</strong> 6a) that extend along singleK-feldspar facets in the XZ section. Locally, narrowalbites (An 0.02 ) margin residual oligoclase (An 0,17 ) grainsFigure 6. Details <strong>of</strong> BSE images (type 3, mylonite) showing microstructures interpreted to mimic thetopology <strong>of</strong> crystallized interstitial melt. (a) Interstitial grains <strong>of</strong> quartz <strong>and</strong> albite between grains <strong>of</strong> K-feldspar aggregate developed especially on grain boundaries perpendicular to lineation (L). K-feldspar,light grey; quartz <strong>and</strong> albite, dark grey; scale bar 500 mm; XZ section. (b) Scarce triangular grain <strong>of</strong> K-feldspar between mica <strong>and</strong> oligoclase grain. Note fine exsolution lamella <strong>of</strong> albite adjacent to oligoclasegrain (arrow). K-feldspar, white; scale bar 250 mm; XZ section. (c) Albite rims on residual oligoclasegrains adjacent to K-feldspar aggregates (arrow 1), partial replacement <strong>of</strong> oligoclase grain by K-feldspar(arrow 2), amoeboid quartz grains at triple point junctions <strong>of</strong> plagioclase, adjacent to K-feldsparaggregates (arrow 3). Scale bar 250 mm; XZ section. (d) K-feldspar in host plagioclase grains (arrow 1),wedge-shaped albite grains within muscovite grains (arrow 2). Dark patches, holes in the specimen; scalebar 250 mm; XZ section. Mineral abbreviations are after Kretz [1983].6<strong>of</strong>15284


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 7. Block diagram <strong>of</strong> the mylonite sample microstructure.The melt field in the legend represents grains <strong>of</strong>phases assumed to have crystallized from interstitial melt.Note the melt topology; grain face pockets in XZ section<strong>and</strong> triple-point pockets in YZ section. Mineral abbreviationsare after Kretz [1983].(Figure 6c). Numerous interstitial round quartz grains withhigh dihedral angle occur in triple point junctions <strong>of</strong> thehost plagioclase aggregates <strong>of</strong>ten close to adjacent K-feldspar b<strong>and</strong>s (Figure 6c). We suggest that these grainsdo not represent crystallized melt but resulted from annealing<strong>of</strong> disintegrated myrmekite fronts [Hanmer, 1982]. Locally,triangular K-feldspar grains with thin albitic ‘‘exsolution’’rim adjacent to oligoclase grains occur in plagioclaseb<strong>and</strong>s (Figure 6b). Thin wedge-shaped albite grains aredeveloped along (001) cleavage planes <strong>of</strong> muscovite grainsin plagioclase aggregates (Figure 6d). Plagioclase grainsadjacent to muscovite grains exhibit a rectangular network<strong>of</strong> K-feldspar <strong>and</strong> albite, which is interpreted in terms <strong>of</strong>melt penetration along the cleavage planes <strong>of</strong> plagioclase(Figure 6d) [Mehnert et al., 1973; Dell’Angelo <strong>and</strong> Tullis,1988]. Quartz–K-feldspar boundaries show cuspate-lobateshapes <strong>and</strong> these boundaries are decorated with numeroussmall amoeboid grains <strong>of</strong> K-feldspar with high dihedral angleadjacent to the K-feldspar cusps (Figure 5).[19] Quartz <strong>and</strong> albite grains in K-feldspar aggregate,albite rims <strong>and</strong> scarce K-feldspar ‘‘pools’’ with albite‘‘exsolutions’’ in plagioclase b<strong>and</strong>s, K-feldspar in plagioclasegrains, together with albite grains within mica crystals<strong>and</strong> amoeboid grains <strong>of</strong> K-feldspar in quartz b<strong>and</strong>s satisfythe criteria for presence <strong>of</strong> crystallized mineral phasesassumed to mimic residual melt topology [Rosenberg <strong>and</strong>Riller, 2000; Sawyer, 2001] <strong>and</strong> will be designated as ‘‘themelt’’ for simplicity. K-feldspar component crystallizedfrom melt probably grew mostly onto the older grainswithin K-feldspar b<strong>and</strong>s, which would also explain theirperthite-free margins [Zulauf et al., 2002]. The crystallization<strong>of</strong> melt as unlike phases in residual aggregates commonlyresults in disappearance <strong>of</strong> the b<strong>and</strong>ed rock texture.Transitions from a b<strong>and</strong>ed orthogneiss into a homogeneousnonfoliated rock can be observed on a single outcrop in thestudied area. However, in type 2 <strong>and</strong> type 3 samples, themelt distributes preferentially in K-feldspar b<strong>and</strong>s.[20] The SEM imagery allowed identifying melt topologyalong intergranular voids as well as in intragranular fractures(Figures 5 <strong>and</strong> 6a). <strong>Quantitative</strong> analysis <strong>of</strong> meltpockets <strong>and</strong> their orientations allows creating a threedimensional(3-D) geometrical reconstruction <strong>of</strong> melt distributionin K-feldspar aggregates <strong>and</strong> analysis. Furthermore,we can define the geometrical relationship <strong>of</strong> intragranularfractures with crystal orientation <strong>and</strong> rock fabric.5.1. Intergranular Melt Topology[21] In order to depict <strong>and</strong> evaluate in detail the grainscalemelt distribution, grain boundaries on backscatteredelectron (BSE) images were traced <strong>and</strong> digitized from bothXZ <strong>and</strong> YZ sections (Figure 7). In moderately deformedorthogneiss type 2, melt in K-feldspar aggregate in XZsection occupies triple point junctions <strong>and</strong> extends into thinwedge-shaped melt films (maximum 30 mm in width) alonggrain faces perpendicular to the stretching lineation. Inhighly deformed orthogneiss type 3, the intergranular meltpockets (or seams [Rosenberg <strong>and</strong> Riller, 2000]) <strong>of</strong> aspectratio 2–3 show two submaxima in R f /f graph inclined at±20° relative to the apparent Z axis <strong>of</strong> the rock fabricellipsoid (Figure 8). Melt topology in YZ section <strong>of</strong> type3 orthogneiss is characterized by presence <strong>of</strong> melt ‘‘droplets’’at triple point junctions <strong>of</strong> K-feldspar grains. Rarely,the melt forms seams (30 mm in width) that line grainboundaries perpendicular to the foliation. These wide pocketsare likely to result from oblique sections <strong>of</strong> seamsoriented at high angle to the principle stretching direction.The melt topology from two perpendicular sections allowscreating a simplified 3-D geometrical model <strong>of</strong> interstitialmelt distribution in K-feldspar aggregate. This model ischaracterized as an interconnected network marked bymelt walls parallel to YZ plane, connected by tubes parallelto X direction along triple point boundaries between threegrains (Figure 9).5.2. Intragranular Fractures Locally Filled by Melt[22] A closer observation <strong>of</strong> XZ section BSE images <strong>of</strong>the type 3 (mylonite) sample revealed that some K-feldspargrains are crosscut by intragranular fractures (Figure 10a)<strong>and</strong> some <strong>of</strong> these fractures are filled with wedge-shapedmelt films. Orientations <strong>of</strong> traces in XZ section <strong>of</strong> theseFigure 8. R f /f diagram <strong>of</strong> the intergranular melt pocketswithin the K-feldspar aggregate from the XZ section.7<strong>of</strong>15285


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 9. Simplified 3-D melt topology model in K-feldspar aggregate <strong>of</strong> the mylonite type 3 orthogneiss. Seetext for explanation.fractures are represented by a rose diagram in Figure 10b<strong>and</strong> show strong maximum subparallel to the Z direction <strong>of</strong>the rock fabric ellipsoid <strong>and</strong> perpendicular to the horizontalstretching lineation (X direction). Furthermore, we havemeasured the CPO <strong>of</strong> grains affected by intragranularfractures, which exhibit clustering <strong>of</strong> poles to (001) planesin stretching lineation direction, while poles to (010) revealmaximum in the Y direction <strong>of</strong> the sample coordinatesystem (Figure 11). In order to evaluate relationship betweenorientations <strong>of</strong> crystallographic directions <strong>and</strong> intragranularfractures within corresponding grains, theorientations <strong>of</strong> fractures measured by U stage <strong>and</strong> graincrystallographic orientations obtained by means <strong>of</strong> EBSDhave been compared. Poles to all measured fractures exhibitstrong maximum close to the X direction, which is extendedtoward the Z direction (Figure 11). Analysis <strong>of</strong> anglesbetween pole <strong>of</strong> fracture <strong>and</strong> pole to (001) <strong>of</strong> K-feldsparcrystal shows that small angles characterize grains with theirpoles to (001) planes subparallel to the stretching lineation.In contrast, this angle is high for grains with (001) poles athigh angle to the stretching lineation (Figure 11). Thestatistics <strong>of</strong> angles between fracture orientations <strong>and</strong> (001)planes clearly shows that the majority <strong>of</strong> fractures aresubparallel to the (001) plane (in 52% <strong>of</strong> the fracturedgrains the angle is less than 30°, see Figure 12). Morphology<strong>of</strong> the fractures shows distinctive features for bothgroups <strong>of</strong> grains with low (60°) angle.In group 1 (60°), the grains show single curved fractures orinward tapering melt intrusions (Figure 10a). In addition,there exist intragranular curved fractures that were initiatedfrom triple point junctions <strong>and</strong> locally adjoin to the cleavage<strong>of</strong> the grain at their tips (Figure 10c).6. Melt Producing Reaction <strong>and</strong> Melt VolumeEstimates[23] An important issue is the origin <strong>of</strong> the interstitial meltdescribed in section 5. The melt could have been producedby in situ melting reaction or by infiltration from externalsources, which is reflected by the melt composition underestimated PT conditions <strong>and</strong> possible reaction textures[Sawyer, 2001]. Another important issue is the melt volumeproduced <strong>and</strong> volume change <strong>of</strong> the melting reaction, asboth these variables control the embrittlement <strong>of</strong> rock atdynamic conditions during breakdown reactions <strong>of</strong> somehydrous phase <strong>and</strong>/or at sufficient heating rates [Connolly etal., 1997; Rosenberg, 2001; Rushmer, 2001; Holyoke <strong>and</strong>Rushmer, 2002]. These data have critical impact on thedescribed melt topology, deformation mechanisms <strong>and</strong>rheology <strong>of</strong> studied orthogneiss <strong>and</strong> will be further discussed.Figure 10. (a) Detail <strong>of</strong> two fractured grains. Arrows <strong>and</strong> numbers designate group 1 <strong>and</strong> 2 fracturedgrains. Scale bar 100 mm. (b) Rose diagram <strong>of</strong> traces <strong>of</strong> the intragranular fracture orientations in K-feldspar b<strong>and</strong>s (XZ section <strong>of</strong> the sample coordinate system). (c) Sketch <strong>of</strong> curved <strong>and</strong> splittingintragranular fractures initiated from triple point junctions. Melt is gray; straight lines in the grains markthe cleavage.8<strong>of</strong>15286


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 11. Stereographic projection <strong>of</strong> CPO data <strong>and</strong> fracture poles <strong>of</strong> 100 fractured grains. X <strong>and</strong> Zdesignate direction <strong>of</strong> lineation <strong>and</strong> pole to the foliation, respectively. Symbols in the center projectiondesignate the misorientation angle <strong>of</strong> (001) planes with respect to the corresponding fracture pole (right).Lower hemisphere, equal-area projection, contoured at multiples <strong>of</strong> uniform distribution.[24] The rocks studied are represented by b<strong>and</strong>ed orthogneissesconsisting <strong>of</strong> K-feldspar (Kfs) (39%), plagioclase(Pl) (25%), quartz (Qtz) (26%), biotite (Bt), <strong>and</strong> muscovite(Ms) (9%) <strong>and</strong> minor garnet (Grt) (1%). P-T pseudosectionfor this rock was calculated with the Perple_X s<strong>of</strong>tware set[Connolly, 1990; Connolly <strong>and</strong> Petrini, 2002] using phaseend-members thermodynamic data by Holl<strong>and</strong> <strong>and</strong> Powell[1998]. Mixing properties <strong>of</strong> phases used for the calculationwere taken from Berman [1990] for garnet, Newton et al.[1980] for plagioclase, Thompson <strong>and</strong> Hovis [1979] for K-feldspar, Powell <strong>and</strong> Holl<strong>and</strong> [1999] for biotite <strong>and</strong> clinopyroxene,<strong>and</strong> White et al. [2001] for melt.[25] The obtained pseudosection calculated inNCKFMASH system with composition taken from thewhole rock analysis <strong>of</strong> a typical b<strong>and</strong>ed orthogneiss (sampleB13) is shown in Figure 13a. Stable mineral assemblageKfs-Pl-Qtz-Ms-Bt-Grt-melt, <strong>and</strong> composition <strong>of</strong> coexistingFe-Mg phases suggest that the highest metamorphic temperaturesassociated with melting are 700°C at a pressure <strong>of</strong>9.5 kbar. Calculated isopleths <strong>of</strong> the melt mode indicatethat for the given melt model <strong>and</strong> estimated P-T conditions,the rock contained 2–4 vol % <strong>of</strong> the melt phase(Figure 13a) at its metamorphic peak. This is, however, aminimum estimate, because the chosen system compositiondoes not take into account the amount <strong>of</strong> water releasedduring melt crystallization. The model <strong>of</strong> White et al. [2001]anticipates that the melt does not contain any CaO componentat the estimated PT conditions, which is in goodagreement with albitic composition <strong>of</strong> plagioclase crystallizedfrom the interstitial melt in the rock. Generalizedreaction leading to melt production in the stability field <strong>of</strong>assemblage Kfs-Pl-Qtz-Ms-Bt-Grt-melt can be derived fromthe changes in modal proportions <strong>of</strong> phases in the stabilityfield <strong>of</strong> interest (Figure 13b). Such changes suggest that theincrease in melt content is associated with the crystallization<strong>of</strong> garnet as a result <strong>of</strong> the reaction: Bt + Ms + Pl + Qtz = Grt +Kfs + melt.[26] The inferred melt-producing reaction correspondswell with the distribution <strong>of</strong> mineral phases in the rockmicrostructure. Small garnet grains (50 mm) form clusters onboundaries between plagioclase b<strong>and</strong>s <strong>and</strong> mica aggregatesin the vicinity <strong>of</strong> quartz grains (Figure 14). Agreement <strong>of</strong>modeled melt composition with that encountered in intergranularvoids <strong>of</strong> K-feldspar indicates that the melt migratedonly in between adjacent feldspar aggregates. Therefore wecan exclude melt loss or melt infiltration from externalsources, which would produce different reaction textures<strong>and</strong> could result in crystallization <strong>of</strong> interstitial plagioclasewith different composition.7. Deformation Mechanisms[27] Micro<strong>structural</strong> features in quartz observed in samples<strong>of</strong> all strain intensities reveal migrated grain boundaries<strong>and</strong> large grain size that increases even more during thedevelopment <strong>of</strong> type 3 microstructure. Microstructures <strong>and</strong>CPO patterns <strong>of</strong> quartz in type 2 <strong>and</strong> 3 samples show clearactivity <strong>of</strong> grain boundary migration (GBM) accommodateddislocation creep mechanism [Jessell, 1987; Hirth <strong>and</strong>Tullis, 1992] <strong>and</strong> CPO patterns correspond to the shapes<strong>of</strong> deformation inferred from the strain analysis.[28] Micro<strong>structural</strong> data show isometric shapes <strong>of</strong> polygonalfeldspar grains, absence <strong>of</strong> shape preferred orientation<strong>and</strong> weak or no CPO for both feldspars in type 2sample. These micro<strong>structural</strong> features <strong>and</strong> high finite strainintensities attained are indicative for grain boundary diffusionaccommodated grain boundary sliding (D gb -GBS) inboth feldspars [Poirier, 1985; Wadsworth et al., 1999].Although large grain size (K-feldspar 300 mm <strong>and</strong> plagioclase200 mm) is not characteristic for such deformationmechanism at experimental strain rates [Tullis <strong>and</strong> Yund,Figure 12. Statistics <strong>of</strong> angles between poles to (001)planes <strong>and</strong> corresponding fractures crosscutting the grainsrepresented by frequency histogram <strong>and</strong> a doughnutdiagram. Colors in the doughnut diagram represent thethree groups <strong>of</strong> grains as depicted in the histogram.9<strong>of</strong>15287


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 13. (a) P-T pseudosection for the sample B13 <strong>of</strong> granitic orthogneiss calculated in theNCKFMASH system <strong>and</strong> presented with quartz in excess. Isopleths <strong>of</strong> the volume percent <strong>of</strong> the melt inthe system are shown as solid black lines. Estimated PT conditions based on the position <strong>of</strong> compositionisopleths <strong>of</strong> coexisting garnet <strong>and</strong> biotite in the Kfs-Pl-Ms-Bt-Grt-melt stability field are marked by theblack field in the lower right inset. (b) Isobaric section (at 9 kbar) showing changes in modal proportions<strong>of</strong> stable phases with increasing temperature. Changes in stable mineral assemblages are shown as labeledsolid vertical lines. Mineral abbreviations are after Kretz [1983].1991], at natural conditions <strong>and</strong> presence <strong>of</strong> melt, whicheffectively enhances diffusion creep, it may be possible[Dell’Angelo et al., 1987]. In type 3 microstructure, increase<strong>of</strong> finite strain in both K-feldspar <strong>and</strong> plagioclase is associatedwith significant reduction <strong>of</strong> grain size (in contrast toquartz) <strong>and</strong> an important strengthening <strong>of</strong> CPO, which isremarkable especially in K-feldspar, while the shape preferredorientation (SPO) remains low. This is explained byincreased activity <strong>of</strong> dislocation creep [e.g., Tullis, 1983] <strong>of</strong>both feldspars in comparison with type 2 orthogneiss. Theopposite trend <strong>of</strong> grain size increase in quartz probablyreflects lower relative stress <strong>and</strong> strain dissipation in quartzwith respect to both feldspars. However, strong contribution<strong>of</strong> GBS accommodated by grain boundary diffusion flow(D gb -GBS) is considered to be responsible for accommodation<strong>of</strong> extreme finite strains in feldspars.[29] The suggested operative deformation mechanism infeldspars <strong>of</strong> type 3 sample (D gb -GBS + dislocation creep<strong>and</strong> climb within grains) was reported to result in extremefinite strains in several alloys deformed at relatively highstrain rates [Wadsworth et al., 1999; Weietal., 2003]. Finegrainedalloys undergoing D gb -GBS mechanism can beelongated up to several hundreds percents prior to creepfailure [Poirier, 1985; Čadek, 1988]. This behavior is called‘‘superplasticity’’ <strong>and</strong> it was also attributed to be responsiblefor deformation <strong>of</strong> some natural mylonitic rock b<strong>and</strong>s composed<strong>of</strong> fine (10 mm), isometricaly shaped <strong>and</strong> CPOlacking grains [Boullier <strong>and</strong> Gueguen, 1975; Behrmann<strong>and</strong> Mainprice, 1987]. In general, it is very little knownabout high-temperature GBS mechanism operating inquartzo-feldspathic rocks, although it is regarded as one <strong>of</strong>the most important strain accommodation mechanisms activein mineral aggregates (Langdon <strong>and</strong> Vastava [1982] as citedby Zhang et al. [1994]; also Ranalli [1995]).8. Discussion[30] Detailed <strong>analyses</strong> <strong>of</strong> different micro<strong>structural</strong> typespresented in this work provided a unique view on processesoperating in the studied rock at high metamorphic conditions<strong>and</strong> extreme finite strains. The mylonite type 3microstructure shows intergranular voids filled by melt<strong>and</strong> intragranular fractures both developed exclusively inK-feldspar aggregates. Micro<strong>structural</strong> observations showFigure 14. Micrograph <strong>of</strong> the studied orthogneiss indicatingthe melting reaction; scale bar 500 mm. Mineral abbreviationsare after Kretz [1983].10 <strong>of</strong> 15288


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 15. Illustration <strong>of</strong> the dilatancy mechanism <strong>and</strong>fracturing <strong>of</strong> K-feldspar grains. Lineation parallel to s 3 isvertical in geographic coordinates. (a) Low proportion <strong>of</strong>melt shuffled in intergranular spaces during GBS accommodatedby melt-enhanced diffusion creep along grainboundaries <strong>and</strong> dislocation creep within grains in bothfeldspars. (b) Dilatancy driven by cavitation or by meltoverpressure in K-feldspar that extracts the melt fromplagioclase intergranular films. Arrows designate thedirection <strong>of</strong> melt flux.that melt was present in both plagioclase <strong>and</strong> K-feldsparaggregates during deformation. At first, origin <strong>of</strong> melt-filledseams <strong>and</strong> intragranular fractures developed within K-feldsparaggregates is critically discussed using petrologicaldata. We further consider the processes responsible forenhancement <strong>of</strong> solid-state flow, given by the presence <strong>of</strong>melt phase, <strong>and</strong> the role <strong>of</strong> its redistribution on the rheology<strong>of</strong> polyphase <strong>and</strong> strongly deformed rock in terms <strong>of</strong>draining <strong>and</strong> accumulation <strong>of</strong> melt in different mineralaggregates. Finally, we discuss the apparent rheologicalrelationship between ‘‘weak’’ feldspars <strong>and</strong> ‘‘strong’’quartz.8.1. Cavitation Versus Fracturing Driven by MeltOverpressure[31] In studied rocks, the melt topology is characterizedby preferential distribution <strong>of</strong> melt in K-feldspar intergranularpockets, which is similar to experiments conductedwith low amounts <strong>of</strong> melt <strong>and</strong> at relatively high differentialstress [Daines <strong>and</strong> Kohlstedt, 1997; Gleason et al., 1999;Rosenberg, 2001; Holtzman et al., 2003]. In addition, K-feldspar grains are affected by fractures <strong>of</strong> the same orientation,which show affinity to the (001) cleavage. Ourpetrological investigations have shown that the melt wasproduced by metamorphic reaction within mica-plagioclaseb<strong>and</strong>s (source) <strong>and</strong> migrated into K-feldspar b<strong>and</strong>s (sink).This implies that individual mineral aggregates building therock can be characterized as open systems (Figure 15).There are two possible mechanisms, which could haveproduced the observed melt topology <strong>and</strong> intragranularfractures in K-feldspar.[32] If a fluid is introduced to the rock, e.g., by meltproducing metamorphic reaction, its pressure may overcomethe least principal stress s 3 as well as cohesion ortensile strength <strong>of</strong> suitably oriented planes in the aggregate[Hubbert <strong>and</strong> Rubey, 1959; Price <strong>and</strong> Cosgrove, 1990].These planes would in our case be represented by grainboundaries <strong>and</strong> (001) crystallographic planes oriented perpendicularto the stretching lineation (s 3 direction). As aresult <strong>of</strong> melt overpressure, the K-feldspar aggregate dilates<strong>and</strong> the melt is accumulated in intergranular pockets <strong>and</strong>some intragranular fractures. The aggregate hardens due toincreased frictional stress on melt free boundaries, becausemodeling <strong>of</strong> melt productivity suggests that no significantamount <strong>of</strong> melt was gained from external sources (‘‘dilationhardening’’ <strong>of</strong> Renner et al. [2000]).[33] The second model, explaining production <strong>of</strong> intergranularmelt pockets <strong>and</strong> intragranular fractures is theprocess <strong>of</strong> cavitation. Cavitation, formation <strong>of</strong> submicroscopiccavities driven primarily by diffusion <strong>and</strong> accumulation<strong>of</strong> vacancies, is the direct consequence <strong>of</strong> GBS. WhenGBS cannot be fully compensated by diffusion <strong>and</strong>/ordislocation creep controlled grain shape change, cavitiesnucleate at first on grain boundaries at high angle to thetensional direction (direction <strong>of</strong> s 3 )[Čadek, 1988; Kassner<strong>and</strong> Hayes, 2003]. Cavities then grow <strong>and</strong> coalesce witheach other to form intergranular voids. Further cavitycoalescence can be caused by nucleation <strong>and</strong> growth <strong>of</strong>cavities on grain boundaries at low angle to the tensionaldirection due to (1) formation <strong>of</strong> tensile <strong>and</strong> compressiveledges, where boundaries are not straight due to inhomogeneousplastic deformation <strong>of</strong> the grains or (2) dislocationpileups at grain boundaries (Zener-Stroh mechanism) oraround impurities in crystal lattice (Figure 16) [Vollbrecht etal., 1999; Kassner <strong>and</strong> Hayes, 2003]. Plastic deformation isfurther being localized to the ‘‘bridges’’ (intact grain boundarysegments) affected by further cavitation. The bridgesloose its stability <strong>and</strong> locally contract to form finally afracture which is driven by coalescence <strong>of</strong> cavities at its tip(Figure 16). In this way, clear macroscopic intragranularfractures develop [Čadek, 1988].[34] According to the cavitation model, opening <strong>of</strong> intergranularvoids <strong>and</strong> intragranular fractures during GBSproduces local underpressure (the volume <strong>of</strong> the systemincreases due to creation <strong>of</strong> voids in K-feldspar <strong>and</strong> thegrains support the voids as a load bearing framework),Figure 16. (a) Illustration <strong>of</strong> the cavitation mechanismgenerated by GBS (grain boundary sliding) <strong>and</strong> formation<strong>of</strong> intergranular <strong>and</strong> intragranular fractures at the onset <strong>of</strong>final creep failure <strong>of</strong> the aggregate. T <strong>and</strong> C designate thetensional <strong>and</strong> compressive sector <strong>of</strong> a grain boundary ledge.Z-S designates illustration <strong>of</strong> the Zener-Stroh mechanism <strong>of</strong>cavity formation [Kassner <strong>and</strong> Hayes, 2003]. See descriptionin text.11 <strong>of</strong> 15289


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 17. PT plot with contours <strong>of</strong> molar volume in Jbar 1 (equivalent to 10 1 cm 1 mol1 ) calculated for thesame system as shown in Figure 13. Mineral abbreviationsare after Kretz [1983].which soaks up the melt produced within plagioclase b<strong>and</strong>s(the melt is passively redistributed, Figure 15). Consequently,plagioclase could not deform any further due to the loss <strong>of</strong>intergranular melt that worked as a creep enhancing medium.[35] It has been shown that creation <strong>of</strong> melt-bearingporosity can be induced by fast melt production due tobreakdown <strong>of</strong> some hydrous phase in the rock (e.g., viadiscontinuous reaction) or due to high heating rate causingsubstantial overstepping <strong>of</strong> continuous melting reactions<strong>and</strong> thus fast melt production as a result <strong>of</strong> the system’seffort to reach equilibrium state [e.g., Connolly et al., 1997;Rushmer, 2001]. In contrast, change in molar volume <strong>of</strong> therock during slow progress <strong>of</strong> continuous reactions will beprobably small.[36] In order to test this hypothesis, the P-T phasediagram section in Figure 13a was contoured for molarvolume <strong>of</strong> the system to see, whether the observed meltingreaction at estimated PT conditions may lead to rapidvolume increase. The result (Figure 17) shows that the onlyrapid increase in molar volume <strong>of</strong> the rock in given PTrange results from dehydration melting <strong>of</strong> muscovite at PTconditions higher than those reached by studied orthogneisssamples. In addition, the role <strong>of</strong> melt volume change <strong>and</strong>velocity <strong>of</strong> the reaction is speculative for aggregates with‘‘incoherent’’ grain boundaries like both feldspars undergoingGBS in studied orthogneiss than rather ‘‘intact’’ solidphases around reaction sites in the experiments <strong>of</strong> Connollyet al. [1997].[37] The calculated volume <strong>of</strong> melt (2–4 vol %) suggeststhat the melt formed isolated melt films, pools or pockets. Inthe case <strong>of</strong> ‘‘overpressure’’ model, sufficient amount <strong>of</strong> meltneeds to be produced to induce considerable melt overpressure<strong>and</strong> melt redistribution. This critical melt volumecorresponds to creation <strong>of</strong> interconnected network <strong>of</strong> meltbetween the framework <strong>of</strong> grains, so that the melt can startincreasing its hydrostatic pressure toward the level <strong>of</strong>maximum compressive stress [Renner et al., 2000]. Thismelt volume is given by the melt connectivity transition(MCT) <strong>of</strong> F =0.07[Rosenberg <strong>and</strong> H<strong>and</strong>y, 2005] or similar‘‘liquid percolation threshold’’ (LPT) <strong>of</strong> F =0.08[Vigneresseet al., 1996], although this critical level is likely to depend onseveral variables (e.g., dihedral angles). It is possible that themelt migrates from locally overpressured isolated pockets athigh angle to the maximum compressive stress (s 1 )tointergranular boundaries subparallel with this direction(s 1 ). However, it is unlikely that the cohesion <strong>of</strong> grainboundaries <strong>and</strong> tensile strength <strong>of</strong> (001) intragranular cleavage<strong>of</strong> K-feldspar are significantly smaller than intergranularcohesion <strong>of</strong> plagioclase, which would be necessary conditionfor preferential hydraulic failure <strong>of</strong> K-feldspar, although themelt was produced in plagioclase b<strong>and</strong>s.[38] In contrast, during cavitation, underpressure arisingfrom opening <strong>of</strong> incipient intergranular voids in K-feldsparis theoretically susceptible to ‘‘extract’’ melt from isolatedpools <strong>and</strong> films within plagioclase aggregates, in spite <strong>of</strong> itslittle amount. If distortion <strong>of</strong> plagioclase grains is easierthan for K-feldspar grains to accommodate GBS, then plagioclaseaggregates should also ‘‘extrude’’ excessive intergranularmelt into possible sink sites. This is in agreementwith analogue modeling results <strong>of</strong> Walte et al. [2005], whereaggregates composed <strong>of</strong> ‘‘weaker’’ grains (more susceptibleto plastic deformation) show higher ‘‘GBS locking transition’’than relatively stronger grains. Melt migration alongmutual boundaries <strong>of</strong> quartz was probably negligible incomparison to both feldspars. This is again controlled byrelative ‘‘weakness’’ <strong>of</strong> quartz grains with respect to ‘‘stronger’’grains <strong>of</strong> both feldspars <strong>and</strong> ‘‘welded’’ boundaries <strong>of</strong>quartz. The viscosity ratio between melt <strong>and</strong> host aggregategrains thus increases from quartz to plagioclase <strong>and</strong> ishighest for K-feldspar [Walte et al., 2005].[39] Considering little amount <strong>of</strong> melt produced by themelting reaction <strong>and</strong> absence <strong>of</strong> intergranular melt pocketsperpendicular to the stretching lineation in plagioclase, weconsider the ‘‘cavitation’’ model to be the most likelymicrophysical mechanism explaining the dilation, micr<strong>of</strong>racturation<strong>and</strong> preferential melt distribution in K-feldsparaggregates <strong>of</strong> type 3 microstructure.8.2. Impact <strong>of</strong> Impurities <strong>and</strong> MechanicalAnisotropy <strong>of</strong> Crystals on GBS[40] Experiments with aluminum in GBS regime withrelatively high-purity metals (e.g., aluminum <strong>of</strong> purity 4N(where N expresses the degree <strong>of</strong> purity <strong>of</strong> a material; 2N =99% purity, 4N = 99.99% purity, etc.), 650°K, strain rate10 9 s 1 ) or at high strain rates (e.g., 4N copper at 10 2 s 1 ,873°K [Čadek, 1988]) are marked by high relative elongationse, distinct contraction at the area <strong>of</strong> failure (necking)associated with intracrystalline fracturing. On theother h<strong>and</strong>, experiments with relatively low-purity metals(2N aluminum at 10 9 s 1 , 650°K) or at low strain rates(4N copper at 10 9 s 1 , 873 K) show small relativeelongations to failure, regular contraction along the deformedspecimen (homogeneous strain) <strong>and</strong> intergranularfracturing (Figure 18) [Sklenička et al., 1977; Čadek, 1988;Mohamed, 2002]. Therefore the higher purity material canwithst<strong>and</strong> higher relative elongations to creep failure (<strong>and</strong> atfaster strain rates) than its lower purity equivalent. Incontrast, stress accumulations around impurities <strong>and</strong> grainboundary ledges enhance diffusion-driven cavitation, cavitieseasily form at dislocation pileups around impuritieswithin the grains <strong>and</strong> creep failure will occur after smallerrelative elongations (<strong>and</strong> slower strain rates).12 <strong>of</strong> 15290


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Figure 18. Final creep failure microstructure <strong>and</strong> specimendistortion from tensile experiments with metals atconstant strain rate. [Sklenička et al., 1977; Čadek, 1988].Material <strong>of</strong> purity 2N contains more impurities than 4Nmaterial.[41] These results can be compared with different intensities<strong>of</strong> deformation <strong>and</strong> microstructures <strong>of</strong> both feldspars inthe studied orthogneiss (type 3 microstructure) sample.Feldspars can be also regarded as solid solutions like alloysor impure metals. We consider the three component K-feldspar with clearly defined solvus as a ‘‘less pure’’ equivalentto the two component oligoclase. This interpretation issupported by the presence <strong>of</strong> cryptoperthite exsolutionswithin some K-feldspar grains (lower left part <strong>of</strong>Figure 10a), along which the tips <strong>of</strong> cavitation drivenfractures could nucleate <strong>and</strong> propagate. The perthite intergrowthsare commonly oriented along (001) plane [Spry,1969], which is fully compatible with orientation <strong>of</strong> most <strong>of</strong>the intragranular fractures reported in this work. In contrast,plagioclase shows higher finite strains in type 3 orthogneissthan K-feldspar <strong>and</strong> no intragranular fractures, which isconsistent with higher purity aluminum experiments orhigher purity metals in general [Čadek, 1988; Mohamed,2002]. Fractured grains in K-feldspar exhibit strong crystallographicpreferred orientation marked by (001) planesperpendicular to the axis <strong>of</strong> stretching direction (X direction)<strong>and</strong> their (010) plane coinciding with the Y direction <strong>of</strong> therock fabric’s coordinate system. Because propagation <strong>of</strong>fractures (group 1 fractures, Figure 10a) driven by coalescence<strong>of</strong> cavities is initiated from boundaries oriented at highangles to the maximum principal compressive stress direction,we can suggest that cavity coalescence was mosteffective along the K-feldspar (001) plane in [100] crystallographicdirection. Selection <strong>of</strong> suitably oriented K-feldspargrains for intragranular cavitation reflects the strong mechanicalcrystallographic anisotropy <strong>of</strong> K-feldspar [Smith,1974]. Group 2 fractures (Figure 10a) <strong>and</strong> curved or splittingfractures (Figure 10c) could develop due to cavitation <strong>and</strong>propagation <strong>of</strong> fractures from triple-point junction boundaries.The propagation <strong>of</strong> fractures in K-feldspars in the case<strong>of</strong> ‘‘melt overpressure’’ model would probably follow similarpathways as cavitation-driven fractures.8.3. Effect <strong>of</strong> Melt Phase on GBS[42] Several tensile experiments on creep properties <strong>of</strong>metals have shown dominantly GBS accommodated superplasticbehavior with peak <strong>of</strong> finite elongations at temperaturescoinciding with occurrence <strong>of</strong> melt phase along grain boundaries[Mabuchi et al., 1997]. This weakening was associatedwith effective contribution <strong>of</strong> the melt phase to annihilation <strong>of</strong>gliding dislocations, therefore prevention <strong>of</strong> dislocation pileups,increased velocity <strong>of</strong> diffusive mass transfer <strong>and</strong> accommodation<strong>of</strong> stress concentrations that was also reflected bylower cavity densities in comparison with ‘‘melt-free’’ experiments[Koike et al., 1998]. Melt-enhanced weakening isfavored by low wettability <strong>of</strong> melt (high dihedral angles) <strong>and</strong>its relatively small amount. In contrast, high wettability (lowdihedral angle) melts thickly coating grain boundaries producedpremature necking <strong>and</strong> creep failure due to the loss <strong>of</strong>grain boundary cohesion [Mabuchi et al., 1997; Koike et al.,1998]. The same transition from superplastic flow to prematurefailure coinciding with almost complete wetting <strong>of</strong> grainboundaries (at length proportion <strong>of</strong> 70%) was demonstratedalso during compressive experiments [Pharr et al., 1989;Baudelet et al., 1992].[43] In summary, the material science experimentalresults show distinct rheological transition between melt(liquid)–enhanced GBS <strong>and</strong> collapse marked by increasedcavitation velocities at a melt volume threshold coincidingwith complete wetting <strong>of</strong> grain boundaries. This threshold issimilar to the concept <strong>of</strong> melt embrittlement at the onset <strong>of</strong>melt connectivity transition (MCT) proposed by Rosenberg<strong>and</strong> H<strong>and</strong>y [2005] from experimental data on quartz<strong>of</strong>eldspathicrocks. Collapse (rather than weakening) at theMCT was associated with localized intergranular <strong>and</strong> intragranularmicrocracking, frictional sliding <strong>and</strong> limited bodyrotation leading to the development <strong>of</strong> cataclastic zones<strong>and</strong> produces heterogeneous <strong>and</strong> restricted deformation[Rutter <strong>and</strong> Neumann, 1995; Rosenberg <strong>and</strong> H<strong>and</strong>y,2005]. Rosenberg <strong>and</strong> H<strong>and</strong>y [2005] have regarded MCTas a rheological threshold even more important than thepreviously emphasized rheological critical melt percentage(RCMP) at F = 0.1–0.3 [Arzi, 1978; van der Molen <strong>and</strong>Paterson, 1979] at experimental conditions. However, theweakening mechanisms associated with MCT are incompatiblewith our observations.[44] Our strain measurements <strong>and</strong> material science experimentalresults show that weakening associated with meltenhancedGBS at low melt volumes (below the MCT) issubstantial <strong>and</strong> the aggregate is deformed relatively homogeneously.We therefore suggest that the degree <strong>of</strong> weakeninginduced by melt-enhanced GBS can be higher <strong>and</strong>more important on large scales at natural conditions than theweakening mechanisms operative at experimental strainrates at the onset <strong>of</strong> melt connectivity transition as proposedby Rosenberg <strong>and</strong> H<strong>and</strong>y [2005].[45] The critical melt volume for transition from meltenhancedGBS to premature creep failure due to increasedmelt volume is likely to depend on several variables, such asdifferential stress [e.g., Bordeaux et al., 1994; Rosenberg,2001], strain rate, grain size [Bordeaux et al., 1994; Renner etal., 2000], melt volume <strong>and</strong> melt framework viscosity ratio[McKenzie, 1984; Pharr et al., 1989; Walte et al., 2005].[46] The presence <strong>of</strong> water without melt in the systemcould effectively enhance diffusion <strong>and</strong> dislocation creepaccommodated GBS <strong>of</strong> feldspars <strong>and</strong> produce similarmicrostructures as in this study [Tullis <strong>and</strong> Yund, 1991].However, micro<strong>structural</strong> observations <strong>and</strong> petrological datashow that melt was present in the system during deformation.The combined effect <strong>of</strong> water <strong>and</strong> melt on GBSenhancement is speculative (according to the weakeningmechanisms <strong>of</strong> melt discussed above), because the influence<strong>of</strong> water on dihedral angles <strong>of</strong> melt in quartzo-feldspathicrocks is not yet well understood [Rushmer, 1996].13 <strong>of</strong> 15291


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB102108.4. Comparative Rheology Between Feldspars <strong>and</strong>Quartz[47] Experimental studies show that the creep strength <strong>of</strong>feldspars deformed at dislocation creep is higher than that <strong>of</strong>quartz for the entire range <strong>of</strong> temperature conditions investigated[Shelton <strong>and</strong> Tullis, 1981; Jaoul et al., 1984; H<strong>and</strong>y,1994]. This assumption is essential in assessing rheology <strong>of</strong>polyphase rocks with different proportion <strong>of</strong> weak quartz<strong>and</strong> ‘‘strong skeleton’’ formed by feldspars in both loadbearingframework <strong>and</strong> interconnected weak layer microstructures[H<strong>and</strong>y, 1990, 1994; Schulmann et al., 1996; Ji etal., 2004; Rybacki <strong>and</strong> Dresen, 2004]. However, change <strong>of</strong>deformation mechanism in feldspars from dislocation todiffusion creep has never been considered in term <strong>of</strong>mineral strength contrast between quartz <strong>and</strong> feldspars.Exceptionally high finite strains <strong>of</strong> both feldspars surroundingmuch less elongated quartz ribbons shows well thatquartz represents the strong phase according to H<strong>and</strong>y[1990, 1994] <strong>and</strong> feldspars form interconnected weaklayers. The explanation is seen in important contribution<strong>of</strong> dislocation creep accommodated GBS mechanism to thedominant melt-enhanced grain boundary diffusion creepoperative in both feldspars that significantly decreases theirstrength with respect to quartz simultaneously deforming viadislocation creep-controlled recrystallization mechanisms.[48] Acknowledgments. This work has been kindly supported byGrant Agency <strong>of</strong> the Czech Academy <strong>of</strong> Sciences (GAAV) projectIAA3111401. J.K. appreciates the financial support <strong>of</strong> the Ministry <strong>of</strong>Education, Youth <strong>and</strong> Sports <strong>of</strong> the Czech Republic through the ScientificCentre ‘‘Advanced Remedial Technologies <strong>and</strong> Processes’’ (identificationcode 1M0554). We appreciated constructive reviews <strong>of</strong> P. Bons, J. Weertman,<strong>and</strong> an anonymous material scientist <strong>and</strong> editorial h<strong>and</strong>ling <strong>of</strong> J. C. Mutter.This study was also made possible thanks to the ANR project ‘‘LFO inorogens’’ funding as well as to financial support <strong>of</strong> CNRS (UMR 7516) toOndrej Lexa.ReferencesArzi, A. (1978), Critical phenomena in the rheology <strong>of</strong> partially meltedrocks, Tectonophysics, 44, 173–184.Baudelet, B., M. C. Dang, <strong>and</strong> F. Bordeaux (1992), Mechanical behaviour<strong>of</strong> an aluminium alloy with fusible grain boundaries, Scr. Metall. Mater.,26(4), 573–578.Behrmann, J. H., <strong>and</strong> D. Mainprice (1987), Deformation mechanisms in ahigh-temperature quartz-feldspar mylonite: evidence for superplastic flowin the lower continental crust, Tectonophysics, 140, 297–305.Berman, R. G. (1990), Mixing properties <strong>of</strong> Ca-Mg-Fe-Mn garnets, Am.Mineral., 75, 328–344.Bordeaux, F., M. C. Dang, <strong>and</strong> B. Baudelet (1994), Percolation <strong>and</strong> damagebehaviour in materials with liquid grain boundaries, J. Mater. Sci., 29,3023–3030.Boullier, A. M., <strong>and</strong> Z. Gueguen (1975), SP-Mylonites: origin <strong>of</strong> somemylonites by superplastic flow, Contrib. Mineral. Petrol., 50, 93–104.Brown, M., <strong>and</strong> T. Rushmer (1997), The role <strong>of</strong> deformation in the movement<strong>of</strong> granitic melt: views from the laboratory <strong>and</strong> the field, in Deformation-EnhancedMelt Segregation <strong>and</strong> Metamorphic Fluid Transport,Mineral. Soc. Ser., vol. 8, edited by M. Holness, pp. 111 –144, Chapman<strong>and</strong> Hall, London.Brown, M. A., M. Brown, W. D. Carlson, <strong>and</strong> C. Denison (1999), Topology<strong>of</strong> syntectonic melt-flow networks in the deep crust: Inferences fromthree-dimensional images <strong>of</strong> leucosome geometry in migmatites, Am.Mineral., 84, 1793–1818.Čadek, J. (1988), Creep in Metallic Materials, 372 pp., Elsevier, Amsterdam,Netherl<strong>and</strong>s.Connolly, J. A. D. (1990), Multivariable phase-diagrams—An algorithmbased on generalized thermodynamics, Am. J. Sci., 290(6), 666–718.Connolly, J. A. D., <strong>and</strong> K. Petrini (2002), An automated strategy for calculation<strong>of</strong> phase diagram sections <strong>and</strong> retrieval <strong>of</strong> rock properties as afunction <strong>of</strong> physical conditions, J. Metamorph. Geol., 20, 697–708.Connolly, J. A. D., M. B. Holness, D. C. Rubie, <strong>and</strong> T. Rushmer (1997),Reaction-induced microcracking: An experimental investigation <strong>of</strong> amechanism for enhancing anatectic melt extraction, Geology, 25(7),591–594.Cooper, R. F., <strong>and</strong> D. L. Kohlstedt (1984), Solution-precipitation-enhanceddiffusional creep <strong>of</strong> partially molten olivine-basalt aggregates during hotpressing,Tectonophysics, 107(3–4), 207–233.Cosgrove, J. W. (1997), The influence <strong>of</strong> mechanical anisotropy on thebehaviour <strong>of</strong> the lower crust, Tectonophysics, 280(1–2), 1 –14.Daines, M. J., <strong>and</strong> D. L. Kohlstedt (1997), Influence <strong>of</strong> deformation on melttopology in peridotites, J. Geophys. Res., 102, 10,257–10,271.Dell’Angelo, L. N., <strong>and</strong> J. Tullis (1988), Experimental deformation <strong>of</strong> partiallymelted granitic aggregates, J. Metamorph. Geol., 6(4), 495–515.Dell’Angelo, L. N., J. Tullis, <strong>and</strong> R. A. Yund (1987), Transition fromdislocation creep to melt-enhanced diffusion creep in fine-grained graniticaggregates, Tectonophysics, 139(3–4), 325–332.Flinn, D. (1962), On folding during three-dimensional progressive deformation,Q. J. Geol. Soc. London, 118, 385–433.Franke, W. (2000), The mid-European segment <strong>of</strong> the Variscides: tectonostratigraphicunits, terrane boundaries <strong>and</strong> plate tectonic evolution, inOrogenic Processes: Quantification <strong>and</strong> Modelling in the Variscan Belt,edited by W. Franke et al., Geol. Soc. Spec. Publ., 179, 35–61.Gabriel, A., <strong>and</strong> E. P. Cox (1929), A staining method for the quantitativedetermination <strong>of</strong> certain rock minerals, Am. Mineral., 14, 290–292.Gleason, G. C., V. Bruce, <strong>and</strong> H. W. Green (1999), Experimental investigation<strong>of</strong> melt topology in partially molten quartzo-feldspathic aggregatesunder hydrostatic <strong>and</strong> non-hydrostatic stress, J. Metamorph. Geol., 17,705–722.Gower, R. J. W., <strong>and</strong> C. Simpson (1992), Phase-boundary mobility innaturally deformed, high-grade quartzo-feldspathic rocks—Evidence fordiffusional creep, J. Struct. Geol., 14, 301–313.H<strong>and</strong>y, M. R. (1990), The solid-state flow <strong>of</strong> polymineralic rocks, J. Geophys.Res., 95, 8647–8661.H<strong>and</strong>y, M. R. (1994), Flow laws for rocks containing 2 nonlinear viscousphases—A phenomenological approach, J. Struct. Geol., 16, 287–301.Hanmer, S. (1982), Microstructure <strong>and</strong> geochemistry <strong>of</strong> plagioclase <strong>and</strong>microcline in naturally deformed granite, J. Struct. Geol., 4, 232.Hirth, G., <strong>and</strong> J. Tullis (1992), Dislocation creep regimes in quartz aggregates,J. Struct. Geol., 14, 145–159.Holl<strong>and</strong>, T. J. B., <strong>and</strong> R. Powell (1998), An internally consistent thermodynamicdata set for phases <strong>of</strong> petrological interest, J. Metamorph. Geol.,16(3), 309–343.Hollister, L. S., <strong>and</strong> M. L. Crawford (1986), Melt-enhanced deformation—A major tectonic process, Geology, 14(7), 558–561.Holtzman, B. K., N. J. Groebner, M. E. Zimmerman, S. B. Ginsberg, <strong>and</strong>D. L. Kohlstedt (2003), Stress-driven melt segregation in partially moltenrocks, Geochem. Geophys. Geosyst., 4(5), 8607, doi:10.1029/2001GC000258.Holyoke, C. W., <strong>and</strong> T. Rushmer (2002), An experimental study <strong>of</strong> grainscale melt segregation mechanisms in two common crustal rock types,J. Metamorph. Geol., 20(5), 493–512.Hradecký, P. (Ed.) (2002), Geological map 1:25000, Map 11-221. CzechGeol. Surv., Prague, Czech Republic.Hubbert, M. K., <strong>and</strong> W. W. Rubey (1959), Role <strong>of</strong> fluid pressure in mechanics<strong>of</strong> overthrust faulting, Geol. Soc. Am. Bull., 70, 116–166.Jaoul, O., J. Tullis, <strong>and</strong> A. Kronenberg (1984), The effect <strong>of</strong> varying watercontent on the behaviour <strong>of</strong> Heavitree quartzite, J. Geophys. Res., 89,4298–4312.Jessell, M. W. (1987), Grain boundary migration microstructures in a naturallydeformed quartzite, J. Struct. Geol., 9, 1007–1014.Ji, S. C., Z. T. Jiang, E. Rybacki, R. Wirth, D. Prior, <strong>and</strong> B. Xia (2004),Strain s<strong>of</strong>tening <strong>and</strong> micro<strong>structural</strong> evolution <strong>of</strong> anorthite aggregates <strong>and</strong>quartz-anorthite layered composites deformed in torsion, Earth Planet.Sci. Lett., 222(2), 377–390.Jurewicz, S. R., <strong>and</strong> E. B. Watson (1984), Distribution <strong>of</strong> partial melt in afelsic system - the importance <strong>of</strong> surface energy, Contrib. Mineral. Petrol.,85(1), 25–29.Kassner, M. E., <strong>and</strong> T. A. Hayes (2003), Creep cavitation in metals, Int. J.Plasticity, 19(10), 1715–1748.Koike, J., K. Miki, K. Maruyama, <strong>and</strong> H. Oikawa (1998), Effects <strong>of</strong> theliquid phase on the high-temperature tensile ductility: from embrittlementto superplasticity, Philos. Mag. A, 3, 599–614.Konopásek, J., <strong>and</strong> K. Schulmann (2005), Contrasting Early Carboniferousfield geotherms: Evidence for accretion <strong>of</strong> a thickened orogenic root <strong>and</strong>subducted Saxothuringian crust (central European Variscides), J. Geol.Soc., 162, 463–470.Konopásek, J., K. Schulmann, <strong>and</strong> O. Lexa (2001), Structural evolution <strong>of</strong>the central part <strong>of</strong> the Krušné hory (Erzgebirge) Mountains in the CzechRepublic—Evidence for changing stress regime during Variscan compression,J. Struct. Geol., 23, 1373–1392.Kotková, J., A. Kröner, W. Todt, <strong>and</strong> J. Fiala (1996), Zircon dating <strong>of</strong> northBohemian granulites, Czech Republic: Further evidence for the Lower14 <strong>of</strong> 15292


B10210ZÁVADA ET AL.: EXTREME DUCTILITY OF FELDSPAR AGGREGATESB10210Carboniferous high-pressure event in the Bohemian Massif, Geol.Rundsch., 85(1), 154–161.Kretz, R. (1983), Symbols for rock-forming minerals, Am. Mineral., 68,277–279.Kruse, R., H. Stünitz, <strong>and</strong> A. Kunze (2001), Dynamic recrystallizationprocesses in plagioclase porphyroclasts, J. Struct. Geol., 23, 1781–1802.Langdon, T. G., <strong>and</strong> R. B. Vastava (1982), An evaluation <strong>of</strong> deformationmodels for grain boundary sliding, in Mechanical Testing for DeformationModel Development, edited by R. W. Rohde <strong>and</strong> J. C. Swearengen,Am. Soc. for Test. <strong>of</strong> Mater., Philadelphia, Pa.Lexa, O., P. Štípská, K. Schulmann, L. Baratoux, <strong>and</strong> A. Kröner (2005), Contrastingtextural record <strong>of</strong> two distinct metamorphic events <strong>of</strong> similar P-Tconditions <strong>and</strong> different durations, J. Metamorph. Geol., 23(8), 649–666.Lister, G. S., <strong>and</strong> U. F. Dornsiepen (1982), Fabric transitions in the Saxonygranulite terrain, J. Struct. Geol., 4, 81–92.Lister, G. S., <strong>and</strong> B. E. Hobbs (1980), The simulation <strong>of</strong> fabric developmentduring plastic deformation <strong>and</strong> its application to quartzite: The influence<strong>of</strong> deformation history, J. Struct. Geol., 2, 355–370.Mabuchi, M., H. G. Jeong, K. Hiraga, <strong>and</strong> K. Higashi (1997), Partialmelting at interfaces <strong>and</strong> grain boundaries for high-strain-rate superplasticmaterials, Interface Sci., 4(3–4), 357–368.Marchildon, N., <strong>and</strong> M. Brown (2001), Melt segregation in late syn-tectonicanatectic migmatites: An example from the Onawa contact aureole,Maine, USA, Phys. Chem. Earth, Part A, 26(4–5), 225–229.Marshall, D. B., <strong>and</strong> A. C. McLaren (1977), Deformation mechanisms inexperimentally deformed plagioclase feldspars, Phys. Chem. Mineral., 1,351–370.Martelat, J. E., K. Schulmann, J. M. Lardeaux, C. Nicollet, <strong>and</strong> H. Cardon(1999), Granulite micr<strong>of</strong>abrics <strong>and</strong> deformation mechanisms in southernMadagascar, J. Struct. Geol., 21, 671–687.McKenzie, D. (1984), The generation <strong>and</strong> compaction <strong>of</strong> partially moltenrock, J. Petrol., 25(3), 713–765.Mehnert, K. R., W. Busch, <strong>and</strong> G. Schneider (1973), Initial melting at grainboundaries <strong>of</strong> quartz <strong>and</strong> feldspar in gneisses <strong>and</strong> granulites, Neues Jahb.Mineral., 4, 165–183.Mohamed, F. A. (2002), The role <strong>of</strong> impurities during creep <strong>and</strong> superplasticityat very low stresses, Metall. Mater. Trans. A., 33, 261–278.Morgan, S. S., <strong>and</strong> R. D. Law (2004), Unusual transition in quartzite dislocationcreep regimes <strong>and</strong> crystal slip systems in the aureole <strong>of</strong> theEureka Valley-Joshua Flat-Beer Creek pluton, California: a case for anhydrousconditions created by decarbonation reactions, Tectonophysics,384(1–4), 209–231.Newton, R. C., T. V. Charlu, <strong>and</strong> O. J. Kleppa (1980), Thermochemistry <strong>of</strong> high<strong>structural</strong> state plagioclases, Geochim. Cosmochim. Acta, 44, 933–941.Olsen, T. S., <strong>and</strong> D. L. Kohlstedt (1984), Analysis <strong>of</strong> dislocations in somenaturally deformed plagioclase feldspars, Phys. Earth Planet. Inter., 11,153–160.Passchier, C. W., <strong>and</strong> R. A. J. Trouw (1996), Microtectonics, 325pp.,Springer Verlag, Berlin.Pharr, G. M., P. S. Godavarti, <strong>and</strong> B. L. Va<strong>and</strong>rager (1989), Effects <strong>of</strong>wetting on the compression creep behaviour <strong>of</strong> metals containing lowmelting intergranular phases, J. Mater. Sci., 24, 784–792.Poirier, J. P. (1985), Creep <strong>of</strong> Crystals, 260 pp., Cambridge Univ. Press,Cambridge, U.K.Powell, R., <strong>and</strong> T. J.B. Holl<strong>and</strong> (1999), Relating formulations <strong>of</strong> the thermodynamics<strong>of</strong> mineral solid solutions; activity modeling <strong>of</strong> pyroxenes,amphiboles <strong>and</strong> micas, Am. Mineral., 84, 1–14.Price, N. J., <strong>and</strong> J. W. Cosgrove (1990), Analysis <strong>of</strong> Geological Structures,502 pp, Cambridge Univ. Press, Cambridge, U.K.Ramsay, J. G., <strong>and</strong> M. I. Huber (1983), The Techniques <strong>of</strong> Modern StructuralGeology, vol. 1, 307 pp., Academic, London.Ranalli, G. (1995), Rheology <strong>of</strong> the Earth, 2nd ed., 413 pp., Chapmann <strong>and</strong>Hall, New York.Renner, J., B. Evans, <strong>and</strong> G. Hirth (2000), On the rheologically critical meltfraction, Earth Planet. Sci. Lett., 181(4), 585–594.Rosenberg, C. L. (2001), Deformation <strong>of</strong> partially molten granite: A review<strong>and</strong> comparison <strong>of</strong> experimental <strong>and</strong> natural case studies, Int. J. EarthSci., 90(1), 60–76.Rosenberg, C. L., <strong>and</strong> M. R. H<strong>and</strong>y (2000), Syntectonic melt pathways duringsimple shearing <strong>of</strong> a partially molten rock analogue (Norcamphor-Benzamide), J. Geophys. Res., 105, 3135– 3149.Rosenberg, C. L., <strong>and</strong> M. R. H<strong>and</strong>y (2005), Experimental deformation<strong>of</strong> partially melted granite revisited: implications for the continental crust,J. Metamorph. Geol., 23(1), 19–28.Rosenberg, C. L., <strong>and</strong> U. Riller (2000), Partial-melt topology in statically<strong>and</strong> dynamically recrystallized granite, Geology, 28(1), 7–10.Rushmer, T. (1996), Melt segregation in the lower crust: How have experimentshelped us?, Trans. R. Soc. Edinburgh Earth Sci., 87(1–2),73–83.Rushmer, T. (2001), Volume change during partial melting reactions: implicationsfor melt extraction, melt geochemistry <strong>and</strong> crustal rheology,Tectonophysics, 342(3–4), 389–405.Rutter, E. H., <strong>and</strong> D. H. K. Neumann (1995), Experimental deformation <strong>of</strong>partially molten Westerly granite under fluid-absent conditions, with implicationsfor the extraction <strong>of</strong> granitic magmas, J. Geophys. Res., 100,15,697–15,715.Rybacki, E., <strong>and</strong> G. Dresen (2004), Deformation mechanism maps forfeldspar rocks, Tectonophysics, 382(3–4), 173–187.Sawyer, E. W. (2001), Melt segregation in the continental crust: Distribution<strong>and</strong> movement <strong>of</strong> melt in anatectic rocks, J. Metamorph. Geol.,19(3), 291–309.Schmid, S. M., R. Heilbronner, <strong>and</strong> H. Stünitz (1999), Deformation mechanismsin nature <strong>and</strong> experiment, Tectonophysics, 303(1–4), vii –ix.Schulmann, K., R. Melka, <strong>and</strong> B. Mlčoch (1996), High-temperature microstructures<strong>and</strong> rheology <strong>of</strong> deformed granite, Erzgebirge, Bohemian Massif,J. Struct. Geol., 18, 719–733.Shelton, G., <strong>and</strong> J. Tullis (1981), Experimental flow laws for crustal rocks,Eos Trans. AGU, 62, 396.Sklenička, V., I. Saxl, <strong>and</strong> J. Čadek (1977), Mezikrystalový lom při VysokoteplotnímCreepu Kovù a Slitin (Intercrystalline Fracturing at HighTemperature Creep <strong>of</strong> Metals <strong>and</strong> Alloys), 108 pp., ČSAV, Prague.Smith, J. V. (1974), Feldspar Minerals. 1. Crystal Structure <strong>and</strong> PhysicalProperties, 627 pp., Springer, Berlin.Spry, A. (1969), Metamorphic Textures, 350 pp., Pergamon, Oxford, U.K.Thompson, J. B., <strong>and</strong> G. L. Hovis (1979), Entropy <strong>of</strong> mixing in sanidine,Am. Mineral., 64, 57–65.Tullis, J. (1983), Deformation <strong>of</strong> feldspars, in Feldspar Mineralogy, Rev.Mineral., vol. 2, edited by P. H. Ribbe, pp. 297–323, Mineral. Soc. <strong>of</strong>Am., Washington, D. C.Tullis, J., <strong>and</strong> R. A. Yund (1991), Diffusion creep in feldspar aggregates—Experimental-evidence, J. Struct. Geol., 13, 987–1000.Ulrich, S., K. Schulmann, <strong>and</strong> M. Casey (2002), Micro<strong>structural</strong> evolution<strong>and</strong> rheological behaviour <strong>of</strong> marbles deformed at different crustal levels,J. Struct. Geol., 24, 979–995.van der Molen, I., <strong>and</strong> M. S. Paterson (1979), Experimental deformation <strong>of</strong>partially-melted granite, Contrib. Mineral. Petrol., 70, 299–318.Vigneresse, J. L., P. Barbey, <strong>and</strong> M. Cuney (1996), Rheological transitionsduring partial melting <strong>and</strong> crystallization with application to felsicmagma segregation <strong>and</strong> transfer, J. Petrol., 37(6), 1579–1600.Vollbrecht, A., M. Stipp, <strong>and</strong> N. Ø. Olesen (1999), Crystallographic orientation<strong>of</strong> microcracks in quartz <strong>and</strong> inferred deformation processes: Astudy on gneisses from the German Continental Deep Drilling Project(KTB), Tectonophysics, 303(1–4), 279–297.Wadsworth, J., O. A. Ruano, <strong>and</strong> O. D. Sherby (1999), Deformation bygrain boundary sliding <strong>and</strong> slip creep versus diffusional creep, paperpresented at the Minerals, Metal <strong>and</strong> Materials Society Annual Meeting,San Diego, Calif.Walte, N. P., P. D. Bons, <strong>and</strong> C. W. Passchier (2005), Deformation <strong>of</strong> meltbearingsystem-insight from in situ grain-scale analogue experiments,J. Struct. Geol., 27, 1666–1679.Wei, Y. H., Q. D. Wang, Y. P. Zhu, H. T. Zhou, W. J. Ding, Y. Chino, <strong>and</strong>M. Mabuchi (2003), Superplasticity <strong>and</strong> grain boundary sliding in rolledAZ91 magnesium alloy at high strain rates, Mater. Sci. Eng. A, 360, 107–115.White, R. W., R. Powell, <strong>and</strong> T. J. B. Holl<strong>and</strong> (2001), Calculation <strong>of</strong> partialmelting equilibria in the system Na 2 O-CaO-K 2 O-FeO-MgO-Al 2 O 3 -SiO 2 -H 2 O (NCKFMASH), J. Metamorph. Geol., 19(2), 139–153.Willaime, C., <strong>and</strong> M. G<strong>and</strong>ais (1977), Electron microscope study <strong>of</strong> plasticdefects in experimentally deformed alkali feldspars, Bull. Soc. Fr.Mineral. Cristall., 100(5), 263–271.Willaime, C., J. M. Christie, <strong>and</strong> D. Mainprice (1979), Experimental deformation<strong>of</strong> K-feldspar single crystals, Bull. Mineral., 102, 168–177.Zhang, Y., B. E. Hobbs, <strong>and</strong> M. W. Jessell (1994), The effect <strong>of</strong> grainboundarysliding on fabric development in polycrystalline aggregates,J. Struct. Geol., 16, 1315–1325.Zulauf, G., W. Dörr, J. Fiala, J. Kotková, H. Maluski, <strong>and</strong> P. Valverde-Vaquero (2002), Evidence for high-temperature diffusional creep preservedby rapid cooling <strong>of</strong> lower crust (North Bohemian shear zone,Czech Republic), Terra Nova, 14(5), 343–354.J. Konopásek, Czech Geological Survey, Prague, Czech Republic.(konopasek@cgu.cz)O. Lexa <strong>and</strong> K. Schulmann, Centre de Geochimie de la Surface, UMR7516, Université Louis Pasteur, Strasbourg, 67000 Cedex, France.(lexa@illite.u-strasbg.fr; schulman@illite.u-strasbg.fr)S. Ulrich, Geophysical Institute, Czech Academy <strong>of</strong> Sciences, Prague,Czech Republic. (stano@ig.cas.cz)P. Závada, Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, CharlesUniversity, Prague, Czech Republic. (zavada@natur.cuni.cz)15 <strong>of</strong> 15293


ClickHereforFullArticleJOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113, B10406, doi:10.1029/2007JB005508, 2008Evolution <strong>of</strong> microstructure <strong>and</strong> melt topology in partially moltengranitic mylonite: Implications for rheology <strong>of</strong> felsic middle crustKarel Schulmann, 1 Jean-Emmanuel Martelat, 2 Stanislav Ulrich, 3,4 Ondrej Lexa, 4Pavla Štípská, 1 <strong>and</strong> Jens K. Becker 5Received 19 November 2007; revised 27 April 2008; accepted 30 July 2008; published 24 October 2008.[1] The deformation study <strong>of</strong> midcrustal porphyritic granite reveals exceptionally highstrain intensities <strong>of</strong> feldspar aggregates compared to stronger quartz. Three types <strong>of</strong>microstructures corresponding to evolutionary stages <strong>of</strong> deformed granite wererecognized: (1) the metagranite marked by viscous flow <strong>of</strong> plagioclase around strongalkali feldspar <strong>and</strong> quartz, (2) quartz augen orthogneiss characterized by development <strong>of</strong>b<strong>and</strong>ed mylonitic structure <strong>of</strong> recrystallized plagioclase <strong>and</strong> K-feldspar surroundingaugens <strong>of</strong> quartz, <strong>and</strong> (3) b<strong>and</strong>ed mylonite characterized by alternation <strong>of</strong> quartz ribbons<strong>and</strong> mixed aggregates <strong>of</strong> feldspars <strong>and</strong> quartz. The original weakening <strong>of</strong> alkali feldspar isachieved by decomposition into albite chains <strong>and</strong> K-feldspar resulting from aheterogeneous nucleation process. The subsequent collapse <strong>of</strong> alkaline feldspar <strong>and</strong>development <strong>of</strong> monomineralic layering is attributed to the onset <strong>of</strong> syn-deformationaldehydration melting <strong>of</strong> Mu-Bi layers associated with production <strong>of</strong> 2% melt. The finaldeformation stage is marked by mixing <strong>of</strong> feldspars which is explained by higher meltproduction due to introduction <strong>of</strong> external water. An already small amount <strong>of</strong> melt isresponsible for extreme weakening <strong>of</strong> the feldspar because <strong>of</strong> Melt ConnectivityThreshold effect triggering grain boundary sliding deformation mechanisms. The grainboundary sliding controls diffusion creep at small melt fraction <strong>and</strong> evolves to particulateflow at high melt fractions. Strong quartz shows a dislocation creep deformationmechanism throughout the whole deformation history marked by variations in the activity<strong>of</strong> the slip systems, which are attributed to variations in stress <strong>and</strong> strain rate partitioningwith regard to changing rheological properties <strong>of</strong> the deforming feldspars.Citation: Schulmann, K., J.-E. Martelat, S. Ulrich, O. Lexa, P. Štípská, <strong>and</strong> J. K. Becker (2008), Evolution <strong>of</strong> microstructure <strong>and</strong> melttopology in partially molten granitic mylonite: Implications for rheology <strong>of</strong> felsic middle crust, J. Geophys. Res., 113, B10406,doi:10.1029/2007JB005508.1. Introduction[2] Rheology <strong>of</strong> the continental crust is dominated byquartzo-feldspathic rocks, which are represented mainly bymetagranitoids, orthogneisses <strong>and</strong> felsic volcanics [Carter<strong>and</strong> Tsenn, 1987]. To date, the models <strong>of</strong> crustal rheologyuse laboratory derived laws described by constitutive equationsthat are established for minerals or monomineralicrocks such as quartzites <strong>and</strong> anorthosites [Ranalli, 1995].Most <strong>of</strong> laboratory experiments show that the quartz is1 Centre de Géochimie de la Surface, UMR7516, Université LouisPasteur, CNRS, Strasbourg, France.2 Laboratoire de Géodynamique des Chaînes Alpines, UMR5025,Université Joseph Fourier, Observatoire des Sciences de l’Univers deGrenoble, CNRS, Grenoble, France.3 Geophysical Institute, Czech Academy <strong>of</strong> Sciences, Prague, CzechRepublic.4 Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University,Prague, Czech Republic.5 Institut für Geowissenschaften, Universität Tübingen, Tübingen,Germany.Copyright 2008 by the American Geophysical Union.0148-0227/08/2007JB005508$09.00weaker than plagioclase for the same homologous temperatures[Ranalli <strong>and</strong> Murphy, 1987; Schmid, 1982]. However,the natural quartzo-feldspathic rocks are mixtures withdifferent proportions <strong>of</strong> strong feldspars <strong>and</strong> weak quartzwith variable grain shapes <strong>and</strong> grain size distributions. Thedeformation <strong>of</strong> such natural rocks leads to strain partitioningbetween the different components <strong>and</strong> nonuniform deformation[H<strong>and</strong>y, 1990]. H<strong>and</strong>y [1994a] defined the loadbearingframework structure <strong>and</strong> interconnected weak layerstructure <strong>and</strong> proposed comprehensive empirical equationsthat determine the strength <strong>of</strong> polyphase composites. H<strong>and</strong>y[1994a] applied this concept to quartzo-feldspathic rocks<strong>and</strong> concluded that the proportion <strong>of</strong> weaker quartz controlsthe bulk rheology. The basis <strong>of</strong> these models is the coexistence<strong>of</strong> two nonlinear viscous phases; the bulk rheologyis a consequence <strong>of</strong> the rock structure <strong>and</strong> the relativeproportions <strong>of</strong> the two mineral phases [Ji <strong>and</strong> Zhao, 1994].[3] Micro<strong>structural</strong> studies show that the progressiveorthogneiss deformation is associated with strain partitioning<strong>and</strong> variations in the deformation mechanisms <strong>of</strong>feldspars <strong>and</strong> quartz [H<strong>and</strong>y et al., 1999; Schulmann etal., 1996; Simpson, 1985]. For instance Gapais [1989] <strong>and</strong>B104062951<strong>of</strong>20


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 1. Geological map <strong>of</strong> the central part <strong>of</strong> Bohemian Massif modified after Beneš [1964] <strong>and</strong>Synek <strong>and</strong> Oliverová [1993]. Black stars <strong>and</strong> capitalized letters refer to studied samples (B, D, H, M, R, S,T, <strong>and</strong> V). (a) Lower crustal rocks (granulites <strong>and</strong> eclogites), (b) the midcrustal monotonousmetasedimentary unit, (c) the midcrustal orthogneiss unit, (d) the kyanite micaschist unit, (e) theintrusives, (f) Cretaceous <strong>and</strong> Quaternary rocks, (g) the undifferenciated Lower Palaeozoic rocks, <strong>and</strong>(h) the Neo-Proterozoic metasediments.Schulmann et al. [1996] have shown that at amphibolitefacies the feldspars show evolution from dislocation creepto grain boundary sliding with increasing strain intensity inconjunction with variations in the activity <strong>of</strong> the quartz slipsystems. Therefore, the viscosity contrast between feldspars<strong>and</strong> quartz was not constant but varied with increasingdegree <strong>of</strong> strain as the rheological role <strong>of</strong> individualminerals evolved as shown by a range <strong>of</strong> experimental<strong>and</strong> natural studies [H<strong>and</strong>y, 1994b; Ji et al., 2000; Rybacki<strong>and</strong> Dresen, 2004; Stünitz <strong>and</strong> Fitz Gerald, 1993]. However,the aforementioned studies all neglect a possible role<strong>of</strong> the melt on the deformation <strong>of</strong> the polyphase rocks.[4] The aim <strong>of</strong> this paper is to show, through detailedmicro<strong>structural</strong> study <strong>and</strong> thermodynamical modeling, thecontribution <strong>of</strong> interstitial melt to the rheology <strong>of</strong> progressivelydeformed granites under midcrustal conditions. Weuse natural examples <strong>of</strong> a sequence <strong>of</strong> granite mylonites todocument the melt enhanced rheological inversion <strong>of</strong> disproportionatelystronger quartz compared to the weakfeldspars in midcrustal rocks. This work also shows thatwith increasing melt fraction the deformation mechanisms<strong>of</strong> feldspars varies from the grain boundary sliding accommodateddislocation creep to granular flow while the quartzis only deformed in the dislocation creep field.2. Geological Setting[5] The study area located in the central part <strong>of</strong> theBohemian Massif in the Czech Republic is known for theextreme deformation <strong>of</strong> porphyritic granites in a crustalscaleshear zone [Synek <strong>and</strong> Oliverová, 1993]. The porphyriticgranite mylonites studied are <strong>of</strong> Cambro-Ordovicianprotolith age <strong>and</strong> come from an orthogneiss-bearing, midcrustalunit that overlies kyanite micaschists in the west(Figure 1). In the east a similar rock assemblage occurs inan equivalent <strong>structural</strong> unit, although these orthogneissbodies exist within kyanite-sillimanite bearing micaschists.According to Schulmann et al. [2005] both units record aCarboniferous tectono-metamorphic history between 340<strong>and</strong> 325 Ma. The midcrustal unit is overlain by kyanitebearing migmatites <strong>and</strong> granulites <strong>of</strong> the orogenic lowercrustal unit that contain eclogite lenses with estimatedminimum pressures <strong>of</strong> 18–19 kbar <strong>and</strong> temperatures <strong>of</strong>800–900°C [Medaris et al., 1998]. The P-T conditions <strong>of</strong>the kyanite bearing micaschists in the footwall <strong>of</strong> themylonitic orthogneiss sheet have experienced temperatures<strong>of</strong> 620–710°C at pressures <strong>of</strong> 6.5–9.5 kbar [Kachlík, 1999]similarly to 8–9 kbar <strong>and</strong> 610–660°C estimated in themicaschists in the eastern part <strong>of</strong> the studied area [Pitra <strong>and</strong>Guiraud, 1996].[6] The geological structure <strong>of</strong> the studied rocks wasdescribed by Synek <strong>and</strong> Oliverová [1993] who interpretedthe middle crustal orthogneiss-bearing unit <strong>and</strong> the overlyingorogenic lower crustal unit as a crustal nappe stackresulting from a Carboniferous deformation. On the basis <strong>of</strong>the <strong>structural</strong> position <strong>of</strong> the midcrustal orthogneiss <strong>and</strong> theregional metamorphic field gradient, the PT conditions <strong>of</strong>metamorphism <strong>and</strong> deformation <strong>of</strong> the studied rocks areestimated to be between 9 <strong>and</strong> 18 kbar <strong>and</strong> 650–850°C.More precise P-T estimations were not established because<strong>of</strong> the lack <strong>of</strong> pressure <strong>and</strong> temperature sensitive mineralassemblages that are necessary for st<strong>and</strong>ard thermobarometricmethods.3. Shape Analysis <strong>of</strong> Quartz <strong>and</strong> Feldspars[7] Undeformed porphyritic granitoids may serve as anexcellent example <strong>of</strong> multiphase mixtures <strong>of</strong> originallyspherical (ellipsoidal) clasts with constant phase fractions.When these rocks are subjected to deformation, mineralgrains reach different strain intensities, which can be easilyquantified using st<strong>and</strong>ard finite strain techniques [Ramsay<strong>and</strong> Huber, 1983]. Measurement <strong>of</strong> the shapes <strong>of</strong> naturallydeformed minerals in originally coarse-grained <strong>and</strong> porphyriticgranitoids may thus help to track the viscous behavior<strong>of</strong> individual phases for different bulk strain intensities2<strong>of</strong>20296


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 2. Macroscopic samples <strong>of</strong> deformed metagranite divided in three types according to thedeformation intensity <strong>and</strong> the macroscopic appearance. X, Y, <strong>and</strong> Z refer to the axes <strong>of</strong> the finite strainellipsoid. (a) Sample S1 is a weakly deformed metagranite with large centimeter-sized grains <strong>of</strong> quartz<strong>and</strong> feldspar representing Type I rock. (b, c) Sample M1 is an augen orthogneiss corresponding to Type IIrocks <strong>and</strong> intermediate strain intensity (equivalent to samples T1, T2, M2, <strong>and</strong> R4), (d, e) Sample V1 is ab<strong>and</strong>ed mylonite corresponding to Type III rock <strong>and</strong> the highest intensity <strong>of</strong> deformation (equivalent toother highly strained samples R5, R3, V2, <strong>and</strong> H1).[Treagus, 2002]. This allows constraining the degree <strong>of</strong>strain partitioning in rocks <strong>and</strong> the viscosity ratios betweenthe individual mineral phases.[8] The shape analysis <strong>of</strong> the feldspars <strong>and</strong> the quartzpolycrystalline aggregates was carried out on 17 sectionscut both perpendicular to the foliation <strong>and</strong> parallel to thestretching lineation <strong>and</strong> perpendicular to both the foliation<strong>and</strong> the lineation, i.e., parallel to the XZ <strong>and</strong> YZ sections <strong>of</strong>the finite strain ellipsoid (Figure 2). The K-feldspar cannotbe distinguished from plagioclase in highly deformed macroscopicsamples <strong>and</strong> therefore both minerals were groupedtogether for finite strain measurements. All studied samplesare composed on average <strong>of</strong> 60–70% feldspars, 35–25%quartz <strong>and</strong> up to 10% <strong>of</strong> biotite <strong>and</strong> muscovite. The almostconstant mineral composition for highly variable strainintensities <strong>and</strong> constant bulk rock chemistry shows the lack<strong>of</strong> chemical variations with strain (Table 1).[9] In our study we classified three major types <strong>of</strong>deformed orthogneiss according to the deformation intensitiesat the macroscopic scale: Type I is represented byweakly deformed metagranite (Figure 2a); Type II correspondsto augen orthogneiss with quartz porphyroclast(Figures 2b <strong>and</strong> 2c); Type III is a b<strong>and</strong>ed mylonite orthogneiss(Figures 2d <strong>and</strong> 2e).[10] Mineral shape data are plotted into a Flinn diagram[Flinn, 1965] (Figure 3). The <strong>analyses</strong> <strong>of</strong> feldspar polycrystallineaggregates <strong>and</strong> quartz <strong>of</strong> the individual samples areconnected by tie lines with the vertical ellipse representingthe bulk strain value <strong>of</strong> the whole rock. An importantfeature <strong>of</strong> all the studied samples is that feldspars showhigher strain intensities than quartz for any bulk strain(Figures 3a <strong>and</strong> 3b). In several samples <strong>of</strong> the so-calledDoubravčany pencil gneiss, the strain intensities cannot beidentified because the stretching <strong>of</strong> feldspar <strong>and</strong> quartzlayers exceeds the length <strong>of</strong> the samples (Figure 2d). Thestrain symmetry, represented by the K values <strong>of</strong> Flinn[1965], varies from prolate to oblate shapes (K = 2.7 to0.3). For quartz with prolate shapes, the correspondingfeldspar strain symmetry is close to plane strain, while foroblate quartz, the feldspars show the same shape or slightlymore oblate shapes (Figure 3b). Strain intensities areexpressed using Ramsay’s D value, which is an alternativeexpression <strong>of</strong> viscosity ratio values <strong>of</strong> Gay [1968a, 1968b] alsoused by Schulmann et al. [1996]. The inspection <strong>of</strong> the diagramin Figure 3b shows that the highest strain ratios are achievedbetween D fel /D qtz for Type III orthogneiss <strong>and</strong> some Type IIorthogneiss samples marked by high bulk strain intensities. Therelatively small ratio between D fel /D qtz suggests similar yield-3<strong>of</strong>20297


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Table 1. Bulk Rock Chemistry <strong>of</strong> Midcrustal OrthogneissesSample S1 M1 V1 D1SiO 2 72, 9 70, 9 73, 0 70, 2TiO 2 0, 3 0, 3 0, 2 0, 4Al 2 O 3 13, 3 14, 7 13, 7 14, 7Fe 2 O 3 1, 9 1, 4 1, 8 2, 6FeO 0, 9 1, 0 0, 7 0, 7MnO 0,0 0,1 0,0 0,0MgO 0,5 0,6 0,4 0,8CaO 1, 1 1, 4 0, 8 1, 0Na 2 O 2,3 2,6 2,5 1,8K 2 O 4,5 4,7 4,7 4,6P 2 O 5 0, 2 0, 2 0, 2 0, 2H 2 O- 0, 2 0, 3 0, 3 0, 2H 2 O + 1, 0 1, 2 1, 0 2, 1CO 2 0, 2 0, 2 0, 4 0, 2Total 99, 5 99, 5 99, 7 99, 6ing <strong>of</strong> both mineral phases for low bulk strains for severalsamples <strong>of</strong> weakly deformed Type II orthogneiss.4. Microstructure Development <strong>of</strong> DeformedMetagranitoids[11] The micro<strong>structural</strong> investigations covered the qualitativedescription <strong>of</strong> the rock <strong>and</strong> mineral structure usingoptical microscope <strong>and</strong> scanning electron microscope imaging.The microprobe work complements the scanning electronmicroscope (SEM) study to identify compositional variations<strong>of</strong> the recrystallized feldspars (Tables 2a, 2b, <strong>and</strong> 2c).4.1. Type I: Metagranite <strong>and</strong> Weakly DeformedOrthogneiss[12] The metagranite samples show K-feldspar phenocrystsup to 10 cm in size surrounded by quartz blebs.Elongated plagioclase polycrystalline aggregates range from1 to 3 cm in size (Figure 2a). Light <strong>and</strong> SEM microscopiesshow that the plagioclase forming recrystallized aggregates(labeled Pl1 in Figure 4) corresponds to oligoclase An 15 .The Pl1 plagioclase has an average grain size <strong>of</strong> 5070 mm, straight boundaries <strong>and</strong> subequant shapes (Figure 4<strong>and</strong> Table 2c). The quartz blebs consist <strong>of</strong> large grains (400mm in size) with irregular <strong>and</strong> highly serrated boundaries.Biotite <strong>and</strong> muscovite are present as elongated recrystallizedaggregates. The plagioclase grains within alkaline feldsparphenocrysts are labeled Pl1b here. These plagioclases are 10to 50 mm in size, correspond to albite An 1–8 <strong>and</strong> are usuallyarranged into chains mostly parallel to (100) or locallyalong (010) planes <strong>of</strong> the host K-feldspar (Figures 4a <strong>and</strong>4b <strong>and</strong> Table 2c). The K-feldspars themselves show subgrains<strong>of</strong> 80 mm in size. Subgrain boundaries are straight<strong>and</strong> meet in triple point junctions. At a smaller scale, thefeldspars show kinked domains along which the Pl1b chainsare rotated (Figure 4a). These kinked domains show internalstrain marked by irregular undulatory extinction. Microprobe<strong>analyses</strong> show that K-feldspar cores exhibit ratherconstant composition revealing approximately 10 mol % <strong>of</strong>albite (Table 2a). Only around enclosed Pl1b chains thealkali feldspar shows a rapid decrease <strong>of</strong> the albite component(Figure 5).4.2. Type II: Orthogneiss With Quartz Augens[13] This rock is characterized by the presence <strong>of</strong> isolated<strong>and</strong> elongated quartz augens surrounded by highly elongatedK-feldspar <strong>and</strong> plagioclase polycrystalline aggregates formingalmost monomineralic layers (Figure 6). Biotite <strong>and</strong>muscovite are forming elongated monomineralic aggregateslocated at boundaries between feldspars <strong>and</strong> quartz orwithin plagioclase layers. Quartz augens are composed <strong>of</strong>grains 150–1000 mm in size with serrated boundaries.Plagioclase (Pl1) polycrystalline aggregates are composed<strong>of</strong> oligoclase An 10 – 20 subequant grains (50–150 mm inFigure 3. (a) Flinn diagram after Ramsay <strong>and</strong> Huber [1983] shows the shapes <strong>of</strong> deformation ellipsoidsrepresented as projections <strong>of</strong> X/Y axial ratios on the ordinate <strong>and</strong> Y/Z ratios on the abscissa in the graph.This diagram allows to visualize the shapes <strong>of</strong> the strain ellipsoid <strong>and</strong> the intensity <strong>of</strong> deformation in the2-D diagram. (b) Diagram showing strain intensity expressed as the D parameter on the abscissa againstshape K parameter on the ordinate; K = (X/Y-1)/(Y/Z-1) varies from 0 for oblate shapes, to 1 for planestrain shapes, <strong>and</strong> to infinity for prolate shapes, D = (((X/Y-1) 2 + ((Y/Z-1) 2 ) 1/2 . The values are obtainedby measuring 30 ellipses (harmonic sum) from XZ <strong>and</strong> YZ sections (principal planes <strong>of</strong> finite strainellipsoid). Squares show quartz, <strong>and</strong> circles show undifferentiated feldspar. Rock types are determinedaccording to the bulk strain intensity <strong>and</strong> macroscopic appearance (Figure 2) from weakly deformedmetagranite S1; intermediate strained augen orthogneiss samples T1, T2, M2, M1, <strong>and</strong> R4; <strong>and</strong> highlystrained b<strong>and</strong>ed orthogneiss samples H1, T5, V1, <strong>and</strong> V2.4<strong>of</strong>20298


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Table 2a. K-Feldspar Compositions aSample S1 S1 M1 M1 V1 V1Analysis 50 48 71 79 92 104SiO 2 65, 27 64, 07 64, 09 65, 05 64, 63 64, 96Al 2 O 3 18, 69 18, 42 18, 22 18, 56 18, 73 18, 71CaO 0, 00 0, 03 0, 04 0, 00 0, 03 0, 04Na 2 O 1,20 0,49 0,29 1,37 1,10 0,59K 2 O 15, 87 16, 46 16, 86 15, 52 15, 90 16, 48Total 101, 04 99, 49 99, 50 100, 54 100, 40 100, 82Si 2, 986 2, 985 2, 990 2, 988 2, 978 2, 984Al 1, 008 1, 011 1, 002 1, 005 1, 017 1, 013Ca 0, 000 0, 002 0, 002 0, 000 0, 002 0, 002Na 0, 106 0, 044 0, 026 0, 122 0, 098 0, 053K 0, 926 0, 978 1, 004 0, 910 0, 935 0, 966Total 5, 026 5, 020 5, 024 5, 024 5, 030 5, 017X Or 89, 69 95, 53 97, 27 88, 17 90, 36 94, 66X Ab 10, 31 4, 32 2, 54 11, 83 9, 50 5, 15X An 0, 00 0, 15 0, 19 0, 00 0, 14 0, 19a Structural formulae calculated on the basis <strong>of</strong> 8(O).size) with straight boundaries meeting at triple point junctions(Figures 5 <strong>and</strong> 6c <strong>and</strong> Table 2c). Plagioclase polycrystallineaggregates <strong>of</strong>ten contain biotite flakes parallel tothe foliation or interstitial quartz <strong>and</strong> new plagioclaseAn 1–10 (Pl2). K-feldspar polycrystalline aggregates arecomposed <strong>of</strong> slightly elongate to subequant grains withstraight boundaries forming a well developed triple pointnetwork lined by narrow films <strong>of</strong> pure albite An 1–10 (Pl2).Disintegration <strong>of</strong> the K-feldspar layers into elongate aggregatessurrounded by fine-grained Pl1 matrix was observedin some samples. This process is connected to the development<strong>of</strong> b<strong>and</strong>s filled with recrystallized Pl1 grains up to150 mm wide oblique with respect to the long axis <strong>of</strong> thefeldspar polycrystalline aggregates (Figure 6b). Finally,elongated thin aggregates <strong>of</strong> Pl1 are smeared out in theK-feldspar rich matrix. Quartz also forms weakly elongatedpolymineralic aggregates characterized by highlyirregular boundaries with the surrounding feldspar matrix.Common feature are the lobes <strong>of</strong> quartz <strong>and</strong> cusps <strong>of</strong> theK-feldspar pointing in the direction perpendicular to thelong face <strong>of</strong> the aggregate <strong>and</strong> to the macroscopic foliation.4.3. Type III: B<strong>and</strong>ed Mylonitic Orthogneiss <strong>and</strong>Ultramylonite[14] B<strong>and</strong>ed mylonitic orthogneiss is marked by thinquartz ribbons less than 1000 mm wide. They are surroundedby polymineralic layers <strong>of</strong> plagioclase, quartz <strong>and</strong>Table 2b. Isolated Plagioclase Pl1b <strong>and</strong> Pl2 Compositions aSample S1 T1 T1 M1 V1Analysis 52 119 115 77 105SiO 2 68, 72 69, 06 67, 94 68, 67 67, 69Al 2 O 3 19, 70 19, 74 20, 19 19, 78 20, 69CaO 0, 31 0, 10 0, 78 0, 22 1, 28Na 2 O 11, 84 12, 14 11, 27 11, 77 11, 19K 2 O 0,05 0,12 0,12 0,54 0,23Total 100, 62 101, 19 100, 30 101, 01 101, 09Si 2, 986 2, 986 2, 963 2, 980 2, 938Al 1, 009 1, 006 1, 038 1, 012 1, 058Ca 0, 014 0, 005 0, 037 0, 010 0, 060Na 0, 998 1, 018 0, 953 0, 991 0, 942K 0, 003 0, 007 0, 007 0, 030 0, 013Total 5, 010 5, 022 4, 998 5, 023 5, 010X Or 0, 28 0, 64 0, 67 2, 90 1, 26X Ab 98, 30 98, 91 95, 67 96, 11 92, 87X An 1, 42 0, 45 3, 66 0, 99 5, 87a Structural formulae calculated on the basis <strong>of</strong> 8(O).Table 2c. Large or Interconnected Plagioclase: Pl1 Compositions aSample S1 T1 M1 M1 V1Analysis 43 114 67 91 109SiO 2 64, 60 64, 70 65, 62 63, 75 65, 55Al 2 O 3 22, 72 22, 84 21, 43 23, 06 22, 22CaO 3, 48 3, 76 2, 36 4, 38 3, 07Na 2 O 9, 88 10, 03 11, 03 9, 48 10, 23K 2 O 0,23 0,13 0,10 0,27 0,22Total 100, 91 101, 46 100, 56 100, 97 101, 30Si 2, 827 2, 819 2, 878 2, 797 2, 854Al 1, 172 1, 173 1, 108 1, 192 1, 140Ca 0, 163 0, 176 0, 111 0, 206 0, 143Na 0, 838 0, 847 0, 938 0, 806 0, 864K 0, 013 0, 007 0, 006 0, 015 0, 012Total 5, 013 5, 022 5, 040 5, 017 5, 013X Or 1, 26 0, 70 0, 53 1, 47 1, 20X Ab 82, 65 82, 26 88, 95 78, 49 84, 75X An 16, 09 17, 04 10, 52 20, 04 14, 05a Structural formulae calculated on the basis <strong>of</strong> 8(O).K-feldspar (Figure 7b). Micas are generally dispersed in thematrix or form narrow layers parallel to the foliation(Figures 7c <strong>and</strong> 7d). Disintegrated relics <strong>of</strong> plagioclaseaggregates are composed <strong>of</strong> Pl1 grains An 14 – 20 (Figure 5<strong>and</strong> Table 2c) <strong>and</strong> <strong>of</strong> K-feldspar grains 50–100 mm in size(rarely 200 mm) <strong>and</strong> <strong>of</strong> irregular shapes. The interfacialboundaries with K-feldspar are generally straight <strong>and</strong> commonlylined by rims <strong>of</strong> Pl2 (An 4–12 ) or thin layers <strong>of</strong> quartz.Locally, relics <strong>of</strong> the K-feldspar layers that are a few grainswide (250 mm) occur with straight mutual boundaries linedby Pl2 films. The quartz commonly occurs in the form <strong>of</strong>isolated grains with serrated boundaries <strong>and</strong> cusps pointingperpendicular <strong>and</strong> parallel to the macroscopic foliation. However,the most common are few hundred microns wide quartzribbons composed by elongate, 200–300 mm wide <strong>and</strong> up to1000 mm long grains with highly serrated boundaries.[15] A rather exceptional kind <strong>of</strong> Type III microstructureis represented by the exceptionally coarse-grained, myloniticorthogneiss sample B2. This rock type is characterizedby polyminerallic layers with poorly defined boundaries <strong>and</strong>a large average grain size <strong>of</strong> plagioclase <strong>and</strong> K-feldsparranging from 160 to 200 mm. The Pl1 grains in relictplagioclase aggregates show sutured boundaries <strong>and</strong> widePl2 rims while the interstitial Pl2 grains in the K-feldsparrich aggregates form wide films <strong>and</strong> cuspate pools. Theinterstitial quartz (100 mm in size) occurs in the form <strong>of</strong>serrated grains in triple point junctions <strong>of</strong> K-feldspar <strong>and</strong>Pl1 relict aggregates.4.4. Topology <strong>of</strong> Interstitial Phases in Plagioclase <strong>and</strong>K-Feldspar Aggregates[16] The most common interstitial phase at the grain scaleare thin films <strong>of</strong> Pl2 <strong>and</strong> quartz up to 50 mm long <strong>and</strong> 10 mmwide located along the mutual plagioclase boundaries, <strong>of</strong>tenat high angle to the foliation or in the form <strong>of</strong> cuspate poolsat triple point junctions. The mutual boundaries <strong>of</strong> Pl1aggregate grains are <strong>of</strong>ten lined by narrow rims <strong>of</strong> newplagioclase An 1–10 (Pl2). In Type II orthogneiss samplesshowing strongly elongated recrystallized K-feldspar <strong>and</strong>Pl1 grains (samples T1, T2), the Pl2 films exhibit preferredorientation parallel to the direction <strong>of</strong> maximum stretching,i.e., along the crystal faces parallel to the foliation (Figure 8).In the M1, M2 samples <strong>of</strong> the orthogneiss Type II, the shapepreferred orientation <strong>of</strong> the recrystallized feldspar is low,5<strong>of</strong>20299


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406[18] Another important feature is the development <strong>of</strong> intragranular,wedge shaped fractures that are predominantlydeveloped in elongated feldspar grains with stronger SPOin samples T1 <strong>and</strong> T2 (Figure 8). These gently curvedfractures <strong>of</strong>ten originate at the foliation-parallel grain boundariesor triple point junctions, terminate in the interior <strong>of</strong> thegrains <strong>and</strong> are oriented at low angles (15–30°) to themaximum stretching. In Type III orthogneiss samples (e.g.,sample B2) the intragranular fractures occur as well beingoriented at high angle to the principal stretching direction.Figure 4. Type I microstructure (S1 sample). (a) Type Imetagranite marked by large alkaline feldspar clasts decomposedinto albite chains <strong>and</strong> K-feldspar <strong>and</strong> (b) scanningelectron microscope (SEM) image <strong>of</strong> alkaline feldspardecomposition. Rectangle indicates the position <strong>of</strong> the detailedSEM image. (c) The detail <strong>of</strong> the SEM image showing theshapes <strong>of</strong> new bulbous albite grains. (d) CorrespondingArcView Geographical Information Systems (GIS) image usedfor the quantitative miro<strong>structural</strong> analysis. Scanning electronmicroscope images obtained in backscattered electron mode(Camscan microscope, Institute <strong>of</strong> Petrology <strong>and</strong> StructuralGeology Prague). K-feldspar is represented in light gray,biotite is represented as white laths, quartz <strong>and</strong> plagioclase areshown in dark gray, <strong>and</strong> rectangular white mica laths areshown in light gray. In ArcView GIS images the K-feldspar isrepresented in white, white mica <strong>and</strong> biotite are represented inblack, plagioclase is represented in light gray, <strong>and</strong> quartz isrepresented in dark gray. The ArcView GeographicalInformation System was used as an ideal environment fordigitizing mineral shapes [Lexa et al., 2005]. Mineralabbreviations correspond to those <strong>of</strong> Kretz [1983].which corroborates the development <strong>of</strong> thin albite-quartzfilms along the feldspar faces oriented along two maximawith respect to the macroscopic layering (Figure 8). Themean orientation <strong>of</strong> the Pl2 <strong>and</strong> quartz seams form an angle<strong>of</strong> 10–15° with respect to the foliation trace. Interstitialconvex quartz grains <strong>of</strong>ten occur at triple point junctionsbut locally line the feldspar boundaries as narrow films(Figures 6c <strong>and</strong> 6d).[17] The presence <strong>of</strong> interstitial phases is the most pronouncedin Type III orthogneiss (samples R3, V1 <strong>and</strong> B2 inparticular) where the recrystallized feldspar grains exhibitno shape preferred orientation. Here, wider cuspate-lobatepools <strong>and</strong> narrow films <strong>of</strong> Pl2 <strong>and</strong> quartz occur at a highangle to the macroscopic foliation. The B2 sample shows anextreme orientation <strong>of</strong> Pl2 seams which form an angle <strong>of</strong> upto 70° with the foliation trace.5. <strong>Quantitative</strong> Textural Analysis[19] The quantitative analysis <strong>of</strong> grain shapes <strong>and</strong> boundarieswas carried out in an ArcView GIS environment.Examples <strong>of</strong> analyzed samples are shown in Figures 4d,6d, <strong>and</strong> 7d. The statistical parameters involving grains sizedistributions, shape preferred orientation, degree <strong>of</strong> grainelongation <strong>and</strong> grain contact frequencies were performedusing the MATLAB PolyLX toolbox [Lexa et al., 2005] inorder to quantify the above described micro<strong>structural</strong> sequence<strong>of</strong> rock types. The grain size distribution wasevaluated as an important parameter in deformed rocksbecause <strong>of</strong> its sensitivity to stress <strong>and</strong> temperature [Schmidet al., 1999]. The correct determination <strong>of</strong> grain size isessential in polyphase systems, where stress <strong>and</strong> strain ratepartitioning are expected [H<strong>and</strong>y, 1990]. Another importantparameter is the shape <strong>of</strong> recrystallized grains, which isstrongly dependent on the type <strong>of</strong> deformation mechanisms<strong>and</strong> grain growth history [Boullier <strong>and</strong> Guéguen, 1975;Schmid et al., 1987]. Grain contact distribution in the rockyields information about the evolution <strong>of</strong> spatial distribution<strong>of</strong> the different minerals in rocks with deformation ormelting [Lexa et al., 2005; Hasalová etal., 2008]. This isexpressed by the grain contact frequency (GCF) that statisticallyquantifies the deviation <strong>of</strong> the grain contact distributionfrom r<strong>and</strong>om [Kretz, 1969]. The aggregate distribution,marked by the dominant presence <strong>of</strong> like-like contacts,represents one end-member <strong>of</strong> the spatial distribution <strong>of</strong>grain boundaries that is resulting from the solid statedifferentiation process [McLellan, 1983]. The regular distributionis characterized by the predominance <strong>of</strong> unlikecontacts in rock <strong>and</strong> is considered as a second end-memberFigure 5. Ternary diagrams showing the composition <strong>of</strong>feldspars in different samples. Pl1 corresponds to isolatedplagioclase aggregate (see text for explanation).6<strong>of</strong>20300


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406examined for the interstitial <strong>and</strong> aggregate positions separately.The most striking feature is a progressive increase <strong>of</strong>the aggregate plagioclase grain size from 60 mm (medianvalue) in metagranite to 90 mm in augen orthogneiss to110 mm in b<strong>and</strong>ed orthogneiss, <strong>and</strong> to 170 mm forsample B2 (Figure 9a). We also observe a systematicallyhigher grain size <strong>of</strong> aggregate grains compared to theinterstitial plagioclase grains for all rock types. In addition,the median value <strong>of</strong> the aggregate plagioclase grainsis shifted toward the third quartile value, while in theinterstitial grains it is closer to the first quartile value.[22] In order to obtain information about the crystal nucleation<strong>and</strong> growth, the crystal size distribution (CSD) methodwas applied in the studied samples. The theory <strong>of</strong> CSD is awell-established technique in metallurgy, ceramics <strong>and</strong> chemicalengineering to reveal information about nucleation,growth rates <strong>and</strong> growth times <strong>of</strong> crystals [R<strong>and</strong>olph <strong>and</strong>Larson, 1971]. CSD in many metamorphic <strong>and</strong> igneous rocksshow a log linear relationship between the grain size L <strong>and</strong> thepopulation density N according to the equationN ¼ N 0 e L=GtFigure 6. Type II microstructure. (a) Micrograph <strong>of</strong>sample M2 Type II microstructure characterized by largequartz ribbons <strong>and</strong> fine-grained feldspar matrix, (b) SEMimage <strong>of</strong> the M2 sample showing typical monomineralliclayering <strong>and</strong> plagioclase aggregate bridges, <strong>and</strong> (c) detailedSEM image <strong>of</strong> the T1 sample showing K-feldspar <strong>and</strong>plagioclase polycrystalline aggregate microstructure. Thecharacteristic feature is the high elongation <strong>of</strong> feldspargrains lined with Pl2 films, (d) corresponding ArcView GISdrawing <strong>of</strong> Type II microstructure <strong>of</strong> the T1 sample. TheSEM <strong>and</strong> ArcView GIS colors are the same as in Figure 4.where N 0 <strong>and</strong> Gt are constants <strong>and</strong> may be related to thenucleation density <strong>and</strong> growth rate <strong>of</strong> the crystals [Higgins,1998]. This technique was previously successfully appliedto the metamorphic rocks by Cashman <strong>and</strong> Ferry [1988],Hasalová etal.[2008], <strong>and</strong> Lexa et al. [2005]. CSD plots <strong>of</strong><strong>of</strong> the spatial distribution <strong>of</strong> the grain boundaries thatdevelops mostly because <strong>of</strong> solid state annealing [Flinn,1969], mechanical mixing [Kruse <strong>and</strong> Stünitz, 1999] orcrystallization <strong>of</strong> the interstitial melt [Dallain et al., 1999;Hasalová etal., 2008].5.1. Crystal Size Distribution[20] The grain size distributions <strong>of</strong> recrystallized quartz,K-feldspar <strong>and</strong> <strong>of</strong> plagioclase have been determined on thebasis <strong>of</strong> more than 500 measured grains per thin section foreach sample (except <strong>of</strong> about 200 measurements in coarsegrainedsample B2). Because <strong>of</strong> the ubiquitous lognormaldistribution <strong>of</strong> the grain populations, the median value wasconsidered to be the most reliable statistical value <strong>and</strong> thespread <strong>of</strong> the grain size distribution was evaluated as adifference between the third <strong>and</strong> first quartile instead <strong>of</strong> thest<strong>and</strong>ard deviation.[21] The K-feldspar shows a similar grain size distributionfor all rock types <strong>and</strong> is characterized by the mediangrain size ranging from 60 to 180 mm (Figure 9a). Exceptsample B2 with exceptionally large grain size, the grain sizespread does not follow any systematic pattern. However, theinterstitial quartz shows an increasing median grain sizevalue from 20 to 50 mm in the Type II orthogneiss to 80–100 mm in Type III orthogneiss. The plagioclase wasFigure 7. Type III microstructure <strong>of</strong> b<strong>and</strong>ed orthogneiss.(a) Micrograph <strong>of</strong> sample D1 showing layers <strong>of</strong> quartzalternating with mixed layers <strong>of</strong> feldspars <strong>and</strong> quartz, (b)SEM image <strong>of</strong> the sample V1 showing quartz <strong>and</strong> feldsparmixing <strong>and</strong> poor definition <strong>of</strong> layer boundaries, (c) detailedSEM image <strong>of</strong> K-feldspar <strong>and</strong> plagioclase mixing in thematrix, <strong>and</strong> (d) corresponding ArcView GIS drawing. TheSEM <strong>and</strong> ArcView GIS colors are the same as in Figure 4.7<strong>of</strong>20301


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 8. (a) Detail from the backscattered scanning electron image <strong>of</strong> sample T1 (location <strong>of</strong> Figure 8ais shown in SEM image in Figure 6c). The image shows the distribution <strong>of</strong> albite seams along boundaries<strong>of</strong> K-feldspar grains, character <strong>of</strong> compositional zoning <strong>of</strong> the plagioclase in the top left, <strong>and</strong> intragranularfractures filled with albite (I.F.). (b) Rose diagrams show preferred orientation <strong>of</strong> 100 interstitial quartz<strong>and</strong> Pl2 seams in K-feldspar aggregate. The mean direction <strong>and</strong> st<strong>and</strong>ard circular deviation are shown inthe bottom <strong>of</strong> each diagram. The thick horizontal line represents the orientation <strong>of</strong> lineation (X), <strong>and</strong> thethin line shows the orientation <strong>of</strong> the mean direction.all the samples constructed according to the method byPeterson [1996] exhibit linear correlations between thelogarithm <strong>of</strong> the population density (i.e., the number <strong>of</strong>crystals per size per volume) <strong>and</strong> the crystal size (Figure 9b).Applying the theory <strong>of</strong> CSD, such distributions could beparameterized by the zero size intercept N 0 (nucleationdensity) <strong>and</strong> slope Gt (growth rate multiplied by time).These two parameters are plotted in a N 0 -Gt diagram[Lexa et al., 2005] where the samples form a distincttrend. Trends in the grain size distributions using CSDmethod are visualized in Figure 9b. The crystal sizedistribution plot <strong>of</strong> the plagioclase aggregates is the mostpronounced <strong>and</strong> shows a systematic decrease <strong>of</strong> N 0 <strong>and</strong> anincrease <strong>of</strong> the G t values with increasing degree <strong>of</strong>deformation i.e., from Type I to Type III rocks.5.2. Grain Shapes <strong>and</strong> Shape Preferred Orientation(SPO)[23] The aspect ratio median value <strong>of</strong> plagioclase <strong>and</strong>K-feldspar varies between 1.5 to 3.1 <strong>and</strong> no systematicpattern related to the type <strong>of</strong> rocks <strong>and</strong> the degree <strong>of</strong>deformation is obvious (Figure 10). However, the SPO <strong>of</strong>plagioclase <strong>and</strong> K-feldspar in most <strong>of</strong> Type II <strong>and</strong> IIIorthogneiss samples is higher compared to the Type Isample with exceptionally high SPO for samples T1 <strong>and</strong>T2. There is a difference between the aggregate plagioclase(Pl1) <strong>and</strong> the interstitial albite (Pl2) marked by a systematicallyhigher SPO for the former compared to the latter.5.3. Grain Contact Frequencies (GCF) <strong>and</strong> GrainBoundary Preferred Orientation (GBPO)[24] The combination <strong>of</strong> the GCF analysis with thestudies <strong>of</strong> the preferred orientation <strong>of</strong> the like-like (like–like contacts = boundaries <strong>of</strong> minerals <strong>of</strong> the same species)<strong>and</strong> unlike grain boundaries (GBPO) yields importantinformation about the organization <strong>of</strong> the grain boundarieswith respect to the deformation processes [Lexa et al.,2005]. So far the degree <strong>of</strong> deviation <strong>of</strong> the grain boundarydistributions from the r<strong>and</strong>om distribution has been evaluatedby plotting the observed/expected ratio <strong>of</strong> the like–like8<strong>of</strong>20302


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 9. (a) Grain size statistics for the studied samples presented in box-<strong>and</strong>-whiskers diagrams.Horizontal axis corresponds to the ferret diameter <strong>of</strong> grain size in micrometers, <strong>and</strong> the thick barrepresents the median <strong>of</strong> grain size distribution. Vertical axis shows samples arranged according to thedegree <strong>of</strong> deformation from least deformed S1 to most intensely deformed sample V1. B2 represents anexceptionally coarse-grained mylonitic b<strong>and</strong>ed orthogneiss. For each sample the plagioclase grain sizestatistics for interstitial phases (I = Pl2 grains), <strong>and</strong> for recrystallized grains forming aggregates (A = Pl1grains) are shown. The number <strong>of</strong> grains is indicated. Grain size distributions <strong>of</strong> isolated quartz in thematrix <strong>and</strong> quartz grains forming centimetric ribbons or augen are also differentiated. This diagram showsprogressive coarsening <strong>of</strong> both Pl1 (aggregate) <strong>and</strong> Pl2 (interstitial) grains. (b) Plot <strong>of</strong> crystal sizedistribution (CSD) for plagioclases (I = Pl2 interstitial grain, A = Pl1 recrystallized grains in aggregates).N 0 ln(mm 4 ) values on vertical axis represent density <strong>of</strong> grains per volume, <strong>and</strong> dimensionless Gt valuesreflect the grain size frequency distribution. This diagrams show decrease <strong>of</strong> N 0 values <strong>and</strong> increase <strong>of</strong> Gtvalues which in classical CSD plots represent decreasing values <strong>of</strong> the intercept <strong>of</strong> the CSD curve withvertical axis associated with decreasing slope [Higgins, 1998]. In the CSD theory this evolution meansdecreasing nucleation rate <strong>and</strong> increasing growth rate contribution to the shape <strong>of</strong> the grain size frequencyhistogram [Lexa et al., 2005].contacts <strong>of</strong> the two major minerals against each other [e.g.,McLellan, 1983]. Lexa et al. [2005] proposed a newdiagram where the c valuec ¼ Observed pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiExpectedExpectedis plotted against the ratio <strong>of</strong> orientation tensor eigenvaluesthat represent the degree <strong>of</strong> GBPO. If the solidified melt isidentified in a rock the GBPO may yield information aboutthe types <strong>of</strong> channel networks as defined by Sawyer [2001]<strong>and</strong> quantified by Hasalová etal.[2008] <strong>and</strong> Závada et al.[2007].[25] In this work we use a method proposed by Lexa et al.[2005] where the grain contact frequencies are used toassess the character <strong>of</strong> the spatial distributions <strong>of</strong> grainboundaries. Figure 11a shows that the plagioclase like-likecontacts for the Type I metagranite plots in an intermediatepart <strong>of</strong> the diagram <strong>and</strong> is characterized by a rather smallaggregate distribution. This is due to the presence <strong>of</strong> the twoplagioclase populations resulting from the recrystallization<strong>of</strong> the large plagioclase crystals (high aggregate distribution)<strong>and</strong> from the disintegration <strong>of</strong> the large alkalinefeldspar crystals into a mixture <strong>of</strong> albite <strong>and</strong> K-feldspar(Figure 4). With increasing deformation a clear evolution <strong>of</strong>the grain contact frequency toward strongly aggregateddistribution for weakly deformed samples <strong>of</strong> the Type IImicrostructure can be observed. This is probably due to thecoalescence <strong>of</strong> feldspar <strong>and</strong> plagioclase layers. At very highstrain intensities the plagioclase grain contact frequenciesevolve toward the r<strong>and</strong>om or regular types <strong>of</strong> distribution inType III microstructure. This type <strong>of</strong> evolution indicates analmost perfect mixing <strong>of</strong> the plagioclase with other mineralphases. The degree <strong>of</strong> GBPO <strong>of</strong> the like-like boundaries forplagioclase does not evolve with the above described trendbut remains rather constant for any degree <strong>of</strong> the finitestrain. The K-feldspar grain contact frequency shows asimilar behavior to the plagioclase but for some samples arather strong GBPO coupled with a decreasing grain contactfrequency was observed. The evolution <strong>of</strong> the unlikeplagioclase-K-feldspar grain boundaries exactly mirrorsthe evolution <strong>of</strong> the plagioclase like-like contact frequencybehavior (Figure 11b). For the observed constant grain sizeit is a geometrical necessity that an increase <strong>of</strong> the like-likecontacts causes a corresponding decrease <strong>of</strong> the unlikecontacts <strong>of</strong> feldspars.6. Crystallographic Preferred Orientation[26] The lattice-preferred orientation (LPO) <strong>of</strong> quartz,plagioclase <strong>and</strong> K-feldspar was determined using the electronbackscatter diffraction (EBSD) technique [Bascou et9<strong>of</strong>20303


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 10. Plot <strong>of</strong> the grain shape preferred orientations (SPO)<strong>of</strong> plagioclase <strong>and</strong> K-feldspar in studied samples. The resultsare summarized by a box-type plot <strong>of</strong> axial ratios versuseigenvalue ratios <strong>of</strong> bulk shape preferred orientation forindividual phases. Individual boxes show median <strong>and</strong> first<strong>and</strong> third quartile values. The whiskers represent a statisticalestimate <strong>of</strong> the range <strong>of</strong> data, while outliers are not shown.Vertical axis characterizes the shape <strong>of</strong> grains, while horizontalaxis represents the area-weighted degree <strong>of</strong> preferred orientation.Plagioclase (I = interstitial Pl2 grains, A= recrystallized Pl1 grainsforming aggregates). This diagram shows generally low aspectratios <strong>of</strong> plagioclase grains for all samples <strong>and</strong> exceptionallystrong SPO <strong>of</strong> some samples <strong>of</strong> Type II <strong>and</strong> III microstructures.al., 2001]. The lattice preferred orientation <strong>of</strong> quartz ispresented in the pole figures <strong>of</strong> the crystallographic directions< c > <strong>and</strong> < a > (Figure 12). In the case <strong>of</strong> plagioclase<strong>and</strong> K-feldspar, the list <strong>of</strong> operative slip systems [Kruse etal., 2001; Tullis, 1983] has been used <strong>and</strong> measured data hasbeen plotted in the pole figures <strong>of</strong> these crystallographicplanes <strong>and</strong> directions. Pole figures <strong>of</strong> principal slip directions<strong>and</strong> slip planes showing the best coincidence with themain axes <strong>of</strong> the finite strain ellipsoid <strong>of</strong> every sample arepresented (Figure 13). In the assessment <strong>of</strong> the maximaposition in the pole figures, it has been taken into accountthat samples M1, V1 reveal prolate shapes <strong>of</strong> the strainellipsoid. Hence the maxima <strong>of</strong> poles in the slip planes mightoccur along a girdle perpendicular to the slip direction.6.1. Quartz Augens <strong>and</strong> Ribbons CrystallographicPreferred Orientation (CPO)[27] The quartz c axis fabric for Type I microstructure(sample S1) shows a strong central maximum <strong>and</strong> weakersubmaxima close to the periphery <strong>of</strong> the diagram suggestingan ill-defined oblique cross girdle pattern with openingangles <strong>of</strong> around 45° (Figure 12). The fabric is a result <strong>of</strong>the dominant prism < a > <strong>and</strong> subordinate basal < a > slipsystems <strong>and</strong> a plane strain noncoaxial deformation. Type IImicrostructures (samples T2 <strong>and</strong> M1) show strong, central caxis maxima. There is also a tendency to form a singlegirdle <strong>of</strong> c axes distribution oriented oblique to the foliation(T2). This c axis pattern is commonly interpreted as a result<strong>of</strong> a prism < a > slip activity <strong>and</strong> a noncoaxial deformation[Schmid <strong>and</strong> Casey, 1986]. The Type III microstructure(samples R3 <strong>and</strong> V1) is characterized by maxima locatedeither at the periphery <strong>of</strong> the diagram or in an intermediateposition. The maxima are organized either along singlegirdles oblique to the foliation (sample R3) or symmetricallyin case <strong>of</strong> sample V1. These c axis patterns developedfrom combined activity <strong>of</strong> the basal < a > <strong>and</strong> rhomb < a + c> slip systems with a minor contribution <strong>of</strong> the prism < a >slip during the noncoaxial deformation.6.2. Plagioclase <strong>and</strong> K-Feldspar CPO[28] Plagioclase CPO shows a weakening <strong>and</strong> an increasedactivity <strong>of</strong> secondary <strong>and</strong> tentative slip systemswith micro<strong>structural</strong> evolution from Type I to Type IIImicrostructures (Figure 13). A frequently described slipsystem (010)[100] [Kruhl, 1996; Martelat et al., 1999;Schulmann et al., 1996] has been observed only in the TypeI microstructures (S1) <strong>and</strong> shows an asymmetry <strong>of</strong> theposition <strong>of</strong> the maxima with respect to the foliation. Theplagioclase from the Type II <strong>and</strong> III microstructures showedless common slip systems that are supposed to be secondaryor tentative [Kruse et al., 2001]. In the Type II microstructure(T2 <strong>and</strong> M1), the CPO shows active slip systems withdissociated Burgers vectors [Montardi <strong>and</strong> Mainprice,1987; Olsen <strong>and</strong> Kohlstedt, 1985] namely 1/2[112](111)<strong>and</strong> 1/2[111](101) in the sample T2, <strong>and</strong> 1/2[112](110) <strong>and</strong>1/2[112](201) in the sample M1. Brief inspection <strong>of</strong>both pairs <strong>of</strong> slip systems shows that each pair occupies aclose position in the plagioclase crystal [Kruse et al., 2001,Figure 2] <strong>and</strong> it is very likely that they operated simultaneously.The Type III microstructures show very weak crystallographicpreferred orientation <strong>of</strong> the plagioclase <strong>and</strong> sampleR3 again reveals an activity on the 1/2[112](110) slip systemswith dissociated Burgers vector (Figure 13).[29] The K-feldspar phenocryst <strong>and</strong> albite neoblasts <strong>of</strong>Type I orthogneiss show the same crystallographic orientations(sample S1, Figure 13) that do not coincide with anyknown slip systems. Albite neoblasts originate along thestrings that are parallel to (100) planes. In the Types II <strong>and</strong>III microstructures, the CPO <strong>of</strong> K-feldspar recrystallizedgrains reveal similar pattern indicative <strong>of</strong> the slip directionparallel to the 1/2[110] in all samples that operate on the(001), (111) slip planes (Figure 13). A less pronounced slipsystem [101](101) has been recognized in the augen gneisssample T2 from the Type II microstructure.7. Petrological Modeling[30] The <strong>structural</strong> position <strong>and</strong> shapes <strong>of</strong> interstitialphases, the character <strong>of</strong> mineral zoning <strong>and</strong> the compositionalvariations <strong>of</strong> the plagioclase (Pl2) in both Type II <strong>and</strong>III microstructures indicate the presence <strong>of</strong> some kind <strong>of</strong>fluid during the deformation [Fitz Gerald <strong>and</strong> Stünitz, 1993;Sawyer, 2001; Hasalová etal., 2008; Závada et al., 2007].Therefore, the pseudosection modeling was performed inorder to examine whether the deformation process in thestudied orthogneisses occurred in the presence <strong>of</strong> a melt oran aqueous fluid.10 <strong>of</strong> 20304


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 11. Plot <strong>of</strong> the grain boundary frequencies. Plots <strong>of</strong> the deviations from r<strong>and</strong>om spatialdistribution versus the degree <strong>of</strong> grain boundary preferred orientation. The value <strong>of</strong> the deviation fromr<strong>and</strong>om spatial distribution is obtained by the contact frequency method [Kretz, 1969], <strong>and</strong> the lengthweighteddegree <strong>of</strong> preferred orientation is estimated as the ratio <strong>of</strong> eigenvalues <strong>of</strong> the bulk matrix <strong>of</strong>inertia. The plot <strong>of</strong> the like-like <strong>and</strong> unlike boundaries is separated. (a) The diagram <strong>of</strong> feldspars like-likeboundaries shows a trend from the weak aggregate distribution <strong>of</strong> Type 1 microstructure, to the highlyaggregate distribution in Type II microstructures, <strong>and</strong> almost r<strong>and</strong>om distribution in highly mixed TypeIII microstructures. (b) The diagram <strong>of</strong> feldspar unlike boundaries mirrors the like-like evolutionarytrend.7.1. Modeling Method[31] The pseudosections were calculated in the system Na 2 O-CaO-K 2 O-FeO-MgO-Al 2 O 3 -SiO 2 -H 2 O (NCKFMASH). All theFe was treated as FeO <strong>and</strong> molar amounts <strong>of</strong> the consideredoxides were recalculated to 100 mol %. The calculationswere performed using THERMOCALC 3.25 [Powell et al.,1998] <strong>and</strong> the data set 5.5 [Holl<strong>and</strong> <strong>and</strong> Powell, 1998]. Themixing models for the most solid solutions were takenfrom White et al. [2001] <strong>and</strong> the THERMOCALC documentation[Powell <strong>and</strong> Holl<strong>and</strong>, 2004]. The feldspars areformulated using the model <strong>of</strong> Holl<strong>and</strong> <strong>and</strong> Powell [2003],the paragonite-muscovite solution is after Coggon <strong>and</strong>Holl<strong>and</strong> [2002]. The quartz, K-feldspar <strong>and</strong> plagioclaseare in excess <strong>and</strong> the pseudosections are contoured forliquid mole isopleths.7.2. Results[32] The first pseudosection is calculated with the amount<strong>of</strong> H 2 O, M(H 2 O) = 2.68 mol % in the whole rock composition,which is the amount <strong>of</strong> H 2 O tied in micas in theassemblage biotite-muscovite-plagioclase-K-feldspar-quartz(Bt-Ms-Pl-Kfs-Qtz) at 8 kbar <strong>and</strong> 600°C (Figure 14a). Thelimiting maximum pressure for the studied rock with theassemblage Bt-Ms-Pl-Kfs-Qtz is the appearance <strong>of</strong> garnet at8–9 kbar, the maximum temperature limit is the beginning<strong>of</strong> biotite dehydration melting marked by the appearance <strong>of</strong>garnet at temperatures <strong>of</strong> 680–700°C <strong>and</strong> the breakdown <strong>of</strong>muscovite at 680–700°C below 6 kbar that is not observed.The minimum pressure <strong>and</strong> temperature conditions cannotbe determined from the assemblage <strong>of</strong> the studied sample,therefore, we used the P-T path <strong>of</strong> Pitra <strong>and</strong> Guiraud [1996]that was determined in the neighboring metapelites. It ischaracterized by decompression from 8–9 kbar <strong>and</strong> 610–660°C to 4–5 kbar <strong>and</strong> 600–650°C. Path 1 (Figure 14a) ischaracterized first by heating followed by decompression.Melting in the assemblage Bt-Ms-Pl-Kfs-Qtz without externalH 2 O starts at c. 660°C <strong>and</strong> 8 kbars <strong>and</strong> the maximumamount <strong>of</strong> melt produced in the field <strong>of</strong> Bt-Ms-Pl-Kfs-Qtz-Liq is less then 1 mol %.[33] The evolution <strong>of</strong> assemblages that the rock compositionproduces with a varying amount <strong>of</strong> H 2 O(M(H 2 O) =0–4.78 mol %) on heating at 8 kbar is examined inFigure 14b. The evolution for the amount <strong>of</strong> H 2 O tied inmicas (=2.68 mol %) in a rock with the starting assemblageBt-Ms-Pl-Kfs-Qtz is indicated by path 1, <strong>and</strong> is equivalentto the isobaric heating path in Figure 14a. The assemblageBt-Ms-Pl-Kfs-Qtz starts to melt at c. 660°C, <strong>and</strong> producesless then 0.5% <strong>of</strong> melt at c. 685°C (path 1 in Figure 14b).The upper temperature limit is the appearance <strong>of</strong> garnet at c.690°C that is not observed in the studied rocks. The loweramount <strong>of</strong> H 2 O in the rock stabilizes the anhydrous phasessuch as the garnet or kyanite <strong>and</strong> higher amounts <strong>of</strong> H 2 Oindicates the presence <strong>of</strong> free aqueous fluid below 630°C.[34] If above the temperature <strong>of</strong> c. 630°C appears freeH 2 O in the rock that follows the path 1, the rock compositionis immediately drawn to the right side, <strong>and</strong> starts to11 <strong>of</strong> 20305


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406melt (for example path 2 in Figure 14a). This leucocraticmelt is generated by the water released by the dehydrationprocess in which water migrates upward into the melt-fertilerocks that are above the wet-solidus temperature [Scaillet etal., 1990; Thompson <strong>and</strong> Connolly, 1995]. The quantity <strong>of</strong>produced melt depends on the amount <strong>of</strong> available externalH 2 O. For example the total amount <strong>of</strong> H 2 O for path 2 isM(H 2 O) = 4.28 mol %, which is 1.60 mol % <strong>of</strong> H 2 O addedto 2.68 mol % H 2 O already present in the micas. Such H 2 Oaddition results in the production <strong>of</strong> 5 mol % <strong>of</strong> melt at660°C <strong>and</strong> 8 kbar.[35] The evolution during decompression was also studiedin P-M(H 2 O) sections at 660°C. Following a decompressionfrom 8 to 4.5 kbar, the rock with the original H 2 Otied in micas increases the amount <strong>of</strong> melt to 1 mol %. In arock that contains added H 2 O, the melt fraction increases by1–2.5%, for example for path 2 the total amount <strong>of</strong> melt at4.5 kbar is 7.5 mol % because <strong>of</strong> 1.6 mol % added H 2 Oonthe prograde path.8. Discussion[36] Here we discuss the anomalously weak K-feldspar<strong>and</strong> plagioclase compared to quartz in light <strong>of</strong> the strain data<strong>and</strong> quantitative micro<strong>structural</strong> <strong>and</strong> textural analysis. Thequartz-feldspar rheology inversion is discussed as a result <strong>of</strong>the alkaline feldspar breakdown <strong>and</strong> the formation <strong>of</strong> a finegrainedfeldspar matrix. Finally, we propose a model <strong>of</strong>mixing <strong>of</strong> the recrystallized plagioclase <strong>and</strong> K-feldspargrains with the silicate melt associated with grain boundarysliding controlled diffusion creep followed by a granularflow at higher melt proportions.8.1. Comparative Quartz <strong>and</strong> K-Feldspar Rheology[37] On the basis <strong>of</strong> the number <strong>of</strong> experimental data, thequartz can be considered to be significantly weaker comparedto the K-feldspar <strong>and</strong> plagioclase for a wide range <strong>of</strong>temperatures [Jaoul et al., 1984; Kronenberg <strong>and</strong> Tullis,1984; Tullis, 1990]. This is supported by a number <strong>of</strong>studies <strong>of</strong> quartzo-feldspathic rocks deformed at mediumtohigh-temperature conditions [Gapais, 1989; H<strong>and</strong>y,1994a; Schulmann et al., 1996]. It is only at greenschistfacies conditions <strong>and</strong> in the presence <strong>of</strong> hydrous fluids,when feldspars become weaker than quartz via destabilization<strong>and</strong> breakdown to mixture <strong>of</strong> retrograde fine-grainedreaction products as white mica, quartz (K-feldspar breakdown)<strong>and</strong> epidote, albite ± garnet (plagioclase breakdown)[Fitz Gerald <strong>and</strong> Stünitz, 1993; H<strong>and</strong>y, 1990; Stünitz <strong>and</strong>Fitz Gerald, 1993]. Nevertheless, it is assumed that thestrength <strong>of</strong> feldspars is generally several orders <strong>of</strong> magnitudehigher than that <strong>of</strong> quartz [H<strong>and</strong>y, 1994a; H<strong>and</strong>y et al.,1999; Ranalli <strong>and</strong> Murphy, 1987].[38] The shapes <strong>of</strong> quartz <strong>and</strong> feldspar polycrystallineaggregates measured in this work clearly show that for12 <strong>of</strong> 20Figure 12. Electron backscatter diffraction in situ measurementson quartz ribbons. Pole diagrams showingcontoured crystallographic orientation (projected in lowerhemisphere equal area). Contoured at multiples <strong>of</strong> uniformdistribution (maximum 10 uniform). The foliation normal isN-S, <strong>and</strong> the stretching lineation is E-W. Black squarescorrespond to the pole <strong>of</strong> the mean orientation. The polefigures show evolution <strong>of</strong> active slip systems withincreasing deformation from activity <strong>of</strong> combined prism <strong>and</strong> basal < a > slip systems for weakly deformed Type Imicrostructures, via the activity <strong>of</strong> prism < a > slip systemin intermediate Type II microstructures to combined activity<strong>of</strong> basal < a > <strong>and</strong> rhomb < a + c > slip systems in highlystrained Type III microstructures.306


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 13. Contoured pole figures <strong>of</strong> the most characteristic slip direction <strong>and</strong> slip plane <strong>of</strong> K-feldspar<strong>and</strong> plagioclase from the Type I to the Type III microstructures. Equal area projection, lower hemisphere.Contoured at interval 1.0 times <strong>of</strong> the uniform distribution. Foliation (full line) is horizontal, <strong>and</strong> lineationis in this plane in the E-W direction. N is the number <strong>of</strong> measured grains. Maximum densities (blacksquare in the pole figures) are marked below the pole figure. Plagioclase shows commonly reported slipsystem only for Type I microstructure <strong>and</strong> activity <strong>of</strong> secondary <strong>and</strong> tentative slip systems withdissociated Burgers vectors for Type II <strong>and</strong> III microstructures. The top right pole figures showdependency <strong>of</strong> CPO <strong>of</strong> the new exsolved albite grains Pl1b on the orientation <strong>of</strong> alkali feldspar host.K-feldspar reveals also slip systems with dissociated Burgers vectors for Types II <strong>and</strong> IIImicrostructures. See text for discussion.various deformation intensities the quartz aggregates areless deformed than the feldspar aggregates <strong>and</strong> the difference<strong>of</strong> D values between feldspar <strong>and</strong> quartz aggregatesincreases with increasing bulk deformation. We suggest thatthe studied samples reveal a higher competency <strong>of</strong> quartzcompared to the feldspar polycrystalline aggregates for allexamined strain intensities at high-temperature conditions.The strength reversal between the quartz <strong>and</strong> feldspar inexperimentally deformed aplite was documented byDell’Angelo <strong>and</strong> Tullis [1996]. These authors showedthat at 700°C, the dispersed quartz grains in the apliteremain less deformed than the feldspar grains while athigher temperatures the interconnected quartz becomesweaker than feldspars.8.2. Structural Evolutionary Trends <strong>of</strong> PolyphaseMylonites[39] We discuss here a hypothesis <strong>of</strong> micro<strong>structural</strong>evolutionary trend in which the Types I, II <strong>and</strong> III microstructuresrepresent different deformation stages along a13 <strong>of</strong> 20307


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 14. (a) Pseudosection showing the evolution <strong>of</strong> assemblages that the rock composition produceswith varying amounts <strong>of</strong> H 2 O(M(H 2 O) = 0–3.88 mol %) on heating at 8 kbar. (b) The pseudosectioncalculated with the amount <strong>of</strong> H 2 O tied in micas, deduced from pseudosection (Figure 14a). Thepseudosection shows a prograde path between 7 <strong>and</strong> 8 kbar along which the melt appears above 640°Cinthe stability <strong>of</strong> Bt-Ms-Liq-Pl-Ksp-Qtz. For comparison, the contours <strong>of</strong> melt production in other fields arealso shown. Mineral abbreviations correspond to those <strong>of</strong> Kretz [1983].progressive deformation path. Theoretically, the loadbearingframework (LBF) minerals (LBF = interconnectedstrong phase) should shield a weak phase from viscousdeformation for weak phase fractions lower than 20%[H<strong>and</strong>y, 1990]. With increasing strain the weak phase startsto interconnect along localized shear b<strong>and</strong>s, leading finallyto the development <strong>of</strong> a b<strong>and</strong>ed structure <strong>of</strong> alternatingmonomineralic layers [Jordan, 1988]. This corresponds toan evolution from high-viscosity contrast toward lowviscositycontrast interconnected weak layer (IWL) structures[H<strong>and</strong>y, 1994a].[40] The microstructure <strong>of</strong> the Type I orthogneisses showinterconnected recrystallized aggregates <strong>of</strong> plagioclase (representingthe weakest phase) surrounding strong clasts <strong>of</strong>quartz <strong>and</strong> alkali feldspar. This kind <strong>of</strong> microstructurecorresponds to IWL microstructure with the deformationhighly localized into rheologically weak plagioclase <strong>and</strong> arelatively high volume (up to 60 vol. %) <strong>of</strong> strong fraction(Figure 15a). This observation is valid for the macroscopicscale, but different rheological behavior can be observed atthe scale <strong>of</strong> the individual feldspar clasts. The internalstructure <strong>of</strong> the alkali feldspar phenocrysts shows typicalcharacteristics for a LBF structure formed by K-feldspar <strong>and</strong>aggregates <strong>of</strong> recrystallized plagioclase representing anisolated weak phase [Eudier, 1962; Jordan, 1987; Tharp,1983]. A rather high volume <strong>of</strong> weak pockets (20–30%)indicates a low stability <strong>of</strong> the LBF structure for small strainintensities [H<strong>and</strong>y, 1994a]. This is consistent with theobserved onset <strong>of</strong> coalescence <strong>of</strong> plagioclase chains alongmicroshear b<strong>and</strong>s in Figure 6b [Jordan, 1988].[41] The Type II microstructures are characterized by thecollapse <strong>of</strong> the internal LBF structure <strong>of</strong> alkali feldsparthrough the interconnection <strong>of</strong> recrystallized plagioclase <strong>and</strong>K-feldspar. This evolution is documented by the presence <strong>of</strong>oblique ‘‘bridges’’ <strong>of</strong> recrystallized plagioclases crossingrecrystallized K-feldspar aggregates <strong>and</strong> connecting surroundingplagioclase matrix (Figure 6b). From a rheologicalpoint <strong>of</strong> view, the volume <strong>of</strong> the weak matrix increased from 30 to 60 vol. % through the addition <strong>of</strong> entirely recrystallizedplagioclase <strong>and</strong> K-feldspar from original phenocrysts<strong>of</strong> Type I rock to already recrystallized plagioclase <strong>of</strong>Type II orthogneiss (Figure 15b). This evolution leads tothe development <strong>of</strong> IWL structures at all scales marked bya low volume <strong>of</strong> strong phases (quartz) <strong>and</strong> a moderateviscosity contrast indicated by the elongated shapes <strong>of</strong>quartz aggregates. An important feature <strong>of</strong> this stage is thecoalescence <strong>of</strong> K-feldspar <strong>and</strong> plagioclase leading to thedevelopment <strong>of</strong> almost monomineralic layers (Figure 6).[42] The Type III microstructure is characterized by thedestruction <strong>of</strong> feldspar monomineralic layering throughprogressive mixing <strong>of</strong> the feldspars <strong>and</strong> the subordinatesmall quartz grains forming a fine-grained matrix. The finalmicrostructure is represented by highly elongated quartzribbons surrounded by the homogeneously deformed feldspar-quartzmatrix. This type <strong>of</strong> rock microstructure correspondsto the IWL structures <strong>and</strong> is marked by a verylow-viscosity contrast between the quartz <strong>and</strong> the weakphases. The stronger quartz can be seen as a deformableinclusions in weak matrix (Figure 15c). In order to developsuch a microstructure the deformation mechanisms <strong>of</strong> feldsparshave to evolve toward similar efficiency.14 <strong>of</strong> 20308


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 15. The evolutionary trend <strong>of</strong> the metagranite deformation. (a) Orthogneiss Type I showinginterconnected weak layer (IWL) structure with strain concentration in plagioclase. The K-feldspar showsinternal load-bearing framework structure with weak plagioclase strings, while quartz exhibits weakCPO, both indicating shielding <strong>of</strong> quartz <strong>and</strong> feldspar from viscous flow. (b) Orthogneiss Type II ismarked by almost uniform flow in plagioclase <strong>and</strong> K-feldspar monomineralic layers around the strongquartz-IWL structure with low viscosity contrast. The rock shows development <strong>of</strong> monomineraliclayering <strong>and</strong> intense yielding <strong>of</strong> quartz marked by strong CPO <strong>and</strong> activity <strong>of</strong> prism < a > slip.(c) Orthogneiss Type III shows a uniform flow <strong>of</strong> all mineral phases <strong>and</strong> mixing <strong>of</strong> feldspars associatedwith important crystal growth. The quartz texture exhibits weakening comparing to the previous stage<strong>and</strong> activity <strong>of</strong> the basal < a > slip.8.3. Mechanism <strong>of</strong> Alkali Feldspar Breakdown:Transition From Type I to II Rock[43] The key element controlling the rheological development<strong>of</strong> the bulk rock is the mechanism <strong>of</strong> decomposition<strong>of</strong> originally large alkali feldspar phenocrysts into chains <strong>of</strong>plagioclase grains <strong>and</strong> adjacent K-feldspar hosts (Figure 4).EBSD measurements have shown that the texture <strong>of</strong> thenew plagioclase crystals in any <strong>structural</strong> position is almostidentical to that <strong>of</strong> the K-feldspar hosts (Figure 13), indicatingan orientation relationship between the host <strong>and</strong> theinclusion.[44] The above described features do not have an unambiguousexplanation. The coherent texture <strong>of</strong> the plagioclase<strong>and</strong> host K-feldspar grain may reflect a heterogeneousnucleation process [Putnis et al., 2003; Ribbe, 1983]. TheTime-Temperature-Transformation diagrams for the feldsparexsolution during cooling suggest that a process <strong>of</strong> heterogeneousnucleation is the most likely mechanism because <strong>of</strong>the slow cooling rates (which are likely to occur in deepseated intrusions) [Putnis et al., 2003]. Our micro<strong>structural</strong><strong>analyses</strong> show that the shape <strong>and</strong> size <strong>of</strong> the exsolutionpatterns is determined by the structure <strong>of</strong> the host namely by(100) <strong>and</strong> (010) planes <strong>of</strong> alkali feldspars. The experiments<strong>of</strong> Putnis et al. [2003] show that the albite rich regionsoriginate as a monoclinic feldspar exsolution in the K-richhost. With falling temperatures the albite changes to a highalbitetriclinic structure. The strain generated at the lamellainterface leads to the segmentation <strong>of</strong> the lamella by thealbite twinning reducing the strain energy across the interfaceregion. It is possible that such twinned lamellae coarsenduring the subsequent annealing which may lead to thedevelopment <strong>of</strong> new albite grains. The process <strong>of</strong> coarseningmay have been enhanced by the thermally induceddeformation so that first subgrains <strong>and</strong> subsequently developednew grains originated from the progressively deformingtwinned lamellae. All these processes may result in thedevelopment <strong>of</strong> chains <strong>of</strong> albite grains subparallel to theformer exsolution domains. The process <strong>of</strong> the heterogeneousnucleation is supported by the compositional pr<strong>of</strong>ileswhich show a decrease <strong>of</strong> albite components in the hostK-feldspar toward the albite boundary <strong>of</strong> rather constantcomposition similarly to the pr<strong>of</strong>iles published by Putniset al. [2003]. In theory the exsolution process should berelated either to the cooling history <strong>of</strong> crystallization <strong>of</strong>the granite or to the deformation-metamorphism processas suggested by White <strong>and</strong> Mawer [1986, 1988]. However,the spatial distribution <strong>of</strong> plagioclase chains in thefeldspar host <strong>and</strong> the crystallographic coherency suggestthat the transformation <strong>of</strong> the original alkali feldspar wasachieved by heterogeneous nucleation.8.4. Consequences <strong>of</strong> Feldspar Breakdown:Monomineralic Layering in Type II Rocks[45] The progressive textural evolution toward aggregatedistribution in the Type II microstructure is a typical feature<strong>of</strong> the high-grade deformation <strong>of</strong> granitoids [Gapais, 1989;H<strong>and</strong>y, 1994a; Schulmann et al., 1996]. The onset <strong>of</strong> granitedeformation is marked by the high stress concentrations inthe plagioclase grains due to the relative rheological inactivity<strong>of</strong> the quartz <strong>and</strong> K-feldspar at the onset <strong>of</strong> themetagranite deformation (Type I microstructure). The grainsize increase <strong>of</strong> the recrystallized plagioclase <strong>and</strong> K-feldsparassociated with the development <strong>of</strong> monomineralic layeringcan be interpreted in terms <strong>of</strong> stress relaxation [Hobbs,1981] coupled with a flow stress increase in the quartzaggregates indicated by the activity <strong>of</strong> the prism < a > slipsystem compared to the basal < a > in the Type I microstructure[Schmid <strong>and</strong> Casey, 1986]. Alternatively, theevolution <strong>of</strong> the quartz slip system can be attributed to theincreasing strain intensity in the Type II microstructure[Heilbronner <strong>and</strong> Tullis, 2006]. Rather unusual feldsparslip systems like once in Type II rocks (Figure 13) havebeen reported by Baratoux et al. [2005] from high-grademylonitic metagabbros <strong>and</strong> by Franěk et al. [2006] frompartially molten granulites. These authors interpreted theobserved slip systems as a result <strong>of</strong> a grain boundary slidingaccompanied with a crystal plastic deformation. All this15 <strong>of</strong> 20309


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406implies a progressive homogenization <strong>of</strong> the stress fieldduring this stage. Differences in the efficiency <strong>of</strong> thedeformation mechanisms lead to the development <strong>of</strong> monomineraliclayering precluding effective mechanical mixing.We emphasize that the involvement <strong>of</strong> feldspars in thehomogeneous flow increases the proportion <strong>of</strong> the weakmaterial up to 60%, which provided a significant drop <strong>of</strong> thebulk strain rate or stress assuming constant far field stressesor a bulk strain rate, respectively (Figure 15b). The effectiveweakening mechanism could be regarded in the presence <strong>of</strong>fluid or melt phase at grain boundaries in Type II rocks whichis represented by the interstitial albitic Pl2 <strong>and</strong> quartz formingthin films parallel to the K-feldspar <strong>and</strong> Pl1 boundaries or assmall cuspate grains in triple point junctions.8.5. Reasons for Transition From Type II to Type IIIRock: Melt-Assisted Mineral Mixing[46] Type III microstructures correspond to a mixing <strong>of</strong>fine-grained plagioclase <strong>and</strong> K-feldspar by two competitiveprocesses: a mechanical mixing accompanied by syndeformationalmass transfer either by diffusion [Baratouxet al., 2005; Kruse <strong>and</strong> Stünitz, 1999] or by melt/fluidredistribution along grain boundaries during the in situmelting [Závada et al., 2007; Hasalová et al., 2008] orinfiltration <strong>of</strong> hydrous fluids, respectively [Stünitz <strong>and</strong> FitzGerald, 1993]. The mechanical mixing process is supportedby the progressive thinning <strong>of</strong> the plagioclase aggregates ina K-feldspar matrix associated with the extreme stretching<strong>of</strong> the rock at very high strains. This process is inevitablyassociated with a grain boundary sliding mechanism <strong>and</strong>diffusion dominated creep. The growth <strong>of</strong> albite <strong>and</strong> quartzrims around the K-feldspar grains <strong>and</strong> the overall grain sizeincrease <strong>of</strong> both feldspars can result from the above mentionedmass transfer mechanisms.[47] The intensity <strong>of</strong> quartz CPO in Type III microstructuresdecreases compared to Type II microstructures. This isdue to a switch <strong>of</strong> the active slip system from prism < a >toward basal < a > <strong>and</strong> rhomb < a + c > slip systems in TypeIII microstructure (Figure 15c). These slip systems are notcommon in partially molten high-grade mylonites where theactivity <strong>of</strong> prism < c > glide is reported [e.g., Gapais <strong>and</strong>Barbarin, 1986; Martelat et al., 1999]. Our observations donot indicate a late reworking during a low-temperature event<strong>and</strong> we therefore suggest that the activity <strong>of</strong> the ‘‘lowtemperature’’slip systems in high-grade mylonites resultsfrom strain rate or stress conditions which are currentlyunconstrained [Závada et al., 2007; Hasalová etal., 2008].Consequently, the deformation mechanism in quartz is thedislocation creep <strong>and</strong> the evolution <strong>of</strong> slip systems suggest adecrease <strong>of</strong> the flow stress or strain rate compared toprevious stages [H<strong>and</strong>y, 1990; Herwegh et al., 1997;Schmid <strong>and</strong> Casey, 1986]. However, the diffusional creepin the feldspar matrix supported by grain shapes <strong>and</strong>crystallographic preferred orientations (Figure 13) showsthat the Type III microstructures represent a deformationstage marked by the overall flow stress drop associated withthe increase <strong>of</strong> homogeneity <strong>of</strong> the strain distribution. Thisoverall weakening is consistent with interstitial phasesforming a substantial proportion <strong>of</strong> the rock volume <strong>and</strong>shapes <strong>of</strong> pools or thick films <strong>of</strong> Pl2 <strong>and</strong> quartz grains thatare similar to the melt topology described by Sawyer [2001]or Marchildon <strong>and</strong> Brown [2003].8.6. <strong>Quantitative</strong> Micro<strong>structural</strong> <strong>and</strong> PetrologicalArguments for Syn-deformational Melting[48] The CSD curves (Figure 9b) show decreasing N 0 <strong>and</strong>increasing G t values in relict plagioclase aggregates towardType III rocks. The trend from Type II to Type III rocks(Figure 9b) consistent with the rapid nucleation <strong>of</strong> interstitialPl2 is marked by the presence <strong>of</strong> thin isolated films inSEM images <strong>of</strong> Type II microstructures (Figure 6, samplesT1 <strong>and</strong> M2, <strong>and</strong> Figure 8). The decrease <strong>of</strong> N 0 <strong>and</strong> increase<strong>of</strong> G t values in Type III microstructures correspond to agrowth <strong>of</strong> Pl2 pools <strong>and</strong> large new Pl2 rims in a K-feldspardominated matrix (Figure 7, samples V1 <strong>and</strong> D1). Thisevolution corrobarates the weakening <strong>of</strong> the aggregatedistribution in Type II microstructures toward an almostr<strong>and</strong>om distribution in Type III microstructures <strong>and</strong> anassociated development <strong>of</strong> important proportion <strong>of</strong> theinterstitial phases (Figure 11).[49] The pseudosection modeling indicates that the onlyfluid that could be present during the deformation <strong>of</strong> theorthogneiss at the estimated temperatures <strong>of</strong> 650–680°C isa silicate melt (Figure 14). Either the temperature increaseor the pressure drop along the path estimated by Pitra <strong>and</strong>Guiraud [1996] or Tajcmanová etal.[2006] for associatedmetapelites can be responsible for the dehydration melting<strong>of</strong> the muscovite-biotite bearing orthogneiss assemblage.The amount <strong>of</strong> melting may even be increased by externalfluids introduced from the surrounding metapelites[Thompson <strong>and</strong> Connolly, 1995].[50] The nucleation dominated part <strong>of</strong> the deformationmeltinghistory in Type II microstructures is probablyassociated with the onset <strong>of</strong> the melting reaction <strong>and</strong>heterogeneous nucleation <strong>of</strong> the melt droplets at highenergytriple points or noncoherent grain boundaries. Thegrowth dominated process associated with the Type IIImicrostructures reflects a more advanced melting leadingto the development <strong>of</strong> large pools <strong>and</strong> the coalescence <strong>of</strong>small nuclei along all boundaries in deformed aggregates.This process is best exemplified by the sample B2 whichresembles a migmatitic structure at the macroscopic scale.8.7. Grain Boundary Sliding Diffusional Creep <strong>and</strong>Variations in Melt Topology With Increasing MeltFraction[51] The studied micro<strong>structural</strong> sequence is characterizedby the increasing deformation intensity (Figure 3), theincreasing melt fraction (Figures 4, 6, <strong>and</strong> 7), grain size(Figure 9) <strong>and</strong> r<strong>and</strong>om grain contact distribution as well asthe low aspect ratio <strong>of</strong> grains (Figures 10 <strong>and</strong> 11). Thisevolutionary trend is accompanied by a change in orientation<strong>of</strong> the melt seams (Figures 8 <strong>and</strong> 16) from an orientationsubparallel to the foliation to seams oriented at highangle to the foliation.[52] Melt seams preferentially located along grain boundariesparallel to the foliation in Type II microstructures arecommonly reported in natural samples [Sawyer, 2001;Rosenberg <strong>and</strong> Berger, 2001] <strong>and</strong> in some experimentalor analog studies [Daines <strong>and</strong> Kohlstedt, 1997; Groebner<strong>and</strong> Kohlstedt, 2006; Walte et al., 2005]. This geometry canbe observed in microstructures <strong>of</strong> the T1 <strong>and</strong> T2 samples<strong>and</strong> is compatible with the combined activity <strong>of</strong> grainboundary sliding <strong>and</strong> crystal-plastic deformation mechanismsat very low melt fractions (


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Figure 16. The evolution <strong>of</strong> (a) shear plane-parallel melt preferred orientation at low melt fractions <strong>and</strong>(b) shear plane-perpendicular melt preferred orientation at high melt fractions. At low melt fractions,shear is accommodated by the grain boundary sliding parallel to the shear direction. Since grain boundarysliding means a loss <strong>of</strong> the cohesion along these boundaries, they represent an easy path for melt whichaccumulates at these boundaries (bottom left, Figure 16a). At higher melt fractions, a switch from shearplane parallel to shear plane-perpendicular melt preferred orientation can be observed. This is probablybecause <strong>of</strong> stress concentrations at grain boundaries oriented perpendicular to simple shear <strong>and</strong>accompanying dilation/cavitation <strong>of</strong> these boundaries (bottom right, Figure 16b). See text for furtherexplanations.orientation <strong>of</strong> melt seams can be explained by a loss <strong>of</strong>cohesion at boundaries oriented parallel to the foliation.Because <strong>of</strong> shearing, the melt concentrated at triple pointssuffers an increasing overpressure. We suggest that thecombination <strong>of</strong> loss <strong>of</strong> cohesion <strong>and</strong> build up <strong>of</strong> meltoverpressure leads to the injection <strong>of</strong> melt between activelymoving grain boundaries oriented at a high angle to theprincipal compressive stress, i.e., parallel to the macroscopicfoliation (Figure 16a). Oblique intragranular fractures canbe interpreted as Riedel shears channeling melt from overpressured,dilatant grain boundaries accompanying movementsalong foliation-parallel grain boundaries [Rosenberg<strong>and</strong> H<strong>and</strong>y, 2000]. The M1 <strong>and</strong> M2 samples are characterizedby a general constrictional deformation, lower aspectratio <strong>of</strong> grains <strong>and</strong> low melt fraction (


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406suitably oriented planes in the aggregate [Hubbert <strong>and</strong>Rubey, 1959]. As a result <strong>of</strong> melt overpressure, the feldsparaggregate dilates <strong>and</strong> the melt is accumulated in intergranular<strong>and</strong> intragranular pockets <strong>and</strong> fractures.[56] It is likely that the evolution in the orientation <strong>of</strong> meltpockets is caused by the increasing melt fraction <strong>and</strong>/or bythe change in differential stress magnitude <strong>and</strong> orientation[Cosgrove, 1997]. The implication <strong>of</strong> the increasing meltfraction <strong>and</strong> connectivity certainly influences the distribution<strong>of</strong> local compressive <strong>and</strong> shear stress magnitudes<strong>and</strong> orientations <strong>and</strong> the active deformation mechanism(Figure 16). In this context, it is not likely that the meltpreferred orientation is controlled by the variations indifferential stress [Gleason et al., 1999], hydrostatic annealing[Daines <strong>and</strong> Kohlstedt, 1997] or variations in thewetting angle [Walte et al., 2005]. In order to further explainthis relationship between melt seam orientations <strong>and</strong> meltfractions more analogs <strong>and</strong> possibly <strong>numerical</strong> modeling isneeded.[57] In Type II microstructures the diffusion creep rates<strong>of</strong> both feldspars were effectively enhanced by the interstitialmelt phase wetting their boundaries <strong>and</strong> leading to arapid strength drop within the MCT field as proposed by[Rosenberg <strong>and</strong> H<strong>and</strong>y, 2005]. This is consistent with a lowamount <strong>of</strong> melt


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406Gapais, D., <strong>and</strong> B. Barbarin (1986), Quartz fabric transition in a coolingsyntectonic granite (Hermitage massif, France), Tectonophysics, 125,357–370, doi:10.1016/0040-1951(86)90171-X.Gay, N. C. (1968a), Pure shear <strong>and</strong> simple shear deformation <strong>of</strong> inhomogeneousviscous fluids. 1. Theory, Tectonophysics, 5, 211–234,doi:10.1016/0040-1951(68)90065-6.Gay, N. C. (1968b), Pure shear <strong>and</strong> simple shear deformation <strong>of</strong> inhomogeneousviscous fluids. 2. The deformation <strong>of</strong> the total finite strain in arock from objects such as deformed pebbles, Tectonophysics, 5, 295–302, doi:10.1016/0040-1951(68)90033-4.Gleason, G. C., V. Bruce, <strong>and</strong> H. W. Green (1999), Experimental investigation<strong>of</strong> melt topology in partially molten quartzo-feldspathic aggregatesunder hydrostatic <strong>and</strong> non-hydrostatic stress, J. Metamorph. Geol., 17,705–722, doi:10.1046/j.1525-1314.1999.00228.x.Groebner, N., <strong>and</strong> D. L. Kohlstedt (2006), Deformation induced metal meltnetworks in silicates: Implications for core-mantle interactions in planetarybodies, Earth Planet. Sci. Lett., 245, 571 – 580, doi:10.1016/j.epsl.2006.03.029.H<strong>and</strong>y, M. R. (1990), The solid-state flow <strong>of</strong> polymineralic rocks, J. Geophys.Res., 95, 8647–8661, doi:10.1029/JB095iB06p08647.H<strong>and</strong>y, M. R. (1994a), The energetics <strong>of</strong> steady-state heterogeneous shearin mylonitic rock, Mater. Sci. Eng. A, 175, 261–272, doi:10.1016/0921-5093(94)91065-0.H<strong>and</strong>y, M. R. (1994b), Flow laws for rocks containing two non-linearviscous phases: A phenomenological approach, J. Struct. Geol., 16,287–301, doi:10.1016/0191-8141(94)90035-3.H<strong>and</strong>y, M. R., S. Wissing, <strong>and</strong> J. E. Streit (1999), Strength <strong>and</strong> structure <strong>of</strong>mylonite with combined frictional-viscous rheology <strong>and</strong> varied bimineraliccomposition, Tectonophysics, 303, 175 – 191, doi:10.1016/S0040-1951(98)00251-0.Hasalová, P., K. Schulmann, O. Lexa, P. Štípská, <strong>and</strong> F. Hrouda (2008),Origin <strong>of</strong> felsic migmatites by progressive destruction <strong>of</strong> metamorphiclayering during ductile shearing <strong>and</strong> melt infiltration, J. Metamorph.Geol., 26, 29–53.Heilbronner, R., <strong>and</strong> J. Tullis (2006), Evolution <strong>of</strong> c axis pole figures <strong>and</strong>grain size during dynamic recrystallization: Results from experimentallysheared quartzite, J. Geophys. Res., 111, B10202, doi:10.1029/2005JB004194.Herwegh, M., M. R. H<strong>and</strong>y, <strong>and</strong> R. Heilbronner (1997), Temperature- <strong>and</strong>strain-rate-dependent micr<strong>of</strong>abric evolution in monomineralic mylonite:Evidence from in situ deformation <strong>of</strong> norcamphor, Tectonophysics, 280,83–106, doi:10.1016/S0040-1951(97)00139-X.Higgins, M. D. (1998), Origin <strong>of</strong> anorthosite by textural coarsening: <strong>Quantitative</strong>measurements <strong>of</strong> a natural sequence <strong>of</strong> textural development, J.Petrol., 39, 1307–1325, doi:10.1093/petrology/39.7.1307.Hobbs, B. E. (1981), The influence <strong>of</strong> metamorphic environment upon thedeformation <strong>of</strong> minerals, Tectonophysics, 78, 335 – 383, doi:10.1016/0040-1951(81)90020-2.Holl<strong>and</strong>, T. J. B., <strong>and</strong> R. Powell (1998), An internally consistent thermodynamicdata set for phases <strong>of</strong> petrological interest, J. Metamorph. Geol.,16, 309–343, doi:10.1111/j.1525-1314.1998.00140.x.Holl<strong>and</strong>, T. J. B., <strong>and</strong> R. Powell (2003), Activity-composition relations forphases in petrological calculations: An asymmetric multicomponent formulation,Contrib. Mineral. Petrol., 145, 492–501, doi:10.1007/s00410-003-0464-z.Hubbert, M. K., <strong>and</strong> W. W. Rubey (1959), Role <strong>of</strong> fluid pressure in mechanics<strong>of</strong> overthrust faulting, Geol. Soc. Am. Bull., 70, 116–166.Jaoul, O., J. Tullis, <strong>and</strong> A. Kronenberg (1984), The effect <strong>of</strong> varying watercontents on the creep behavior <strong>of</strong> Heavitree Quartzite, J. Geophys. Res.,89, 4298–4312, doi:10.1029/JB089iB06p04298.Ji, S., <strong>and</strong> P. Zhao (1994), Strength <strong>of</strong> two-phase rocks: A model based onfiber-loading theory, J. Struct. Geol., 16, 253–262, doi:10.1016/0191-8141(94)90108-2.Ji, S., R. Wirth, E. Rybacki, <strong>and</strong> J. Zhenting (2000), High-temperatureplastic deformation <strong>of</strong> quartz-plagioclase multilayers by layer-normalcompression, J. Geophys. Res., 105, 16,651 –16,664, doi:10.1029/2000JB900130.Jordan, P. G. (1987), The deformational behaviour <strong>of</strong> bimineralic limestone-haliteaggregates, Tectonophysics, 135, 185– 197, doi:10.1016/0040-1951(87)90160-0.Jordan, P. G. (1988), The rheology <strong>of</strong> polymineralic rocks — An approach,Geol. Rundsch., 77, 285–294, doi:10.1007/BF01848690.Kachlík, V. (1999), Relationship between Maldanubicum (Central Bohemia,Czech Republic): A result <strong>of</strong> the polyphase Variscan nappe tectonics,J. Czech Geol. Soc., 44, 201–291.Kassner, M. E., <strong>and</strong> T. A. Hayes (2003), Creep cavitation in metals, Int. J.Plast., 19, 1715–1748, doi:10.1016/S0749-6419(02)00111-0.Kretz, R. (1969), On the spatial distribution <strong>of</strong> crystals in rocks, Lithos, 2,39–66, doi:10.1016/S0024-4937(69)80005-8.Kretz, R. (1983), Symbols for rock-forming minerals, Am. Mineral., 68,277–279.Kronenberg, A., <strong>and</strong> J. Tullis (1984), Flow strengths <strong>of</strong> quartz aggregates;grain size <strong>and</strong> pressure effects due to hydrolytic weakening, J. Geophys.Res., 89, 4281–4297, doi:10.1029/JB089iB06p04281.Kruhl, J. H. (1996), Prism- <strong>and</strong> basal-plane parallel subgrain boundaries inquartz: A micro<strong>structural</strong> geothermobarometer, J. Metamorph. Geol., 14,581–589, doi:10.1046/j.1525-1314.1996.00413.x.Kruse, R., <strong>and</strong> H. Stünitz (1999), Deformation mechanisms <strong>and</strong> phasedistribution in mafic high-temperature mylonites from the Jotun Nappe,southern Norway, Tectonophysics, 303, 223 – 249, doi:10.1016/S0040-1951(98)00255-8.Kruse, R., H. Stünitz, <strong>and</strong> K. Kunze (2001), Dynamic recrystallizationprocesses in plagioclase porphyroclast, J. Struct. Geol., 23, 1781–1802, doi:10.1016/S0191-8141(01)00030-X.Lexa, O., P. Štípská, K. Schulmann, L. Baratoux, <strong>and</strong> A. Kröner (2005),Contrasting textural record <strong>of</strong> two distinct metamorphic events <strong>of</strong> similarP-T conditions <strong>and</strong> different durations, J. Metamorph. Geol., 23, 649–666, doi:10.1111/j.1525-1314.2005.00601.x.Marchildon, N., <strong>and</strong> M. Brown (2003), Spatial distribution <strong>of</strong> melt-bearingstructures in anatectic rocks from Southern Brittany: Implications formelt-transfer at grain- to orogen-scale, Tectonophysics, 364, 215–235,doi:10.1016/S0040-1951(03)00061-1.Martelat, J.-E., K. Schulmann, J.-M. Lardeaux, <strong>and</strong> C. Nicollet (1999),Granulite micr<strong>of</strong>abric <strong>and</strong> deformation mechanisms in southern Madagascar,J. Struct. Geol., 21, 671– 687, doi:10.1016/S0191-8141(99)00052-8.McLellan, E. L. (1983), Contrasting textures in metamorphic <strong>and</strong> anatecticmigmatites: An example from the Scottish Caledonides, J. Metamorph.Geol., 1, 241–262, doi:10.1111/j.1525-1314.1983.tb00274.x.Medaris, L. G., J. H. Fournelle, E. D. Ghent, E. Jelínek, <strong>and</strong> Z. Mísař(1998), Prograde eclogite in the Gföhl Nappe, Czech Republic: Newevidence on Variscan high-pressure metamorphism, J. Metamorph. Geol.,16, 563–576, doi:10.1111/j.1525-1314.1998.00158.x.Montardi, Y., <strong>and</strong> D. Mainprice (1987), A transmission electron microscopicstudy <strong>of</strong> the plastic deformation <strong>of</strong> calcic plagioclase (An 68–70), Bull. Mineral., 110, 1–14.Olsen, T. S., <strong>and</strong> D. L. Kohlstedt (1985), Natural deformation <strong>and</strong> recrystalization<strong>of</strong> some intermediate plagioclase feldspars, Tectonophysics,111, 107–131, doi:10.1016/0040-1951(85)90067-8.Park, Y., <strong>and</strong> W. D. Means (1996), Direct observation <strong>of</strong> deformationprocesses in crystal mushes, J. Struct. Geol., 18, 847–858,doi:10.1016/S0191-8141(96)80017-4.Peterson, T. D. (1996), A refined technique for measuring crystal sizedistributions in thin section, Contrib. Mineral. Petrol., 124, 395–405,doi:10.1007/s004100050199.Pitra, P., <strong>and</strong> M. Guiraud (1996), Probable anticlockwise P-T evolution inextending crust: Hlinsko region, Bohemian Massif, J. Metamorph. Geol.,14, 49–60, doi:10.1111/j.1525-1314.1996.t01-1-00049.x.Powell, R., <strong>and</strong> T. J. B. Holl<strong>and</strong> (2004), Course notes for ‘‘THERMOCALCWorkshop 2004: Calculating metamorphic phase equilibria [CD-ROM],report, Eidg. Tech. Hochsch., Zurich, Zurich, Switz.Powell, R., T. Holl<strong>and</strong>, <strong>and</strong> B. Worley (1998), Calculating phase diagramsinvolving solid solutions via non-linear equations, with examples usingTHERMOCALC, J. Metamorph. Geol., 16, 577 – 588, doi:10.1111/j.1525-1314.1998.00157.x.Putnis, A., C. M. Pina, J. M. Astilleros, L. Fernández-Díaz, <strong>and</strong> M. Prieto(2003), Nucleation <strong>of</strong> solid solutions crystallizing from aqueous solutions,Philos. Trans. R. Soc. London, Ser. A, 361, 615–632,doi:10.1098/rsta.2002.1142.Ramsay, J. G., <strong>and</strong> M. I. Huber (1983), The Techniques <strong>of</strong> Modern StructuralGeology, vol. 1, Strain Analysis, 307 pp., Elsevier, London.Ranalli, G. (1995), Rheology <strong>of</strong> the Earth, 2nd ed., 366 pp., CRC Press,London.Ranalli, G., <strong>and</strong> D. C. Murphy (1987), Rheological stratification <strong>of</strong> thelithosphere, Tectonophysics, 132, 281– 295, doi:10.1016/0040-1951(87)90348-9.R<strong>and</strong>olph, A. D., <strong>and</strong> M. A. Larson (1971), Theory <strong>of</strong> Particle Processes,251 pp., Elsevier, San Diego, Calif.Ribbe, P. H. (1983), Chemistry structure <strong>and</strong> nomenclature <strong>of</strong> feldspars, inFeldspar Mineralogy, Rev. Mineral., vol. 2, 2nd ed., edited by P. H.Ribbe, pp. 21–55, Mineral. Soc. <strong>of</strong> Am., Washington, D. C.Rosenberg, C. L., <strong>and</strong> A. Berger (2001), Syntectonic melt pathways ingranite, <strong>and</strong> melt-induced transition in deformation mechanisms, Phys.Chem. Earth, Part A, 26, 287–293.Rosenberg, C. L., <strong>and</strong> M. R. H<strong>and</strong>y (2000), Syntectonic melt pathways duringsimple shearing <strong>of</strong> a partially molten rock analogue (Norcamphor-Benzamide), J. Geophys. Res., 105, 3135–3149, doi:10.1029/1999JB900371.Rosenberg, C. L., <strong>and</strong> M. R. H<strong>and</strong>y (2005), Experimental deformation <strong>of</strong>partially melted granite revisited: Implications for the continental crust,19 <strong>of</strong> 20313


B10406SCHULMANN ET AL.: RHEOLOGY OF PARTIALLY MOLTEN GNEISSESB10406J. Metamorph. Geol., 23, 19– 28, doi:10.1111/j.1525-1314.2005.00555.x.Rybacki, E., <strong>and</strong> G. Dresen (2004), Deformation mechanism mapsfor feldspar rocks, Tectonophysics, 382, 173–187, doi:10.1016/j.tecto.2004.01.006.Sawyer, E. W. (2001), Melt segregation in the continental crust: Distribution<strong>and</strong> movement <strong>of</strong> melt in anatectic rocks, J. Metamorph. Geol., 19,291–309, doi:10.1046/j.0263-4929.2000.00312.x.Scaillet, B., C. France-Lanord, <strong>and</strong> P. Le Fort (1990), Badrinath-Gangotriplutons (Garhwal, India): Petrological <strong>and</strong> geochemical evidence for fractionationprocesses in a high Himalayan leucogranite, J. Volcanol.Geotherm. Res., 44, 163–188, doi:10.1016/0377-0273(90)90017-A.Schmid, S. M. (1982), Micr<strong>of</strong>abric studies as indicators <strong>of</strong> deformationmechanisms <strong>and</strong> flow laws operative in mountain building, in MountainBuilding Processes, edited by K. J. Hsü, pp. 95–110, Elsevier, London.Schmid, S. M., <strong>and</strong> M. Casey (1986), Complete fabric analysis <strong>of</strong> somecommonly observed quartz c-axis patterns, in Mineral <strong>and</strong> Rock Deformation:Laboratory Studies, Geophys. Monogr. Ser., vol. 36, edited byB. E. Hobbs <strong>and</strong> H. C. Heard, pp. 263–286, AGU, Washington, D. C.Schmid, S. M., R. Panozzo, <strong>and</strong> S. Bauer (1987), Simple shear experimentson calcite rocks: Rheology <strong>and</strong> micr<strong>of</strong>abric, J. Struct. Geol., 9, 747–778,doi:10.1016/0191-8141(87)90157-X.Schmid, S. M., R. Heilbronner, <strong>and</strong> H. Stünitz (1999), Deformation mechanismsin nature <strong>and</strong> experiment (editorial remarks), Tectonophysics,303, vii–ix, doi:10.1016/S0040-1951(98)00302-3.Schulmann, K., B. Mlcoch, <strong>and</strong> R. Melka (1996), High-temperature microstructures<strong>and</strong> rheology <strong>of</strong> deformed granite, Erzgebirge, Bohemian Massif,J. Struct. Geol., 18, 719–733, doi:10.1016/S0191-8141(96)80007-1.Schulmann, K., A. Kröner, E. Hegner, I. Wendt, J. Konopásek, O. Lexa,<strong>and</strong> P. Štípská (2005), Chronological constraints on the pre-orogenichistory, burial <strong>and</strong> exhumation <strong>of</strong> deep-seated rocks along the easternmargin <strong>of</strong> the Variscan Orogen, Bohemian Massif, Czech Republic,Am. J. Sci., 305, 407–448, doi:10.2475/ajs.305.5.407.Simpson, C. (1985), Deformation <strong>of</strong> granitic rocks across the brittle-ductiletransition, J. Struct. Geol., 7, 503–511, doi:10.1016/0191-8141(85)90023-9.Sklenička, V., J. Saxl, <strong>and</strong> J. Čadek (1977), Intercrystalline fracturing athigh temperature creep <strong>of</strong> metals <strong>and</strong> alloys (in Czech), report, 108 pp.,Czech. Acad. <strong>of</strong> Sci., Prague.Stünitz, H., <strong>and</strong> J. D. Fitz Gerald (1993), Deformation <strong>of</strong> granitoids at lowmetamorphic grades. II. Granular flow in albite-rich mylonites, Tectonophysics,221, 299–324, doi:10.1016/0040-1951(93)90164-F.Synek, J., <strong>and</strong> D. Oliverová (1993), Terrane character <strong>of</strong> the north-eastmargin <strong>of</strong> the Moldanubian Zone: The Kutná Hora Crystalline Complex,Bohemian Massif, Geol. Rundsch., 82, 566 –582, doi:10.1007/BF00212417.Tajcmanová, L., J. Konopásek, <strong>and</strong> K. Schulmann (2006), Thermal evolution<strong>of</strong> the orogenic lower crust during exhumation within a thickenedMoldanubian root <strong>of</strong> the Variscan belt <strong>of</strong> Central Europe, J. Metamorph.Geol., 24, 119–134, doi:10.1111/j.1525-1314.2006.00629.x.Tharp, T. M. (1983), Analogies between the high-temperature deformation<strong>of</strong> polyphase rocks <strong>and</strong> the mechanical behavior <strong>of</strong> porous powder metal,Tectonophysics, 96, 1–11, doi:10.1016/0040-1951(83)90216-0.Thompson, A. B., <strong>and</strong> J. A. D. Connolly (1995), Melting <strong>of</strong> the continentalcrust: Some thermal <strong>and</strong> petrological constraints on anatexis in continentalcollision zones <strong>and</strong> other tectonic settings, J. Geophys. Res., 100,15,565–15,580, doi:10.1029/95JB00191.Treagus, S. H. (2002), Modelling the bulk vicosity <strong>of</strong> two-phase mixtures interm <strong>of</strong> clast shape, J. Struct. Geol., 24, 57 – 76, doi:10.1016/S0191-8141(01)00049-9.Tullis, J. (1983), Deformation <strong>of</strong> feldspars, in Feldspar Mineralogy, Rev.Mineral., vol. 2, 2nd ed., edited by P. H. Ribbe, pp. 297–323, Mineral.Soc. <strong>of</strong> Am., Washington, D. C..Tullis, J. (1990), Experimental studies <strong>of</strong> deformation mechanisms <strong>and</strong>microstructures in quartzo-feldspathic rocks, in Deformation Processesin Minerals, Ceramics <strong>and</strong> Rocks, edited by D. Barbour <strong>and</strong> P. Meredith,pp.190–227, CRC Press, Cambridge, U. K.Walte, N. P., P. D. Bons, <strong>and</strong> C. W. Passchier (2005), Deformation <strong>of</strong> meltbearingsystems — Insight from in situ grain-scale analogue experiments,J. Struct. Geol., 27, 1666–1679, doi:10.1016/j.jsg.2005.05.006.White, J. C., <strong>and</strong> C. K. Mawer (1986), Extreme ductility <strong>of</strong> feldspars from amylonite, Parry Sound, Canada, J. Struct. Geol., 8, 133 –143,doi:10.1016/0191-8141(86)90104-5.White, J. C., <strong>and</strong> C. K. Mawer (1988), Dynamic recrystallization <strong>and</strong>associated in perthites: Evidence <strong>of</strong> deep crustal thrusting, J. Geophys.Res., 93, 325–337, doi:10.1029/JB093iB01p00325.White, R. W., R. Powell, <strong>and</strong> T. J. B. Holl<strong>and</strong> (2001), Calculation <strong>of</strong> partialmelting equilibria in the system Na 2 O–CaO–K 2 O–FeO–MgO–Al 2 O 3 –SiO 2 –H 2 O (NCKFMASH), J. Metamorph. Geol., 19, 139–153, doi:10.1046/j.0263-4929.2000.00303.x.Závada, P., K. Schulmann, J. Konopásek, O. Lexa, <strong>and</strong> S. Ulrich (2007),Melt topology in deformed quartzo-feldspathic rocks: Implicationfor rheology <strong>and</strong> grain-scale migration <strong>of</strong> melt in partially molten crust,J. Geophys. Res., 112, B10210, doi:10.1029/2006JB004820.J. K. Becker, Institut für Geowissenschaften, Universität Tübingen,Sigwartstrasse 10, D-72076 Tübingen, Germany.O. Lexa, Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, CharlesUniversity, Albertov 6, 12843 Praha 2, Czech Republic.J.-E. Martelat, Laboratoire de Géodynamique des Chaînes Alpines,UMR5025, Université Joseph Fourier, Observatoire des Sciences del’Univers de Grenoble, CNRS, F-38041 Grenoble Cedex 9, France.K. Schulmann <strong>and</strong> P. Štípská, Centre de Géochimie de la Surface,UMR7516, Université Louis Pasteur, CNRS, F-67084 Strasbourg Cedex,France.S. Ulrich, Geophysical Institute, Czech Academy <strong>of</strong> Sciences, Boční II/1401, 14131 Praha 4, Czech Republic.20 <strong>of</strong> 20314


J. metamorphic Geol., 2008, 26, 29–53 doi:10.1111/j.1525-1314.2007.00743.xOrigin <strong>of</strong> migmatites by deformation-enhanced melt infiltration<strong>of</strong> orthogneiss: a new model based on quantitativemicro<strong>structural</strong> analysisP. HASALOVÁ, 1,2 K. SCHULMANN, 1 O. LEXA, 1,2 P. ŠTÍPSKÁ, 1 F. HROUDA, 2,3 S. ULRICH, 2,4J. HALODA 5 AND P. TÝCOVÁ 51 Université Louis Pasteur, CGS/EOST, UMR 7517, 1 rue Blessig, Strasbourg 67084, France (hasalovap@seznam.cz)2 Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University, Albertov 6, 12843 Prague, Czech Republic3 AGICO, Ječná 29a, 621 00 Brno, Czech Republic4 Institute <strong>of</strong> Geophysics, Czech Academy <strong>of</strong> Sciences, Boční II/1401, 14131 Praha 4, Czech Republic5 Czech Geological Survey, Klárov 3, 118 21 Prague 1, Czech RepublicABSTRACTA detailed field study reveals a gradual transition from high-grade solid-state b<strong>and</strong>ed orthogneiss viastromatic migmatite <strong>and</strong> schlieren migmatite to irregular, foliation-parallel bodies <strong>of</strong> nebulitic migmatitewithin the eastern part <strong>of</strong> the Gfo¨ hl Unit (Moldanubian domain, Bohemian Massif). The orthogneiss tonebulitic migmatite sequence is characterized by progressive destruction <strong>of</strong> well-equilibrated b<strong>and</strong>edmicrostructure by crystallization <strong>of</strong> new interstitial phases (Kfs, Pl <strong>and</strong> Qtz) along feldspar boundaries<strong>and</strong> by resorption <strong>of</strong> relict feldspar <strong>and</strong> biotite. The grain size <strong>of</strong> all felsic phases decreases continuously,whereas the population density <strong>of</strong> new phases increases. The new phases preferentially nucleate alonghigh-energy like–like boundaries causing the development <strong>of</strong> a regular distribution <strong>of</strong> individual phases.This evolutionary trend is accompanied by a decrease in grain shape preferred orientation <strong>of</strong> all felsicphases. To explain these data, a new petrogenetic model is proposed for the origin <strong>of</strong> felsic migmatites bymelt infiltration from an external source into b<strong>and</strong>ed orthogneiss during deformation. In this model,infiltrating melt passes pervasively along grain boundaries through the whole-rock volume <strong>and</strong> changescompletely its macro- <strong>and</strong> microscopic appearance. It is suggested that the individual migmatite typesrepresent different degrees <strong>of</strong> equilibration between the host rock <strong>and</strong> migrating melt duringexhumation. The melt topology mimicked by feldspar in b<strong>and</strong>ed orthogneiss forms elongate pocketsoriented at a high angle to the compositional b<strong>and</strong>ing, indicating that the melt distribution wascontrolled by the deformation <strong>of</strong> the solid framework. The microstructure exhibits features compatiblewith a combination <strong>of</strong> dislocation creep <strong>and</strong> grain boundary sliding deformation mechanisms. Themigmatite microstructures developed by granular flow accompanied by melt-enhanced diffusion <strong>and</strong>/ormelt flow. However, an AMS study <strong>and</strong> quartz micr<strong>of</strong>abrics suggest that the amount <strong>of</strong> melt present didnot exceed a critical threshold during the deformation to allow free movements <strong>of</strong> grains.Key words: crystal size distribution; melt infiltration; melt topology; migmatites; quantitative texturalanalysis.INTRODUCTIONMovement <strong>of</strong> a large volume <strong>of</strong> granitic melt is animportant factor in the compositional differentiation<strong>of</strong> the continental crust (Fyfe, 1973; Collins & Sawyer,1996; Brown & Rushmer, 2006) <strong>and</strong> the presence <strong>of</strong>melt in rocks pr<strong>of</strong>oundly influences their rheology(Arzi, 1978). The migration <strong>of</strong> melt through the crust iscontrolled by melt buoyancy <strong>and</strong> pressure gradientsresulting from the combination <strong>of</strong> gravity forces <strong>and</strong>deformation (Wickham, 1987; Sawyer, 1994). Thereare three major mechanisms controlling melt migrationthrough the continental crust: (i) diapirism resulting inupward motion <strong>of</strong> low-density magma through higherdensity rocks (Ch<strong>and</strong>rasekhar, 1961; Ramberg, 1981);(ii) dyking that describes melt migration by hydr<strong>of</strong>racturing<strong>of</strong> the host rock <strong>and</strong> transport <strong>of</strong> meltthrough narrow dykes (Lister & Kerr, 1991; Petford,1995); (iii) <strong>and</strong> migration <strong>of</strong> a melt through a network<strong>of</strong> interconnected pores during deformation or compaction<strong>of</strong> solid matrix (McKenzie, 1984; Wickham,1987).Brown & Solar (1998a) <strong>and</strong> Weinberg & Searle(1998) proposed that during active deformation meltmoves by pervasive flow <strong>and</strong> it is essentially pumpedthrough the system parallel to the principal finiteelongation in the form <strong>of</strong> foliation-parallel veins.Based on a number <strong>of</strong> field studies, pervasive meltmigration at outcrop scale controlled by regionaldeformation has been suggested by various authors(Collins & Sawyer, 1996; Brown & Solar, 1998b;V<strong>and</strong>erhaeghe, 1999; Marchildon & Brown, 2003).Ó 2007 Blackwell Publishing Ltd 29315


30 P. HASALOVÁ ET AL.These authors argued that magma intrudes pervasively,parallel to the main anisotropy represented byfoliation planes (John & Stu¨ nitz, 1997), fold hinges <strong>and</strong>interboudin partitions (Brown, 1994; Brown et al.,1995). It is also commonly observed that vein-likeleucosomes are injected into extensional structuresprovided the magma pressure is high enough (Wickham,1987; Lucas & St.Onge, 1995) or parallel to axialsurfaces <strong>of</strong> folds (Vernon & Paterson, 2001).Microscopic studies <strong>of</strong> natural rocks show orientations<strong>of</strong> former melt microstructures that are interpretedin terms <strong>of</strong> grain-scale channel networks(Sawyer, 2001). Melt migration pathways at the grainscale are commonly determined from distribution <strong>of</strong>melt films <strong>and</strong> pools (now glass) in experimentallyprepared samples or by distribution <strong>of</strong> minerals supposedto preserve the original melt topology in naturalrocks (Brown et al., 1999; Rosenberg & Riller, 2000;Rosenberg, 2001). The melt topology in experiments iscontrolled mainly by differential stress, confiningpressure <strong>and</strong> the amount <strong>of</strong> melt in the system(Rosenberg, 2001). At static conditions, the melttopology is characterized by equilibrium dihedral(wetting) angles at triple point junctions (Jurewicz &Watson, 1984; Laporte & Watson, 1995; Laporteet al., 1997; Cmı´ral et al., 1998; Walte et al., 2003) <strong>and</strong>the mobility <strong>of</strong> the melt remains very low, even if themelt phase forms an interconnected network alongtriple-junction grain edges at dihedral angles lowerthan 60° (Laporte & Watson, 1995; Connolly et al.,1997).Experimental studies on rock analogues to investigategrain-scale melt flow under laboratory conditionsshow that during contemporaneous melting <strong>and</strong>deformation melt connection allows the nucleation <strong>of</strong>shear b<strong>and</strong>s along which a melt is further segregated(Rosenberg & H<strong>and</strong>y, 2000, 2001; Barraud et al.,2001). Rosenberg (2001) reviewed the experimentaldata <strong>and</strong> concluded that the melt migration <strong>and</strong> meltflow direction are controlled by incremental shortening<strong>and</strong> melt pressure gradients between source <strong>and</strong> areas<strong>of</strong> melt accumulation.There have only been a few attempts to quantifymelt distribution in rocks using methods <strong>of</strong> quantitative<strong>and</strong> computer aided micro<strong>structural</strong> analysis(Dallain et al., 1999; Tanner, 1999; Marchildon &Brown, 2003). These studies commonly deal with graincontact frequency distributions, grain size evolution<strong>and</strong> orientation <strong>of</strong> former melt films (Dougan, 1983;McLellan, 1983; Rosenberg & Riller, 2000). However,modern quantitative micro<strong>structural</strong> analysis mayprovide further important information about: (i)reorganization <strong>of</strong> the rock structure associated withmelt migration in terms <strong>of</strong> grain contact distributions(Lexa et al., 2005); (ii) characterization <strong>of</strong> dynamic orstatic conditions <strong>of</strong> melt movement through rocksusing analysis <strong>of</strong> grain boundaries <strong>and</strong> shape orientations(Rosenberg & Riller, 2000; Marchildon & Brown,2002); <strong>and</strong> (iii) cooling or heating histories <strong>of</strong> rocksusing crystal size distribution (CSD) theory (Higgins,1998; Berger & Roselle, 2001).In this work, a sequence <strong>of</strong> deformed felsic rocks isstudied, ranging from high-grade b<strong>and</strong>ed orthogneissto fine-grained isotropic migmatite both at macro- <strong>and</strong>microscale using <strong>structural</strong>, petrographic <strong>and</strong> quantitativemicro<strong>structural</strong> <strong>analyses</strong>. It is shown that asequence <strong>of</strong> b<strong>and</strong>ed orthogneiss, stromatic, schlieren<strong>and</strong> nebulitic migmatites results from progressivedeformation in a crustal-scale shear zone in the presence<strong>of</strong> melt. The micro<strong>structural</strong> <strong>and</strong> fabric modificationsconnected with disintegration <strong>of</strong> parentalb<strong>and</strong>ed orthogneiss <strong>and</strong> development <strong>of</strong> r<strong>and</strong>om mineralmicrostructure are quantified. The relationships <strong>of</strong>the individual rocks types <strong>and</strong> the possible origin <strong>of</strong>this sequence are discussed in terms <strong>of</strong> deformation<strong>and</strong> migmatization <strong>of</strong> different protoliths, meltinfiltration from an external source or in situ melting <strong>of</strong>the same protolith during progressive deformation. Itis argued that b<strong>and</strong>ed orthogneiss <strong>and</strong> nebuliticmigmatites can be interpreted as end-members <strong>of</strong> acontinuous sequence resulting from melt infiltrationfrom an external source during deformation. Finally,the role <strong>of</strong> melt for activity <strong>of</strong> grain-scale deformationmechanisms <strong>and</strong> bulk rheological behaviour <strong>of</strong> crustalrocks during melt infiltration is discussed.GEOLOGICAL SETTINGThe Moldanubian zone represents the highest gradeunit <strong>of</strong> the Bohemian Massif <strong>and</strong> is interpreted as aninternal zone <strong>of</strong> the Variscan orogen developed duringthe Variscan convergence (Matte et al., 1990). TheMoldanubian zone is comprised essentially <strong>of</strong> highgradegneisses <strong>and</strong> migmatites containing relicts <strong>of</strong>high-pressure felsic granulites, eclogites <strong>and</strong> peridotitesthat are intercalated with mid-crustal rocks (Fig. 1a).Schulmann et al. (2005) described the <strong>structural</strong> <strong>and</strong>metamorphic evolution <strong>of</strong> high-grade crustal rocks <strong>of</strong>the so-called Gfo¨ hl Unit <strong>and</strong> <strong>of</strong> the adjacent middlecrustal units. For the mechanism <strong>of</strong> exhumation, theyproposed a model <strong>of</strong> vertical extrusion <strong>of</strong> orogeniclower crust <strong>and</strong> its lateral spreading in mid-crustallevels due to an indentation <strong>of</strong> the easterly Bruniapromontory. As a consequence <strong>of</strong> this process, thehigh-pressure rocks were thrust over adjacent middlecrustal units in conjunction with retrogression <strong>of</strong> originalmineral assemblages <strong>and</strong> partial melting <strong>of</strong> all therock types (Sˇtípska´ et al., 2004).Fig. 1. (a) Geological map <strong>of</strong> the eastern margin <strong>of</strong> the Bohemian Massif (modified after Schulmann et al., 2005) with the location <strong>of</strong>the study area (black rectangle). The upper right inset shows the general location <strong>of</strong> the Bohemian Massif within the EuropeanVariscides. (b) Schematic block diagram displaying the main <strong>structural</strong> features in the study area (modified after Schulmann et al.,2005). Dominant S 1 <strong>and</strong> S 2 fabrics with their orientations are shown. This block diagram is not vertically scaled.Ó 2007 Blackwell Publishing Ltd316


ORIGIN OF FELSIC MIGMATITES 31Ó 2007 Blackwell Publishing Ltd317


32 P. HASALOVÁ ET AL.The onset <strong>of</strong> the exhumation process is dated byzircon U–Pb ages <strong>of</strong> c. 340 Ma on felsic granulites,migmatites <strong>and</strong> mantle-derived syn-tectonicallyemplaced plutons (van Breemen et al., 1982; Kro¨ neret al., 1988; Holub et al., 1997; Schulmann et al.,2005). Tajčmanova´ et al. (2006) assigned metamorphicconditions <strong>of</strong> 840 °C at 18–19 kbar <strong>and</strong> 760–790 °C at10–13 kbar to relict steep granulitic fabrics whichoriginated by vertical extrusion <strong>of</strong> lower crust. Theseauthors also estimated the conditions <strong>of</strong> re-equilibration<strong>of</strong> granulites associated with horizontal spreadingstage to 720–770 °C <strong>and</strong> 4–4.5 kbar. High pressurerocks <strong>of</strong> the Gfo¨ hl Unit are accompanied by largebodies <strong>of</strong> biotite–sillimanite Gfo¨ hl orthogneiss spatiallyassociated with K-feldspar–sillimanite paragneisses<strong>and</strong> leucocratic migmatites for which P–Tconditions <strong>of</strong> 7–10 kbar <strong>and</strong> 750 °C were estimated byRacek et al. (2006).The area <strong>of</strong> this study is located at the easternmosttermination <strong>of</strong> the Gfo¨ hl Unit (Fig. 1a). The main rocktype is represented by the Gfo¨ hl orthogneiss withprotolith ages 488 ± 6 Ma (U–Pb SHRIMP: Friedlet al., 2004) including small bodies <strong>of</strong> amphibolite,granulite, eclogite, ultrabasic rock <strong>and</strong> paragneiss. TheGfo¨ hl orthogneiss shows different stages <strong>of</strong> migmatizationcharacterized by the assemblage <strong>of</strong> Kfs +Pl + Qtz + Bt ± Grt ± Sill. This migmatizedorthogneiss complex is heterogeneously deformed bytop to the NE shearing along a large-scale, gentlydipping shear zone (Schulmann et al., 1994). Consequently,the northern margin <strong>of</strong> this complex is thrustover a footwall comprised <strong>of</strong> the Na´meˇsˇtÕ granulitebody <strong>and</strong> Neoproterozoic metagranites <strong>of</strong> the northeasterncontinental margin (Urban, 1992).STRUCTURAL EVOLUTIONMesoscopic structuresTwo major deformation events are recognized. Thefirst event (D 1 ) is represented by a steep, west-dippinghigh-grade foliation S 1 (Fig. 1b). This fabric is preservedin b<strong>and</strong>ed orthogneisses (type I), as an alternation<strong>of</strong> recrystallized monomineralic K-feldspar,plagioclase <strong>and</strong> quartz layers, separated by b<strong>and</strong>s <strong>of</strong>biotite ± sillimanite (Fig. 2a). Lineation L 1 is locallymarked by alignment <strong>of</strong> biotite, sillimanite <strong>and</strong> byelongation <strong>of</strong> quartz <strong>and</strong> feldspar aggregates. Thisdeformation is attributed to an early stage <strong>of</strong> exhumation<strong>of</strong> the lower crust along a vertical channel(c)Type III(d)Type IV(b)Type II1 cm2 cm1cmd(a)Type Icba1cm1.0 mFig. 2. Schematic representation <strong>of</strong> the rock relationships at an outcrop scale <strong>and</strong> photographs <strong>of</strong> the individual rock types. (a) B<strong>and</strong>edorthogneiss (type I); (b) stromatic migmatite (type II); (c) schlieren migmatite (type III); <strong>and</strong> (d) nebulitic migmatite (type IV). Theposition <strong>of</strong> this outcrop in the study area is shown in Fig. 1b.Ó 2007 Blackwell Publishing Ltd318


ORIGIN OF FELSIC MIGMATITES 33during horizontal shortening <strong>of</strong> the thickened orogenicroot (Schulmann et al., 2005).The second deformation (D 2 ) is associated withreworking <strong>and</strong> folding <strong>of</strong> the S 1 compositional layeringin b<strong>and</strong>ed orthogneiss, so that S 1 is only preservedlocally in elongate relict domains (Fig. 2a). The D 2shearing is attributed to horizontal flow <strong>of</strong> hot lowercrust in a zone up to 10 km wide at a mid-crustal levelabove the Brunia promontory over distances <strong>of</strong> severaltens <strong>of</strong> kilometres (Schulmann et al., 2005). Relicdomains with gently folded S 1 fabric are surrounded byhighly deformed zones with tightly folded S 1 fabric(Fig. 2). The composite S 1)2 fabric is characterized bya b<strong>and</strong>ed structure with diffuse boundaries betweenpolymineralic K-feldspar- <strong>and</strong> plagioclase-richdomains similar to a stromatic migmatite structure(type II) (Fig. 2b). Locally the S 1 fabric is completelytransposed <strong>and</strong> a new S 2 foliation is dipping gently tothe SW (Fig. 1b). A sub-horizontal, gently S–SWplunging L 2 lineation (Fig. 1b) is mostly defined bypreferred orientation <strong>of</strong> sillimanite.Detailed field observations reveal that with ongoingdeformation the type II rock gradually pass into moreisotropic rock (type III) composed <strong>of</strong> K-feldspar–quartz <strong>and</strong> plagioclase–quartz aggregates (Fig. 2c) <strong>and</strong>containing rootless folds <strong>of</strong> the deformed S 1 fabric.This rock type alternates with irregular bodies orelongate lenses <strong>of</strong> fine-grained isotropic felsic rock(type IV, Fig. 2d), which in this region traditionallyhas been described as a nebulitic migmatite (Matějovska´,1974). Such a <strong>structural</strong> sequence originatedthrough intense D 2 deformation superimposed on anolder steep anisotropy <strong>and</strong> was identified in outcropscale along several sections. These observations aresupported by the existence <strong>of</strong> macroscopically visibleleucosomes or granitic veins that are also parallel to S 2<strong>and</strong> form isolated elongate pockets <strong>and</strong> lock-up shearb<strong>and</strong>s.This area has been extensively studied by Mateˇjovska´(1974) <strong>and</strong> Dudek et al. (1974) who used theclassical migmatite terminology <strong>of</strong> Mehnert (1971) forthe above-described rock types. These authors identifiedtype I rock as b<strong>and</strong>ed orthogneiss, rock type II asstromatic migmatite <strong>and</strong> rock type IV as nebuliticmigmatite. Rock type III resembles the schlierenmigmatite <strong>of</strong> Mehnert (1971). Because the Gfo¨ hl Unitis considered as one <strong>of</strong> the largest migmatitic terranes<strong>of</strong> the Variscan belt, the traditional migmatite terminologywas adapted to these rocks.MICROSTRUCTURAL OBSERVATIONSThe micro<strong>structural</strong> characteristics including grainsize, grain shape <strong>and</strong> grain boundary geometry werestudied in each <strong>of</strong> the four rock types <strong>and</strong> inK-feldspar- <strong>and</strong> plagioclase-rich domains. Thin sectionswere cut perpendicular to the foliation <strong>and</strong>parallel to L 2 lineation (XZ section). To discriminateK-feldspar from plagioclase, the thin sections werestained according to the method <strong>of</strong> Bailey & Stevens(1960).Type I: b<strong>and</strong>ed orthogneissThis rock type is a fine-grained orthogneiss with 0.25-to 2.0-mm-thick layers <strong>of</strong> recrystallized plagioclase(30 modal%), K-feldspar (40 modal%) <strong>and</strong> quartz(20 modal%), separated by discrete layers <strong>of</strong> biotite(10 modal%) commonly associated with minor sillimanite<strong>and</strong> garnet (Table 1, Fig. 3a).K-feldspar forms completely recrystallized aggregates(0.2–0.8 mm grain size) with straight grainboundaries locally meeting in triple point junctions at120° (Fig. 4a). Numerous rounded inclusions <strong>of</strong> quartz(0.05 mm) occur preferentially at triple points, alongplanar boundaries or in cores <strong>of</strong> feldspar (Fig. 4a).Plagioclase (An 10)20 ) is present in K-feldspar aggregatesas small interstitial grains or forms thin filmspreferentially tracing those K-feldspar boundaries thatare oriented at a high angle to the foliation (Fig. 5a).Rarely, tiny interstitial biotite is present in the K-feldspar-rich b<strong>and</strong>s.Plagioclase aggregates (0.2–0.5 mm) are composed<strong>of</strong> an equidimensional polygonal mosaic with straightboundaries, <strong>and</strong> minor interstitial quartz <strong>and</strong> biotite(Fig. 4b). The plagioclase grains show abundanttwinning <strong>and</strong> form a foam-like texture with a perfecttriple point network <strong>of</strong> grain boundaries. Plagioclaseexhibits normal zoning with homogeneous oligoclasecores (An 24)28 ) <strong>and</strong> more sodic (An 10)18 ), clear,2to10lm-thick rims at boundaries with K-feldspar.Plagioclase grain size continuously decreases from thecentre <strong>of</strong> an aggregate towards its borders. Quartzoccurs as small (0.01–0.05 mm) rounded inclusions orinterstitial grains, whereas K-feldspar exhibits characteristiccuspate shapes (Fig. 5b). Tiny biotite grains(0.1–0.5 mm in length; X Fe ¼ 0.42–0.48, Ti ¼ 0.2–0.27 p.f.u.) commonly occur along the plagioclaseboundaries that are sub-parallel to the foliation(Fig. 3a).Quartz ribbons 0.3–1.0 mm wide are composed <strong>of</strong>elongate grains with straight grain boundaries perpendicularto the ribbon margin (Fig. 3a). Quartz–feldspar boundaries are gently curved, with cusps thatpoint from feldspar to quartz. Biotite-rich layerscommonly show decussate microstructure, which is atextural equivalent <strong>of</strong> the foam-like texture <strong>of</strong> the felsicminerals (Vernon, 1976). Contacts between biotite<strong>and</strong>plagioclase-rich layers are marked by numerous(


34 P. HASALOVÁ ET AL.Table 1. Representative data for the quantitative textural analysis.B<strong>and</strong>ed orthogneiss Stromatic migmatite Schlieren migmatite Nebulitic migmatiteKfs domain P1 domain Kfs domain P1 domain Kfs domain P1 domainGrain size – Feret diameter (mm)Median Kfs 0.430 0.121 0.345 0.065 0.172 0.138 0.137P1 0.134 0.224 0.086 0.225 0.094 0.119 0.110Qtz 0.079 0.076 0.071 0.079 0.070 0.076 0.074Ql Kfs 0.120 0.065 0.194 0.042 0.103 0.082 0.085P1 0.103 0.095 0.055 0.158 0.061 0.084 0.074Qtz 0.046 0.051 0.044 0.050 0.046 0.047 0.039Q3 Kfs 0.630 0.170 0.556 0.101 0.263 0.211 0.237P1 0.257 0.373 0.161 0.350 0.172 0.164 0.160Qtz 0.105 0.114 0.119 0.127 0.105 0.121 0.127Q3 ) Q1 Kfs 0.510 0.105 0.362 0.059 0.161 0.129 0.152P1 0.154 0.278 0.107 0.192 0.111 0.080 0.086Qtz 0.059 0.063 0.075 0.077 0.060 0.074 0.089Crystal size distribution (CSD)N0. (mm )4 ) Kfs 0.0037 – 0.00487 – 0.053 – 0.2124P1 – 0.0303 – 0.0733 – 0.06812 0.1857Qtz 1.008 2.334 1.6448 3.73 2.093 0.6585 2.286Gt Kfs 0.347 – 0.286 – 0.148 – 0.1127P1 – 0.15731 – 0.1269 – 0.11 0.0689Qtz 0.0736 0.0569 0.0669 0.0547 0.0644 0.0813 0.0623Shape preferred orientation (SPO)Eigenvalue ratio (Rg) Kfs 1.48 1.42 1.42 1.32 1.21 1.13 1.15P1 1.17 1.42 1.1 1.23 1.21 1.14 1.13Qtz 1.25 1.47 1 1.24 1.1 1.33 1.32Aspect ratio (median) Kfs 1.66 1.69 1.6 1.5 1.59 1.6 1.55P1 1.51 1.6 1.59 1.44 1.65 1.5 1.61Qtz 1.5 1.5 1.46 1.5 1.5 1.5 1.49Bt 2.14 2.7 2 2.2 2.2 2.35 2.2Grain boundary preferred orientation (GBPO)Eigenvalue ratio (Rb) Kfs–Kfs 1.34 – 1.25 – 1.06 – 1.5Kfs–P1 1.15 1.18 1.09 1.13 1.14 1.13 1.12Kfs–Qtz 1.17 – 1.15 – 1.17 – 1.17P1–P1 – 1.18 – 1.15 – 1.15 1.36Modal proportion (%)Kfs 70–80 10 70–80 5 50 20–25 30P1 10 60 10 60 20 40 30Qtz 10–20 20 10–20 25 30 30 30Bt


ORIGIN OF FELSIC MIGMATITES 35(a)Type Iplagioclase (Fig. 4c). Biotite (10–15 modal%; X Fe ¼0.76–0.79, Ti ¼ 0.18–0.19 p.f.u.) is homogeneouslydispersed <strong>and</strong> is most prevalent in the plagioclase–quartz domains. Atoll-shaped garnet (0.05–0.25 mm insize; X Fe ¼ 0.96–0.97) appears inside the felsic aggregates,rather than along contacts with biotite.Type IV: nebulitic migmatiteThis type <strong>of</strong> rock is composed <strong>of</strong> almost equal amounts<strong>of</strong> plagioclase, K-feldspar <strong>and</strong> quartz, <strong>and</strong> containsminor biotite (X Fe ¼ 0.91–0.93, Ti ¼ 0.01–0.04 p.f.u.),sillimanite <strong>and</strong> garnet (X Fe ¼ 0.98–1.00) (Fig. 3d),with a weakly developed preferred orientation <strong>of</strong> thebiotite <strong>and</strong> sillimanite; modes are given in Table 1. K-feldspar (0.10–0.25 mm in size) occurs in the form <strong>of</strong>irregular grains embayed with quartz <strong>and</strong> plagioclase.Commonly, the intensity <strong>of</strong> quartz <strong>and</strong> plagioclaselobes correlates well with highly cuspate irregularforms <strong>of</strong> corroded relics <strong>of</strong> K-feldspar (Fig. 4e). Similarly,the relics <strong>of</strong> irregular plagioclase (0.05–0.15 mmin size; An 6)10 in the core <strong>and</strong> An 0)4 at the rim) showcuspate boundaries, but with curvature less pronouncedthan that <strong>of</strong> the corroded relics <strong>of</strong> K-feldspargrains. An important feature is the presence <strong>of</strong> newplagioclase (An 0)1 )–K-feldspar intergrowths embayingcorroded relics <strong>of</strong> K-feldspar grains (Fig. 4f). Quartz(0.04–0.07 mm) with highly lobate boundaries is uniformlydistributed in the rock. Biotite <strong>of</strong> low aspectratio shows highly corroded cuspate forms filled withquartz, K-feldspar <strong>and</strong> plagioclase.(b)(c)1 mm1 mmType II1 mmType IIISummary <strong>of</strong> modal changesModal composition <strong>of</strong> the feldspar aggregates in thetype II migmatite does not change significantly comparedwith the type I orthogneiss. However, the typeIII migmatite is characterized by an important increasein quartz content in feldspar domains (up to 30 modal%)associated with a slight increase in interstitialplagioclase in K-feldspar-rich domains <strong>and</strong> K-feldsparin plagioclase-rich domains. The proportions <strong>of</strong> thefelsic minerals are equal in the type IV migmatite.(d)KfsPlQtzBt,Sill,Grt1 mmType IVFig. 3. Representative digitalized microstructures (XZ sections)for individual textural types (note differences in scales whenmaking comparisons). (a) B<strong>and</strong>ed orthogneiss (type I) with distinctmonomineralic layers composed <strong>of</strong> a polygonal mosaic <strong>of</strong>well-equilibrated plagioclase, K-feldspar <strong>and</strong> quartz polycrystallineribbons separated by discrete layers <strong>of</strong> biotite ± sillimanite± garnet (sample PH60/B). (b) Stromatic migmatite (type II)composed <strong>of</strong> K-feldspar-rich, plagioclase-rich <strong>and</strong> quartz-richaggregates separated by relicts <strong>of</strong> biotite ± sillimanite-rich layers(sample PH60/A). (c) Schlieren migmatite (type III) showingalternation <strong>of</strong> K-feldspar- <strong>and</strong> plagioclase-rich domains interpretedto correspond to an original spatial distribution (K-feldspardomain is shown, sample PH90). (d) Isotropic nebuliticmigmatite without any gneissosity (type IV) composed <strong>of</strong> equalamounts <strong>of</strong> K-feldspar, plagioclase <strong>and</strong> quartz (sample PH59/D).Ó 2007 Blackwell Publishing Ltd321


36 P. HASALOVÁ ET AL.Evidence <strong>of</strong> meltingSawyer (1999, 2001) summarized criteria for recognition<strong>of</strong> former melt at grain scale in metamorphicrocks. The three most important features are: (i) mineralpseudomorphs after thin melt films along crystalfaces, a feature typically observed in melting experimentsunder dynamic conditions (Jin et al., 1994); (ii)rounded <strong>and</strong> corroded reactant minerals embayed bysurrounding mineral pseudomorphs after melt (Bu¨ sch(a)(b)KfsKfsQPlBt0.5 mm0.2 mm(c)Qtz(d)PlPlKfsKfsBtPlKfsQtzKfsQtzKfsPl0.5 mmQtz100 µm(e)(f)PlQtzKfsKfsQtzQtzKfs-PlintergrowthPl100 µmBtPlKfsPl0.2 mmQtz Pl Myrmekites Irregular embayments <strong>of</strong> relict feldsparÓ 2007 Blackwell Publishing Ltd322


ORIGIN OF FELSIC MIGMATITES 37(a)Pl(An25)(b)QtzQtzQtzQtzAn 10-20An 10-20An 10-20QtzQtzKfsPl(An25)PlKfsQtzKfsKfsQtzQtzQtz0.5 mm100 µm(c)Pl core(An 12-16 )QtzKfs(d)PlKfs(An 15 )(An 0-4)Pl rim(An 1-4 )KfsPl rim(An 1-4)Pl(An 12-16 )(An 1-4 )QtzPlrelict grain(An 12-16)100 µmKfsPlQtzPl(An15)QtzPlQtzKfsQtz200 µmFig. 5. SEM backscatter images showing the inferred former melt topology (note differences in scales when making comparisons). (a)Type I b<strong>and</strong>ed orthogneiss: interstitial plagioclase (An 10)20 ), representing the plagioclase component crystallized from the anatecticmelt (grey arrow), tracing the K-feldspar boundaries sub-perpendicular to the foliation (sample PH60/B). Black arrows show smallrounded quartz grains crystallized along feldspar boundaries. (b) Type I b<strong>and</strong>ed orthogneiss: inferred former melt pools with cuspatemargins in a plagioclase b<strong>and</strong> (sample PH60/B). The former melt has crystallized to K-feldspar (cuspate melt pools), plagioclase(growing on the old plagioclase grains) <strong>and</strong> quartz (forming small rounded grains along the feldspar boundaries (black arrow)). (c)Type III schlieren migmatite: more developed interstitial plagioclase (grey arrow) with normal zoning (core ¼ An 12)16 ; rim ¼ An 1)4 )<strong>and</strong> distinct albite rims (An 1)4 ) on relict feldspar grains (white arrow) (sample PH90). The interstitial plagioclase is not in opticalcontinuity with any residual plagioclase grains adjacent to it <strong>and</strong> does not show any preferred orientation, in contrast to plagioclase intypes I <strong>and</strong> II. (d) Type III schlieren migmatite: new plagioclase inferred to have crystallized from melt (growing on an old plagioclasegrain in the form <strong>of</strong> the discrete albite rims (white arrow)) <strong>and</strong> quartz grains that resorb relict K-feldspar grains (sample PH14/D).Fig. 4. Photomicrographs showing characteristic textures <strong>of</strong> the rock sequence (note differences in scales when making comparisons).(a) Type I b<strong>and</strong>ed orthogneiss: recrystallized K-feldspar aggregate with straight grain boundaries <strong>and</strong> numerous smaller roundedquartz grains (white triangles) along the boundaries or in the cores <strong>of</strong> feldspar (sample PH60/B). (b) Type I b<strong>and</strong>ed orthogneiss: welldevelopedplagioclase polygonal foam-like texture with straight grain boundaries, interstitial quartz (white triangles) <strong>and</strong> biotite(sample PH60/B). (c) Type III schlieren migmatite: typical microstructure with irregularly shaped feldspar <strong>and</strong> quartz grains withhighly lobate boundaries. Myrmekitic aggregates commonly develop along the K-feldspar boundaries (black arrow). New smallinterstitial plagioclase (grey triangles), K-feldspar <strong>and</strong> quartz (white triangles) grains trace almost all the relict feldspar boundaries.Interstitial quartz forms preferentially rounded shapes different from plagioclase which forms thin elongated grains/films coatingK-feldspar boundaries (sample PH90). Such a microstructure is typical also for the type IV. (d) Type III schlieren migmatite: irregularcuspate K-feldspar grain embayed with newly crystallized quartz <strong>and</strong> plagioclase (sample PH90). (e) Type IV nebulitic migmatite:corroded relics <strong>of</strong> K-feldspar grains (sample PH59/D). (f) Type III nebulitic migmatite: plagioclase-K-feldspar intergrowths embayingrelict K-feldspar grain (sample PH14/D). White arrows in (d), (e) <strong>and</strong> (f) point to irregular embayments <strong>of</strong> relict K-feldspar originatedthrough resorption <strong>of</strong> old K-feldspar grains by newly crystallized material.Ó 2007 Blackwell Publishing Ltd323


38 P. HASALOVÁ ET AL.PlQtz70% 30%B<strong>and</strong>ed orthogneiss (Type I)Stromatic migmatite (Type II)Schlieren migmatite (Type III)Nebulitic migmatite (Type IV)50%30%et al., 1974); <strong>and</strong> (iii) cuspate <strong>and</strong> lobate areas inferredto represent pools <strong>of</strong> crystallized melt (Jurewicz &Watson, 1984).The former presence <strong>of</strong> melt at grain scale was inferredfrom the following microstructures (Figs 4 & 5).(i) Plagioclase films between adjacent K-feldspargrains, inferred to represent a plagioclase componentcrystallized from melt (Fig. 5a, c). This plagioclase ischaracterized by more albitic composition <strong>and</strong> by differenttopology compared with original grains. (ii) Pl–Kfs–Kfs <strong>and</strong> Kfs–Kfs–Pl dihedral angles commonlylower than 30° (Fig. 5a, c), as observed in granitic meltcrystallized under experimental conditions (e.g.Laporte et al., 1997). (iii) Cuspate K-feldspar pools inplagioclase aggregates (Fig. 5b), inferred to represent aK-feldspar component crystallized from melt (Jurewicz& Watson, 1984; Sawyer, 1999, 2001). (iv) Normalzoning <strong>of</strong> plagioclase from An 10)30 to An 0)15 (Sawyer,1998; Marchildon & Brown, 2001) lining K-feldsparboundaries (Fig. 5c, d). An important feature is thepreferential orientation <strong>of</strong> plagioclase films coating K-feldspar boundaries in type I orthogneiss <strong>and</strong> type IImigmatite sub-perpendicular to the foliation (Fig. 5a),in contrast to the types III <strong>and</strong> IV migmatites, wherethese films are wider <strong>and</strong> do not show any opticallyvisible preferred orientation. Bulbous myrmekite(Fig. 4d) <strong>and</strong> new highly irregular lobate grains thatovergrow partially resorbed corroded feldspar grains(e.g. Fig. 4c) are similar to microstructures describedas typical <strong>of</strong> minerals reacting with melt (Mehnertet al., 1973; Bu¨ sch et al., 1974; McLellan, 1983).QUANTITATIVE TEXTURAL ANALYSISKfsFig. 6. Modal changes in both plagioclase (open symbols) <strong>and</strong>K-feldspar (closed symbols) aggregates in different rock typesplotted in a quartz–plagioclase–K-feldspar triangle. Arroweddashed lines indicate evolutionary trend from type I b<strong>and</strong>edorthogneiss to type IV nebulitic migmatite.The quantitative analysis <strong>of</strong> texture is based onstatistical evaluation <strong>of</strong> grain size distributions(Kretz, 1966, 1994; Ashworth, 1976; Ashworth &McLellan, 1985; Cashman & Ferry, 1988; Cashman& Marsh, 1988; Higgins, 1998; Berger & Roselle,2001), spatial distribution <strong>of</strong> minerals <strong>and</strong> GBPOs(Panozzo, 1983), <strong>and</strong> grain contact frequencies(Flinn, 1969; Kretz, 1969; McLellan, 1983; Kruse &Stu¨ nitz, 1999). In simple chemical systems, thesetextural parameters are more sensitive to changes <strong>of</strong>physical conditions than compositional characteristics.This is due to the high activation energies <strong>of</strong>chemical reactions needed to produce new crystalgrowth compared with the small amount <strong>of</strong> latticestrain energy <strong>and</strong> grain boundary energy required todrive recrystallization processes (Spry, 1969; Stu¨ nitz,1998).In this study, the textures <strong>of</strong> three samples wereanalysed from each rock type, <strong>and</strong> in each samplemore than 1000 grains were evaluated in thin section.Due to significant textural variations, the individualK-feldspar-rich <strong>and</strong> plagioclase-rich domains wereanalysed separately. Maps <strong>of</strong> grains with full topologywere manually traced into the ESRI ArcViewDesktop GIS environment <strong>and</strong> grain boundaries weregenerated using the ArcView PolyLX extension (Lexa,2003). The ÔshapefilesÕ <strong>of</strong> individual digitalized thinsections are attached in Appendix S1 (Supplementarymaterial). Analysis <strong>of</strong> grain size, CSD, grain shapepreferred orientation (SPO), grain boundary preferredorientation (GBPO) <strong>and</strong> grain contact frequencieswere obtained using the MATLAB TM PolyLX toolbox(Lexa, 2003; http://petrol.natur.cuni.cz/ ondro/). Thegrain sizes <strong>of</strong> the minerals were evaluated in terms<strong>of</strong> Feret diameter (diameter <strong>of</strong> a circle having thesame area as the grain). Two methods were used todetermine the grain SPO: (1) mean directions usingcircular statistics; <strong>and</strong> (2) eigenvalue analysis <strong>of</strong>Scheidegger’s bulk orientation tensor calculated fromindividual long axes weighted by grain size (Lexaet al., 2005), where degree <strong>of</strong> SPO is expressed as theeigenvalues ratio Rg. GBPO was assessed by similartechniques, but the bulk orientation tensor is formedfrom the decomposed grain boundaries betweenchosen phases (Lexa et al., 2005) <strong>and</strong> the degree <strong>of</strong>GBPO is expressed as the eigenvalues ratio Rb. Graincontact frequency, used to examine statistical deviationfrom a r<strong>and</strong>om spatial distribution <strong>of</strong> contactrelations between the individual minerals, was evaluatedin a manner similar to the method <strong>of</strong> Kretz(1969, 1994), except that contact frequencies wereobtained directly from grain map topologies instead<strong>of</strong> using line intercepts.Results <strong>of</strong> the quantitative micro<strong>structural</strong> <strong>analyses</strong>show an evolutionary trend from the b<strong>and</strong>ed orthogneiss,through the migmatite types II <strong>and</strong> III to thenebulitic migmatite. Therefore, in the following sectionsthe rock types are discussed as a sequence inwhich the type I orthogneiss <strong>and</strong> type IV nebuliticmigmatite are considered to be end-members <strong>of</strong> acontinuous micro<strong>structural</strong> evolution.Ó 2007 Blackwell Publishing Ltd324


ORIGIN OF FELSIC MIGMATITES 39Grain size analysisThe CSD is an important tool to estimate residencetime <strong>of</strong> magmas in magma chambers, cooling rates inrapidly quenched lavas, as well as to quantify texturesrelated to phenocrysts accumulation <strong>and</strong> fractionation(Cashman & Marsh, 1988; Marsh, 1988; Higgins,1998). In metamorphic petrology, CSD is used toobtain quantitative information concerning crystalnucleation <strong>and</strong> growth rates <strong>and</strong> nucleation density<strong>and</strong>/or annealing (R<strong>and</strong>olph & Larson, 1971; Cashman& Ferry, 1988; Carlson, 1989; Waters & Lovegrove,2002). Hickey & Bell (1996) proposed thatduring dynamic recrystallization decreasing strain rateto temperature ratio (_e/T) leads to decrease in the ratio<strong>of</strong> nucleation <strong>and</strong> growth rate (N/G) <strong>and</strong> development<strong>of</strong> coarser grain size, whereas increasing _e/T leads toincreasing N/G <strong>and</strong> therefore to grain size decrease.This hypothesis is well documented in experimentalstudies with steel alloys (Sakai & Jonas, 1984) supportedby Azpiroz & Ferna´ndez (2003) <strong>and</strong> Lexa et al.(2005) in naturally deformed rocks. These authorsevaluate the role <strong>of</strong> recrystallization mechanisms on N/G ratio <strong>of</strong> the CSD. The CSD is commonly used informerly partially molten rocks to evaluate combinedprocess <strong>of</strong> resorption <strong>and</strong> grain size decrease in reactingphases in mesosome <strong>and</strong> nucleation <strong>and</strong> graingrowth <strong>and</strong> coarsening <strong>of</strong> minerals crystallizing inleucosomes (Dougan, 1983; McLellan, 1983; Ashworth& McLellan, 1985; Dallain et al., 1999). Because theprocesses controlling grain size distributions in thecrystallization <strong>of</strong> partially molten rocks are complex<strong>and</strong> interpretations uncertain, CSD has only rarelybeen used to describe textural evolution <strong>of</strong> migmatites(Berger & Roselle, 2001). In this work, the CSDmethods are used as a practical approach to parameterizegrain size frequency histograms <strong>and</strong> visualizetheir trends in a simple manner.Grain size statistics were evaluated for the four rocktypes for plagioclase, K-feldspar <strong>and</strong> quartz <strong>and</strong> theresults are presented in the form <strong>of</strong> average grain size,expressed as a median value <strong>of</strong> the Feret diameter, <strong>and</strong>grain size range expressed as the difference between thethird <strong>and</strong> first quartiles instead <strong>of</strong> st<strong>and</strong>ard deviationbecause <strong>of</strong> the log-normal distribution <strong>of</strong> measureddata (Fig. 7a, Table 1). The results are also summarizedas CSD curves (plot <strong>of</strong> logarithms <strong>of</strong> populationdensity against crystal size) that were constructed usingthe method <strong>of</strong> Peterson (1996); values <strong>of</strong> the zero-sizeintercept (N 0 – population density interpreted as theratio <strong>of</strong> nucleation rate to growth rate) <strong>and</strong> negativeinverse <strong>of</strong> slope (Gt interpreted as a function <strong>of</strong> growthrate) <strong>of</strong> the linear parts <strong>of</strong> the CSD curves are plottedin Fig. 7b, c.Both plagioclase <strong>and</strong> K-feldspar in the type Iorthogneiss are characterized by log-normal grain sizedistribution exhibiting average grain size <strong>of</strong> 0.2 <strong>and</strong>0.2–0.5 mm respectively. Interstitial quartz yields significantlysmaller average grain size <strong>of</strong> 0.05–0.1 mm inboth domains. Quartz grains from polycrystalline ribbonswere not evaluated statistically but their grain size<strong>of</strong> 0.5–2.0 mm was estimated using an optical microscope.The grain size <strong>of</strong> new interstitial plagioclase inthe K-feldspar aggregates is close to 0.1 mm. The grainsize distributions from type I orthogneiss to type II,type III <strong>and</strong> type IV migmatites are characterized bythe following features. The average grain size <strong>of</strong> plagioclase<strong>and</strong> K-feldspar decreases compared with typeI orthogneiss (Fig. 7a, Table 1). This is accompaniedby a continuous decrease in grain size range for bothfeldspars. The interstitial quartz grain size remainsfairly constant throughout all the stages <strong>of</strong> texturalevolution, ranging between 0.05 <strong>and</strong> 0.1 mm, beinglarger in the K-feldspar than in the plagioclasedomains (Fig. 7a). The grain size <strong>of</strong> minor plagioclasein the K-feldspar domains shows a bimodal distributionthat is attributed to the presence <strong>of</strong> small newlynucleated grains (0.06–0.1 mm) <strong>and</strong> to larger plagioclasegrains (0.2 mm) already present in the feldsparaggregates.The CSD <strong>of</strong> plagioclase indicate continuous increasein N 0 (nucleation density) values coupled with adecrease in Gt (growth rate) values from type Iorthogneiss towards type IV migmatite (Fig. 7b,Table 1). By contrast, K-feldspar shows a decrease inGt values from type I orthogneiss to type III migmatitewithout significant increase in N 0 values, which remainvery low. From type III to type IV migmatite a dramaticincrease in N 0 values is observed for K-feldsparat almost constant Gt values (Fig. 7c). This evolutionis clearly shown by steepening <strong>of</strong> the slopes <strong>of</strong> the CSDcurves accompanied by increase in their upper interceptwith the ordinate axis (insets in Fig. 7b, c).Grain shapes <strong>and</strong> grain shape preferred orientationGrain shape or grain aspect ratio together with grainSPO <strong>analyses</strong> provide important information aboutdeformation during or after leucosome formation(Mehnert, 1971; McLellan, 1983) or about degree <strong>of</strong>inheritance <strong>of</strong> original anisotropy (Ashworth, 1979).Measurements <strong>of</strong> preferred orientations <strong>of</strong> inferredmelt-filled grain boundaries in rocks give insights intoprocesses <strong>of</strong> melt draining <strong>and</strong> melt transfer (Rosenberg& H<strong>and</strong>y, 2000, 2001; Sawyer, 2001).Grain shape <strong>and</strong> SPO statistics were evaluated in allthe textural types for plagioclase, K-feldspar, quartz<strong>and</strong> biotite. The results <strong>of</strong> SPO statistic are summarizedin a boxplot-type diagram, where the axial ratios<strong>of</strong> the individual minerals are plotted against bulk SPO(Lexa et al., 2005) for the corresponding minerals(Fig. 8).Aspect ratios for both K-feldspar <strong>and</strong> plagioclaseshow small median values ranging from 1.5 to 1.7throughout the whole micro<strong>structural</strong> sequence(Fig. 8, Table 1). Quartz exhibits slightly smaller <strong>and</strong>stable aspect ratio close to 1.5. An important feature isthe continuous decrease in SPO <strong>of</strong> K-feldspar <strong>and</strong>Ó 2007 Blackwell Publishing Ltd325


40 P. HASALOVÁ ET AL.(a)(b)(c)Fig. 7. Grain size statistics <strong>and</strong> CSD evolution for the rock sequence. (a) Calculated average grain size (median value <strong>of</strong> the Feretdiameter) <strong>and</strong> range (difference <strong>of</strong> third <strong>and</strong> first quartiles) for plagioclase, K-feldspar <strong>and</strong> quartz. (b,c) Plots <strong>of</strong> crystal size distributionparameters N 0 (corresponding to the nucleation density per size per volume) <strong>and</strong> Gt (non-dimensional value dependent on the growthrate) with examples <strong>of</strong> linearized CSD curves (upper right insets) used for Gt <strong>and</strong> N 0 estimates. (b) Plagioclase, (c) K-feldspar. The CSDcurves show single lines <strong>of</strong> four representative samples corresponding to the individual rock types.plagioclase from type I orthogneiss to type IV nebuliticmigmatite (Fig. 8, Table 1). Rose diagrams for therock types I, II <strong>and</strong> III show that K-feldspar <strong>and</strong>plagioclase (Fig. 8b) have weakly inclined SPO withrespect to the aggregate elongation direction at anangle <strong>of</strong> 15°–30°. Biotite shows a high aspect ratio fortype I orthogneiss (Table 1) <strong>and</strong> strong SPO parallelwith mesoscopic foliation for the types I, II <strong>and</strong> IIImigmatites. In the type IV migmatite, biotite aspectratio <strong>and</strong> preferred orientation are lower <strong>and</strong> the latterparameter shows bimodal distribution with one maximumsub-parallel to the main foliation <strong>and</strong> a secondone almost perpendicular to it. Interstitial K-feldspar,plagioclase <strong>and</strong> quartz exhibit always small aspectratio <strong>and</strong> weakly developed SPO maxima at an angle<strong>of</strong> 40°–60° to the foliation for types I, II <strong>and</strong> III. Theexception is type IV migmatite, where, in similarfashion to biotite, the interstitial plagioclase shows twomaxima, one sub-parallel <strong>and</strong> one perpendicular to thefoliation.Grain contact frequency analysis <strong>and</strong> grain boundarypreferred orientationThe grain contact frequency method (Kretz, 1969)allows an examination <strong>of</strong> the statistical deviationfrom the hypothesis <strong>of</strong> r<strong>and</strong>om distribution <strong>of</strong> phasesin rocks. In r<strong>and</strong>om distribution, the number <strong>of</strong>Ó 2007 Blackwell Publishing Ltd326


ORIGIN OF FELSIC MIGMATITES 41(a)(b)Fig. 8. Plot <strong>of</strong> grain shape preferred orientation (SPO) <strong>of</strong> K-feldspar (a) <strong>and</strong> plagioclase (b). The results are summarized in a box plot<strong>of</strong> aspect ratios (characterizing the shape <strong>of</strong> grains) v. eigenvalue ratios (showing the degree <strong>of</strong> preferred orientation). Individual boxesshow median, <strong>and</strong> first <strong>and</strong> third quartiles <strong>of</strong> the aspect ratio. The whiskers represent a statistical estimate <strong>of</strong> the data range whereoutliers are not plotted. Representative rose diagrams for individual rock types show maxima orientation in respect to the aggregateelongation direction. The degree <strong>of</strong> shading corresponds to the individual rock types.contacts <strong>of</strong> given phases depend only on the totalnumber <strong>of</strong> grains <strong>of</strong> each phase present. There aretwo possible deviations from r<strong>and</strong>om distribution: (i)aggregate distribution, where grains <strong>of</strong> the samephase tend to occur in aggregates in which contactsbetween grains <strong>of</strong> the same phase (like–like contacts)predominate; <strong>and</strong> (ii) regular distribution, where thegrains tend to occur in a regular (chessboard-like)pattern in which contacts between different phases(unlike contacts) are more common. McLellan (1983)reviewed processes responsible for different types <strong>of</strong>grain distributions. A r<strong>and</strong>om distribution shouldtheoretically develop during rapid quenching <strong>of</strong> graniticmelt, whereas regular distribution commonly isinterpreted as resulting from extensive solid-stateannealing under very high temperatures (Flinn, 1969;Vernon, 1976; McLellan, 1983; Lexa et al., 2005).These interpretations are based on the assumption <strong>of</strong>reducing surface energy (Seng, 1936; DeVore, 1959)by elimination <strong>of</strong> high-energy contacts (commonlynon-coherent like–like contacts) either by reduction <strong>of</strong>grain boundary area or by nucleation <strong>and</strong> growth <strong>of</strong>new phases along such a boundary (Kim & Rohrer,2004). In addition, Kruse & Stu¨ nitz (1999) <strong>and</strong> Baratouxet al. (2005) proposed that the regular distributionwas induced by mechanical mixing <strong>and</strong>heterogeneous nucleation. According to Vernon(1976) <strong>and</strong> McLellan (1983), an aggregate distributionresults from a solid-state differentiation associatedmostly with dynamic recrystallization where development<strong>of</strong> monomineralic layers results from unevenefficiency <strong>of</strong> deformation mechanisms simultaneouslyoperating in different phases (Jordan, 1988).Grain contact frequency <strong>and</strong> the GBPO were evaluatedfor K-feldspar, plagioclase <strong>and</strong> quartz in all thetextural types over the full digitized area <strong>of</strong> individualK-feldspar <strong>and</strong> plagioclase domains. The results arepresented in Fig. 9, where the v-valuev ¼ Observed pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiExpectedExpectedor deviation from the r<strong>and</strong>om distribution is plottedagainst the ratio <strong>of</strong> eigenvalues <strong>of</strong> the orientationtensor or the degree <strong>of</strong> GBPO (Lexa et al., 2005).Values <strong>of</strong> expected frequencies are estimated usingLafeber’s method <strong>of</strong> testing for r<strong>and</strong>omness (Lafeber,1963; Kretz, 1969). This diagram <strong>of</strong>fers a simple visualevaluation <strong>of</strong> the relationship between degree <strong>of</strong> deviationfrom expected r<strong>and</strong>om distribution <strong>of</strong> graincontacts <strong>and</strong> GBPOs <strong>of</strong> like–like <strong>and</strong> unlike boundaries.The type I orthogneiss is characterized by a relativelysmall proportion <strong>of</strong> like–like K-feldspar <strong>and</strong>plagioclase contacts indicating a weak regular distribution(slightly negative v-values for like–like <strong>and</strong>positive v-values for unlike contacts; Fig. 9), despitea macroscopically b<strong>and</strong>ed texture in which a strongaggregate distribution should be observed. This featureis attributed to a great proportion <strong>of</strong> minorinterstitial grains (Qtz, Bt, Kfs <strong>and</strong> Pl) lining theK-feldspar <strong>and</strong> the plagioclase boundaries. Additionally,the number <strong>of</strong> like–like K-feldspar <strong>and</strong>plagioclase contacts continuously decreases from typeI orthogneiss to type IV migmatite, whereas thenumber <strong>of</strong> Pl–Kfs, Kfs–Qtz <strong>and</strong> Pl–Qtz unlikeÓ 2007 Blackwell Publishing Ltd327


42 P. HASALOVÁ ET AL.Fig. 9. Grain boundary statistics plotted as the deviation from a r<strong>and</strong>om spatial distribution (grain contact frequency) v. degree <strong>of</strong>grain boundary preferred orientation (GBPO). For details see text. The degree <strong>of</strong> shading corresponds to the individual rock types.contact continuously increases (negative like–like v-values <strong>and</strong> positive unlike v-values) (Fig. 9). This isin a good accordance with the increasing amount <strong>of</strong>interstitial phases towards the type IV migmatite.Quartz exhibits the same strong regular distributionfrom the type I orthogneiss to the type IV migmatitein both feldspar domains.In K-feldspar-rich aggregates, the degree <strong>of</strong> GBPO<strong>of</strong> the K-feldspar like–like boundaries slightly decreasesfrom type I orthogneiss to type III migmatite,whereas type IV migmatite is characterized by an increasein the degree <strong>of</strong> K-feldspar like–like GBPO(Fig. 9, Table 1). The GBPO <strong>of</strong> plagioclase–plagioclaseboundaries in the plagioclase-rich aggregates issimilar to the evolution <strong>of</strong> K-feldspar like–likeboundaries. The GBPO <strong>of</strong> the K-feldspar–quartzboundaries as well as those <strong>of</strong> K-feldspar–plagioclaseboundaries are weak, <strong>and</strong> decrease throughout thetextural evolution (Fig. 9, Table 1).MINERAL FABRICIn rocks deformed in the presence <strong>of</strong> melt, the textures<strong>of</strong> quartz <strong>and</strong> feldspar can be used to evaluate thedeformation mechanisms <strong>of</strong> the solid fraction as wellas the deformation <strong>of</strong> crystallizing intragranular melt(Za´vada et al., 2007). The mineral fabrics <strong>of</strong> ferromagnesianphases can be indirectly assessed usinganisotropy <strong>of</strong> magnetic susceptibility (AMS). TheAMS method has been recently used to determine thedegree <strong>of</strong> susceptibility, shape <strong>of</strong> the fabric ellipsoid<strong>and</strong> relative contribution <strong>of</strong> ferro- <strong>and</strong> para-magneticminerals to the bulk fabric in migmatites (Ferre´ et al.,2003, 2004).Anisotropy <strong>of</strong> magnetic susceptibilityTypes III <strong>and</strong> IV migmatites are macroscopically closeto isotropic, so that the mineral alignment defined bythe orientation <strong>of</strong> dispersed biotite is poorly defined(Fig. 2c, d). To better characterize the fabric, the AMSmethod was used to determine the internal fabric <strong>of</strong>these rocks. Oriented samples were collected using aportable drill at four sampling sites covering a sectionacross the well-defined <strong>structural</strong> sequence. The AMSdata were statistically evaluated using the Anis<strong>of</strong>ts<strong>of</strong>tware package (Jelínek, 1978; Hrouda et al., 1990).The low values <strong>of</strong> mean susceptibility (


ORIGIN OF FELSIC MIGMATITES 43(a)(b)Fig. 10. Plots to show the anisotropy <strong>of</strong> magnetic susceptibility (AMS). (a) P¢–T plot, where the P¢ parameter represents the degree <strong>of</strong>magnetic anisotropy <strong>and</strong> T is a shape parameter that describes the shape <strong>of</strong> the ellipsoid <strong>of</strong> magnetic susceptibility. T can take eitherpositive values (T > 0), characteristic for a planar fabric, or negative values (T < 0), typical for a linear fabric. Dashed ellipses showtwo distinct datasets. For comparison, data obtained by Schulmann K., Edel J.-B., Hasalova´ P., Lexa O., Jezˇek J. & Cosgrove J. W.(unpublished data) <strong>and</strong> Bouchez (1997) are shown. (b) Magnetic foliation (circles), plotted as the minimal susceptibility direction (K 3 ),perpendicular to the magnetic foliation, <strong>and</strong> magnetic lineation (squares), plotted as the maximal susceptibility direction (K 1 ).Lattice preferred orientationTo underst<strong>and</strong> the deformation behaviour <strong>of</strong> individualphases, we measured <strong>and</strong> evaluated statistically thelattice preferred orientation (LPO) <strong>of</strong> aggregate grains(Pl, Kfs <strong>and</strong> Qtz) <strong>and</strong> grains apparently crystallizedfrom melt (Pl, Kfs <strong>and</strong> Qtz) separately (Fig. 12g, h).The LPO <strong>of</strong> quartz, plagioclase <strong>and</strong> K-feldspar weremeasured on a scanning electron microscope Cam-Scan3200 in the Czech Geological Survey using theelectron back-scattered diffraction technique (EBSD)<strong>and</strong> HKL technology (Adams et al., 1993; Bascouet al., 2001). Diffraction patterns were acquired at20 kV <strong>of</strong> accelerating voltage, 5 nA <strong>of</strong> probe current<strong>and</strong> working distance <strong>of</strong> 33 mm from the thin sectionprepared from the <strong>structural</strong> XZ plane. The procedurewas carried out manually due to small differences indiffraction patterns. The chemistry <strong>and</strong> orientation <strong>of</strong>individual grains was controlled using a forescatterdetector with combination <strong>of</strong> orientation <strong>and</strong> chemicalcontrast. Thus, each individual grain is represented byonly one orientation measurement. The resulting polefigures are presented as lower hemisphere equal-areaprojections in which the trace <strong>of</strong> foliation is orientedalong the equator <strong>and</strong> the stretching lineation is in theE–W direction.Old quartz grains in ribbons <strong>of</strong> the type I orthogneissshow c-axes distributed in weak sub-maxima arrangedalong weakly developed small circles close tothe S 1 foliation trace. The most intense sub-maximaare developed close to the lineation direction. This type<strong>of</strong> c-axis pattern may indicate preferential prism Æcæslip-system activity <strong>and</strong> dominantly coaxial deformation.The c-axes <strong>of</strong> large quartz grains in types II, III<strong>and</strong> IV migmatites reveal strong maxima either parallelto the S 2 foliation pole or close to the centre <strong>of</strong> thediagram. These c-axis patterns indicate mainly activity<strong>of</strong> basal Æaæ or rhomb Æa + cæ slip-systems <strong>and</strong> lessfrequently prism Æaæ slip (Fig. 11a). Towards types III<strong>and</strong> IV migmatites, the LPO <strong>of</strong> the matrix quartz becameless well developed, preserving activity <strong>of</strong> thesame slip-systems as in the previous micro<strong>structural</strong>types (Fig. 11a). New quartz grains crystallized frommelt in type I orthogneiss, <strong>and</strong> type II migmatite showvery weak LPO <strong>and</strong> nearly r<strong>and</strong>om distribution <strong>of</strong> allquartz axes (Fig. 11b). Whereas old grains show progressiveweakening <strong>of</strong> the LPO from type II to type IVmigmatite, the new <strong>and</strong> r<strong>and</strong>omly crystallized grainstends to develop weak crystal preferred orientationduring the same micro<strong>structural</strong> evolution from type IIto type IV migmatite (Fig. 11). It is difficult to distinguishold from new quartz grains in the type IV rock<strong>and</strong> therefore the LPO <strong>of</strong> quartz in this microstructurelinks LPO evolution between old <strong>and</strong> new grains in thefinal micro<strong>structural</strong> type.K-feldspar <strong>and</strong> plagioclase commonly show weakLPO in all rock types regardless the origin <strong>of</strong> grains.K-feldspar shows crystallographic patterns which arecompatible with dominant activity <strong>of</strong> the 1/2 [110](001)slip system (Willaime & G<strong>and</strong>ais, 1977; Willaime et al.,1979) (Fig. 12a, c). Contribution <strong>of</strong> other slip systemsas [100](010) (Fig. 12b) <strong>and</strong> [100](001) (Fig. 12d) hasalso been identified in both relict K-feldspar grains <strong>and</strong>in K-feldspar grains apparently crystallized from meltrespectively.Distribution <strong>of</strong> the main lattice directions <strong>of</strong> plagioclaserevealed slip parallel either to 1/2[1 10] on (001)<strong>and</strong> (11 1) planes (Fig. 12e) or to 1/2 [110] on (001) <strong>and</strong>(1 1 1) planes (Fig. 12f) for all types <strong>of</strong> rocks <strong>and</strong> bothaggregate grains <strong>and</strong> plagioclase inferred to havecrystallized from melt (Fig. 12g) (Olsen & Kohlstedt,1984). The textures <strong>of</strong> plagioclase inferred to haveÓ 2007 Blackwell Publishing Ltd329


44 P. HASALOVÁ ET AL.(a)(b)Fig. 11. Characteristic c-axes preferred orientations <strong>of</strong> (a) old/relict quartz grains <strong>and</strong> (b) new quartz grains crystallized from areas <strong>of</strong>inferred former melt for all rock types. The c-axis patterns <strong>of</strong> old/relict quartz grains in type I b<strong>and</strong>ed orthogneiss indicate prism Æcæslip system activity whereas in type II, III <strong>and</strong> IV migmatites basal Æaæ or rhomb Æa + cæ slip systems are dominant with minorprism Æaæ slip. New quartz grains inferred to have crystallized from melt in type I b<strong>and</strong>ed orthogneiss to type IV nebulitic migmatiteshow very weak LPO <strong>and</strong> nearly isotropic distribution <strong>of</strong> all quartz axes. Equal area projections, lower hemisphere, contoured atinterval <strong>of</strong> 0.5 times uniform distribution. Foliation is horizontal <strong>and</strong> lineation is in this plane in the E–W direction. N is the number <strong>of</strong>measured grains. Maximum densities are marked on the bottom right <strong>of</strong> each pole figure. The dashed line represents the lowest contourlevel <strong>and</strong> the grey circle corresponds to the minimum density value.crystallized from melt are commonly weak with theexception <strong>of</strong> strong LPO <strong>of</strong> plagioclase in the type Iorthogneiss (Fig. 12f). Such slip-systems are supposedto be secondary <strong>and</strong> active if grains are in unsuitable(hard) orientation to the dominant slip-system[100](010) (Kruse et al., 2001).DISCUSSIONThis study presents a detailed micro<strong>structural</strong> <strong>and</strong>quantitative textural analysis <strong>of</strong> four types <strong>of</strong> migmatiticrocks identified in one <strong>of</strong> the largest (5000 km 2 )migmatitic complex <strong>of</strong> the eastern Variscan belt. Therock types are interpreted as representing a texturalsequence from b<strong>and</strong>ed orthogneiss via stromatic <strong>and</strong>schlieren migmatites to nebulitic migmatite. The possiblemechanisms that could account for the origin <strong>of</strong>this rock sequence involve: (i) genetically unrelatedmigmatites that have originated from distinct protoliths;(ii) variable degree <strong>of</strong> in situ partial melting <strong>of</strong> asingle protolith or different protoliths; <strong>and</strong> (iii) meltinfiltration from an external source through solid rockin which b<strong>and</strong>ed orthogneiss <strong>and</strong> nebulitic migmatiterepresent genetically linked end-members. Thesehypotheses are discussed further below.Spatial relationships <strong>of</strong> individual migmatite types withinthe shear zoneThe <strong>structural</strong> sequence described in this work indicatesan intimate relationship between types I to IIImigmatites <strong>and</strong> nebulitic type IV migmatite sheets thatcan be interpreted in terms <strong>of</strong> a shear zone, which wasexploited by rising magma (Brown et al., 1995; Collins& Sawyer, 1996; Brown & Solar, 1998b). We haveshown that the D 2 flat fabrics that cross-cut the steepfoliation S 1 developed at high-temperature solid-stateconditions (Fig. 1). Tajcˇmanova´ et al. (2006) <strong>and</strong>Racek et al. (2006) described a similar sequence <strong>of</strong>superposed fabrics in lower crustal rocks several tens<strong>of</strong> kilometres to the north <strong>and</strong> south <strong>of</strong> the studied arearespectively. These authors proposed that the flat D 2deformation fabrics originated due to thrusting <strong>of</strong>orogenic lower crust over middle crustal units along alarge-scale retrograde shear zone. In agreement withthese authors, we suggest that the D 2 fabrics developedin a thrust related crustal-scale shear zone reportedalready by Urban (1992), Schulmann et al. (1994) <strong>and</strong>redefined later by Schulmann et al. (2005). The maindifference between other regions is in the degree <strong>of</strong> D 2reworking, which is so high in the studied area thatÓ 2007 Blackwell Publishing Ltd330


ORIGIN OF FELSIC MIGMATITES 45Fig. 12. Characteristic LPO patterns <strong>of</strong> K-feldspar (a–d) <strong>and</strong> plagioclase (e,f). Both feldspars commonly show weak LPO in all rocktypes regardless <strong>of</strong> the origin <strong>of</strong> the grains. An exception is the strong LPO <strong>of</strong> new plagioclase grains in the type I b<strong>and</strong>ed orthogneiss(f). K-feldspar usually shows activity <strong>of</strong> 1/2[110](001) (a,c; type I), but also <strong>of</strong> [100](010) (b; type IV) <strong>and</strong> [100](001) (d; type IV) slipsystems. Plagioclase reveals activity <strong>of</strong> secondary slip systems such as 1/2[1 10] on (001) <strong>and</strong> (11 1) or 1/2 [1 10] on (001) <strong>and</strong> (1 1 1)(e,f; type I). Equal area projections, lower hemisphere, contoured at intervals <strong>of</strong> 0.5 times uniform distribution. Foliation is horizontal<strong>and</strong> lineation is in this plane in the E–W direction. N is the number <strong>of</strong> measured grains. Maximum densities are marked on thebottom right <strong>of</strong> each pole figure. The dashed line represents the lowest contour level <strong>and</strong> the grey circle corresponds to the minimumdensity value. (f,g) BSE images depicting the micro<strong>structural</strong> appearance <strong>of</strong> examples <strong>of</strong> the measured phases (sample PH60/B).Original plagioclase (g) <strong>and</strong> K-feldspar (h) aggregates with newly crystallized quartz (white arrow), plagioclase (black arrows) <strong>and</strong>K-feldspar (grey arrows) are shown.Ó 2007 Blackwell Publishing Ltd331


46 P. HASALOVÁ ET AL.steep D 1 fabrics are preserved only as rare relics shownin Fig. 2.Our <strong>structural</strong> observations are compatible withprogressive transposition within the ductile shear zoneranging from type I b<strong>and</strong>ed orthogneiss to highlyreworked type III schlieren migmatite. This interpretationis based on progressive folding <strong>of</strong> early steeporthogneiss fabric, development <strong>of</strong> isoclinal folds <strong>and</strong>complete fabric transposition <strong>and</strong> development <strong>of</strong> typeIII migmatite in a ductile shear zone (Fig. 2; Turner &Weiss, 1963). In such a context, the elongated bodies <strong>of</strong>type IV nebulitic migmatite can be seen as veins <strong>of</strong>isotropic granite penetrating parallel to the main S 2mylonitic anisotropy (e.g. Cosgrove, 1997; Brown &Solar, 1998a). Alternatively, the type IV nebuliticmigmatites could be interpreted in terms <strong>of</strong> injectedmelt into hot country rocks (called also magmawedging, Weinberg & Searle, 1998) preventing magmafreezing during D 2 shearing. Finally, the nebuliticmigmatite can be also regarded as the most extremeend-member <strong>of</strong> the <strong>structural</strong> sequence, i.e. completelydisintegrated parental orthogneiss.In summary, the type I orthogneiss to type III migmatiteshow intimate spatial relationships suggestingthat they have originated from the same protolith <strong>and</strong>that they are genetically linked. However, the macroscopicobservations alone cannot distinguish the origin<strong>of</strong> type IV migmatite <strong>and</strong> further arguments arerequired.Micro<strong>structural</strong> <strong>and</strong> petrological arguments for melt–rockinteraction during exhumationWe suggest, in agreement with Sawyer (1999, 2001),that the position <strong>and</strong> topology <strong>of</strong> new plagioclase,quartz <strong>and</strong> K-feldspar grains in type I orthogneiss <strong>and</strong>type II migmatite may be interpreted in terms <strong>of</strong> meltproducts crystallized along boundaries <strong>of</strong> the feldsparin individual aggregates (Fig. 5a–c). The main differencebetween type I orthogneiss <strong>and</strong> type II migmatiteis a more albitic composition <strong>of</strong> plagioclase <strong>and</strong> agreater modal content <strong>of</strong> new phases in the latter.Types III <strong>and</strong> IV migmatites show development <strong>of</strong>highly corroded shapes <strong>of</strong> K-feldspar, plagioclase <strong>and</strong>biotite (Fig. 4c–f). This indicates that all rock typesexhibit features compatible with the presence <strong>of</strong> melt<strong>and</strong> its interaction with the solid rock. Additionally,the degree <strong>of</strong> melt–rock interaction is inferred toincrease from type I orthogneiss towards type IVnebulitic migmatite (Fig. 4).The <strong>structural</strong> sequence exhibits a distinct trend inmodal composition <strong>of</strong> originally monomineralic layersthat are progressively converted into polymineralicaggregates <strong>of</strong> granitic composition (Fig. 6). The compositionalpaths show evolutionary trends from type Iorthogneiss to type III migmatite, interpreted as beingassociated with crystallization <strong>of</strong> melt, culminating inthe type IV migmatite, which has equal amounts <strong>of</strong>plagioclase, K-feldspar <strong>and</strong> quartz (Fig. 6).Plagioclase shows systematic decrease in anorthitecontent for both original plagioclase grains (An 30 toAn 25 ), their rims <strong>and</strong> inter-granular aggregates (An 20to An 10 ) towards type IV nebulitic migmatite. Bothgarnet <strong>and</strong> biotite exhibit systematically increasing X Fetowards type IV migmatite (from 0.7 to 1.00 <strong>and</strong> from0.4 to 0.9 respectively) coupled with decrease in Ticontent in biotite (from 0.2 to 0.04 p.f.u.). The mineralcompositional data suggest systematic equilibration <strong>of</strong>garnet, biotite <strong>and</strong> plagioclase compositions in thestability field <strong>of</strong> sillimanite with decreasing temperature.The full petrological data <strong>and</strong> P–T estimates formelt–rock interaction are presented in a companionpaper (Hasalova´ et al., 2008a). Here, we quote the P–Testimates based on thermodynamic <strong>modelling</strong> usingTHERMOCALC (Powell et al., 1998) to point out thedecrease in temperature from 790 to 850 °C at 7.5 kbarfor type I orthogneiss to 690–770 °C at 4.5 kbar fortype IV nebulitic migmatite.Taken together, the micro<strong>structural</strong> <strong>and</strong> petrologicaldata show paradoxically an increasing degree <strong>of</strong>apparent melt–rock interaction coupled with decreasingequilibration temperature with textural evolutionfrom type I to type IV rock type. The micro<strong>structural</strong><strong>and</strong> modal composition data do not exclude eitherpartial melting or infiltration <strong>of</strong> melt from an externalsource. However, the systematic modification <strong>of</strong>chemical composition <strong>of</strong> minerals across the migmatitesequence is in contradiction with the model <strong>of</strong> in situpartial melting. Namely, as suggested by many field<strong>and</strong> experimental studies, the composition <strong>of</strong> plagioclasewould be shifted towards more anorthitic contents<strong>and</strong> the X Fe <strong>of</strong> garnet <strong>and</strong> biotite would decreaseduring partial melting process (Le Breton & Thompson,1988; Vielzeuf & Holloway, 1988; Gardien et al.,1995; Greenfield et al., 1998; Dallain et al., 1999).Interpretation <strong>of</strong> quantitative micro<strong>structural</strong> dataMicro<strong>structural</strong> studies <strong>of</strong> partially molten rock haverevealed systematic changes in grain size, grain SPO<strong>and</strong> spatial distribution <strong>of</strong> individual phases duringincreasing degree <strong>of</strong> partial melting (e.g. Vernon, 1976;McLellan, 1983; Dallain et al., 1999). Here thesetrends are compared with our quantitative micro<strong>structural</strong>data, <strong>and</strong> the alternative origins that mayresult in the observed micro<strong>structural</strong> sequence arediscussed.Interpretation <strong>of</strong> crystal size distributionsThe most significant result <strong>of</strong> this study is the systematicdecrease in average grain size (Fig. 7a) <strong>and</strong>systematic increase in a population density (nucleationrate) associated with possible decrease in growth ratefor all feldspar from type I orthogneiss to type IVmigmatite (Fig. 7b, c).Results from migmatitic terranes show that the CSDassociated with partial melting is characterized byÓ 2007 Blackwell Publishing Ltd332


ORIGIN OF FELSIC MIGMATITES 47production <strong>of</strong> coarse-grained felsic mineral aggregatesresulting from increase in temperature (e.g. Dougan,1983; McLellan, 1983). This process is commonly followedby textural coarsening (Ashworth & McLellan,1985; Dallain et al., 1999; Berger & Roselle, 2001)explained by two competing approaches: the Lifshitz–Slyozov–Wagner (LSW) model (Lifshitz & Slyozov,1961), <strong>and</strong> the communicating neighbour theory (CN<strong>of</strong> DeH<strong>of</strong>f, 1991). Higgins (1998) showed that texturalcoarsening results in progressive decrease in N 0 value<strong>and</strong> decrease in the slope <strong>of</strong> the CSD curve; he interpretedthis trend as a result <strong>of</strong> rapid undercoolingduring solidification <strong>of</strong> magma followed by reducedundercooling, suppression <strong>of</strong> nucleation <strong>and</strong> texturalcoarsening. However, our textural sequence exhibitsthe opposite trend in evolution <strong>of</strong> CSD curves, which isinterpreted as indicating that in situ partial melting <strong>and</strong>textural coarsening are not responsible for the origin <strong>of</strong>observed CSDs.The observed CSD trend may be explained by one <strong>of</strong>three different mechanisms: (1) solid-state deformationunder decreasing temperature <strong>and</strong> or increasing strainrate (Hickey & Bell, 1996; Azpiroz & Ferna´ndez, 2003;Lexa et al., 2005); (2) a different degree <strong>of</strong> reactionoverstepping (Waters & Lovegrove, 2002; Moazzen &Modjarrad, 2005); <strong>and</strong> (3) a different degree <strong>of</strong> undercooling(Marsh, 1988).The grain size for dynamically recrystallized grainsin a power-law creep regime is a function <strong>of</strong> differentialstress (Twiss, 1977). Such grains are characterized bystrong shape <strong>and</strong> LPO <strong>and</strong> commonly solid-state differentiation(Baratoux et al., 2005; Lexa et al., 2005).However, this micro<strong>structural</strong> study does not revealany features in quartz, plagioclase <strong>and</strong> K-feldspar <strong>of</strong>rock types II, III <strong>and</strong> IV which may indicate a dynamicrecrystallization processes operating under decreasingtemperature. Differences in the degree <strong>of</strong> reactionoverstepping have been documented in contact aureoles,but may be rejected in this case due to theregional nature <strong>of</strong> the metamorphism. However, therole <strong>of</strong> different degrees <strong>of</strong> undercooling relating to anoverall decrease in equilibration temperature cannot beexcluded.Our data indicate that the sequence <strong>of</strong> rock typesreflects the progressive resorption <strong>of</strong> residual grains<strong>and</strong> crystallization <strong>of</strong> new grains from melt in intergranularspaces. Moreover, the trend <strong>of</strong> CSD curvessuggest a progressive increase in nucleation rate <strong>and</strong>decrease in growth rate from type I orthogneiss totype IV nebulitic migmatite. This trend could beexplained by an increase in undercooling consistentwith the decreasing equilibration temperature wereport.The CSD trend is compatible with crystallization <strong>of</strong>melt in a progressively exhuming <strong>and</strong> rapidly coolingsystem. This is in accordance with exceptionally highcooling rates up to several hundred degrees celsius permillion years estimated for nearby granulites byTajcˇmanova´ et al. (2006).Interpretation <strong>of</strong> spatial distributions <strong>of</strong> phasesThe quantitative analysis <strong>of</strong> spatial distributions <strong>of</strong>individual phases shows that the intensification <strong>of</strong>regular distribution (increasing amount <strong>of</strong> unlikecontacts; Fig. 9) correlates with an increasing degree <strong>of</strong>host rock–melt equilibration. The process <strong>of</strong> meltcrystallization leads to new mineral growth on thesurfaces <strong>of</strong> residual grains. This is responsible for theincrease in unlike grain boundaries, which commonlyretain melt–solid geometries. Our case study showsthat the development <strong>of</strong> a regular distribution <strong>of</strong> felsicphases is not related to solid-state annealing, as supposedby some authors (Flinn, 1969; McLellan, 1983;Lexa et al., 2005), but to the process <strong>of</strong> crystallization<strong>of</strong> melt, consistent with precipitation <strong>of</strong> the minorphase on triple points in granular polygonal aggregatesto achieve lower total interfacial energy (Spry, 1969;Vernon, 1974). This process was documented by Dallainet al. (1999), who showed that the predominance<strong>of</strong> unlike contacts in polycrystalline aggregates originatedthrough wetting <strong>of</strong> grain boundaries by fluids ormelt, <strong>and</strong> subsequent precipitation <strong>of</strong> other phases onlike–like contacts. However, we cannot exclude thepossibility that a regular distribution reported fromgranulites <strong>and</strong> high-grade gneisses (Flinn, 1969; Kretz,1994) results from solid-state annealing <strong>of</strong> rocks wheremelt crystallized. Therefore, the regular distributiondeveloped during melt crystallization may be inherited<strong>and</strong> perhaps further accentuated during later thermal<strong>and</strong> textural re-equilibration.Origin <strong>of</strong> micro<strong>structural</strong> <strong>and</strong> compositional trendsThe sequence from type I orthogneiss to type IVmigmatite exhibit continuous trends in all quantitativeparameters (Table 1). The grain size decreases(Fig. 7a) <strong>and</strong> there is a progressive development <strong>of</strong> aregular distribution <strong>of</strong> all felsic phases (Fig. 9), whichis linked with mineral compositional trends indicatingtemperature decrease. These clear evolutionary trendsare incompatible with a process <strong>of</strong> partial melting <strong>of</strong>different protoliths. Partial melting <strong>of</strong> the same protolithmay develop continuous trends, but these shouldshow increase in grain size <strong>of</strong> individual felsic phases(Dallain et al., 1999) <strong>and</strong> different mineral compositionalevolution (e.g. Gardien et al., 1995; Greenfieldet al., 1998). Additionally, we show that the degree <strong>of</strong>regular distribution for K-feldspar- <strong>and</strong> plagioclasedominatedaggregates evolves in the same mannerthroughout the micro<strong>structural</strong> sequence (Fig. 9).However, Dallain et al. (1999) reported significantlymore advanced regular distribution <strong>of</strong> plagioclasecomparedwith K-feldspar-rich aggregates in themicro<strong>structural</strong> sequence originated by partial melting.These authors proposed that this micro<strong>structural</strong>contrast originated due to melting process preferentiallyoperating in mica–plagioclase rich aggregates,whereas the K-feldspar-rich aggregates were moreÓ 2007 Blackwell Publishing Ltd333


48 P. HASALOVÁ ET AL.refractory. In the present case, Hasalova´ et al. (2008b)report continuous trends in whole-rock geochemistry<strong>and</strong> mineral compositions for the sequence <strong>of</strong> rocktypes, but different Nd isotopic composition for thetype I orthogneiss compared with the rest <strong>of</strong> the sequence,which precludes <strong>of</strong> in situ anatexis in a closedsystem.Melt infiltration modelThe discrepancies between the evolutionary trends wereport <strong>and</strong> generally accepted trends for anatecticterranes require an appropriate explanation that isconsistent with the <strong>structural</strong>, quantitative micro<strong>structural</strong><strong>and</strong> mineral compositional data. As a possibleexplanation, we introduce the concept <strong>of</strong> meltinfiltration from an external source, where melt passespervasively along grain boundaries through the wholerockvolume <strong>and</strong> changes macroscopic (Fig. 2) <strong>and</strong>microscopic (Fig. 3) appearance <strong>of</strong> the rock. Thisprocess is characterized by resorption <strong>of</strong> old phases,nucleation <strong>of</strong> new phases along high-energy like–likegrain boundaries <strong>and</strong> modification <strong>of</strong> mineral <strong>and</strong>whole-rock compositions. These gradual changes areaccompanied by grain size reduction (Fig. 7) <strong>and</strong>progressive disintegration <strong>of</strong> former aggregate (layered)distribution <strong>of</strong> original phases (Fig. 9). We suggestthat the individual migmatite types representdifferent degrees <strong>of</strong> equilibration between the host rock<strong>and</strong> migrating melt. It should be emphasized, that allthese processes occur along a retrograde path duringexhumation <strong>of</strong> the Gfo¨ hl Unit. We are aware that adecrease in P–T conditions during melt infiltration is afundamental <strong>and</strong> limiting factor for the model proposed.The amount <strong>of</strong> melt <strong>and</strong> its connectivity are criticalparameters controlling melt mobility <strong>and</strong> the rheologicalbehaviour <strong>of</strong> melt-present rocks. To constrainthese parameters both AMS <strong>and</strong> EBSD were used.Using AMS, it is possible to distinguish between solidstatedominated deformation mechanisms in themelanosome <strong>and</strong> free rigid body particle rotation in theleucosome (e.g. Ferré et al., 2003). On the other h<strong>and</strong>,using the EBSD technique enables us to distinguishdeformation mechanisms in the solid framework <strong>and</strong>to constrain the mechanical role <strong>of</strong> melt during thedeformation.AMS fabric origin: solid framework or melt controlleddeformationThe AMS study shows that the magnetic anisotropy isdominated by biotite. The oblate shape <strong>of</strong> magneticellipsoid <strong>and</strong> high degree <strong>of</strong> anisotropy <strong>of</strong> type I orthogneiss<strong>and</strong> type II migmatite (Fig. 10a) are consistentwith strong preferred orientation <strong>of</strong> biotite <strong>and</strong> thefact that biotite has a intrinsically oblate shape <strong>of</strong> thesingle-grain magnetic ellipsoid (Zapletal, 1990; Martı´n-Hern<strong>and</strong>éz & Hirth, 2003). The type III <strong>and</strong> IV migmatitesreveal partly resorbed biotite flakes uniformlydispersed in the rock marked by slightly weaker degree<strong>of</strong> magnetic anisotropy <strong>and</strong> less oblate fabric ellipsoidcompared with types I <strong>and</strong> II migmatite (Fig. 10a).This contrasts with common granites <strong>and</strong> diatexitesfrom other migmatitic terranes which show significantlylower values <strong>of</strong> degree <strong>of</strong> anisotropy <strong>and</strong> highlyvariable shapes <strong>of</strong> AMS ellipsoids (Fig. 10a; Bouchez,1997; Ferre´ et al., 2003).Numerous natural studies supported by <strong>numerical</strong><strong>modelling</strong> indicate that the magnetic susceptibility inviscously flowing magmas is characterized by a verylow degree <strong>of</strong> anisotropy, pulsatory fabrics <strong>and</strong> dominantlya plane strain AMS ellipsoid shape (Blumenfeld& Bouchez, 1988; Hrouda et al., 1994; Arbaretet al., 2000). A comparison <strong>of</strong> the AMS fabrics withthose <strong>of</strong> diatexites <strong>and</strong> results <strong>of</strong> <strong>numerical</strong> modelsindicate that the intensity <strong>of</strong> the AMS fabric <strong>of</strong> typesIII <strong>and</strong> IV migmatites does not originated from freelyrotated biotite in viscously flowing melt. On the contrary,we argue that the AMS fabric in all types <strong>of</strong>migmatites resembles fabrics usually acquired throughsolid-state deformation <strong>of</strong> a load-bearing framework,similar to melanosomes in migmatites (Ferre´ et al.,2003). To underst<strong>and</strong> the mechanisms responsible fordevelopment <strong>of</strong> such fabrics the grain-scale deformationmechanisms <strong>and</strong> melt behaviour in individual rocktypes is discussed.Deformation mechanismsExperimental studies <strong>of</strong> low melt fraction rocks deformedunder high differential stress show that matrixminerals deform by grain boundary migrationaccommodated dislocation creep (DellÕ Angelo et al.,1987; Walte et al., 2005). Strong shape <strong>and</strong> GBPO <strong>of</strong>feldspar (Figs 8 & 9) as well as LPO <strong>of</strong> residual quartzgrains (Fig. 11a) in the type I orthogneiss may beinterpreted in terms <strong>of</strong> plastic deformation consistentwith a dislocation creep deformation mechanism(Rosenberg & Berger, 2001). However, the weak LPO<strong>of</strong> residual grains <strong>of</strong> both feldspars (Fig. 12a, e) in thetype I orthogneiss suggests a contribution <strong>of</strong> grainboundary sliding during the development <strong>of</strong> themicrostructure. In other words, the type I microstructurecorresponds to a transient microstructure interms <strong>of</strong> decreasing activity <strong>of</strong> dislocation creep <strong>and</strong>enhancement <strong>of</strong> diffusion controlled processes.Decrease in SPO <strong>and</strong> GBPO <strong>and</strong> constantly weakLPO in feldspar <strong>of</strong> type II migmatite (Figs 8 & 9) maybe interpreted as a result <strong>of</strong> melt-enhanced diffusioncreep (Garlick & Gromet, 2004). However, the largequartz grains reveal intense activity <strong>of</strong> basal Æaæ slipsuggesting important plastic yielding <strong>of</strong> this mineral(Fig. 11a). Elongate pockets inferred to represent formermelt oriented at a high angle to the stretchinglineation in the type I orthogneiss (Fig. 5a) <strong>and</strong> type IImigmatite indicate that the melt distribution wascontrolled by the deformation. This is supported by theÓ 2007 Blackwell Publishing Ltd334


ORIGIN OF FELSIC MIGMATITES 49strong LPO <strong>of</strong> interstitial plagioclase (Fig. 12f).Rosenberg & Riller (2000) reported that pockets withinquartz aggregates inferred to have been former meltare oriented at high angle to the foliation plane, possiblyclose to r 1 . Melt distribution in our samples issimilar to their results <strong>and</strong> also to experiments at highdifferential stresses (DellÕ Angelo & Tullis, 1988) <strong>and</strong>high confining pressures. In these experiments, meltaccumulated in pockets along faces <strong>of</strong> the grains subparallelto the main compressional stress direction r 1(DellÕ Angelo & Tullis, 1988; Daines & Kohlstedt,1997). Such a melt topology is also termed ÔdynamicwettingÕ (Jin et al., 1994).In types III <strong>and</strong> IV migmatites both residual <strong>and</strong>new grains <strong>of</strong> K-feldspar <strong>and</strong> plagioclase exhibit lowSPO, GBPO <strong>and</strong> LPO (Figs 8, 9 & 12), indicatingabsence <strong>of</strong> dislocation creep, in contrast to quartz,which exhibits relatively strong crystallographic preferredorientation (Fig. 11a). The topology <strong>of</strong> formermelt is poorly constrained in both types III <strong>and</strong> IVmigmatite but, the SPO <strong>of</strong> the minor phases interpretedto have crystallized from melt shows a bimodaldistribution sub-perpendicular <strong>and</strong> sub-parallel to theS 2 foliation. These observations are neither compatiblewith high differential stress nor low differential stressexperiments, in which the melt occurs primarily intriple point junctions without any SPO (DellÕ Angeloet al., 1987; Gleason et al., 1999). However, in somenatural samples, former melt pockets are preferentiallylocated along grain boundaries parallel to the foliation(John & Stu¨ nitz, 1997; Sawyer, 1999; Rosenberg &Berger, 2001), indicating that the orientation <strong>of</strong> meltpocket in nature is not always in agreement withexperimental studies (Rosenberg, 2001). In this study,the melt pocket orientation sub-parallel to the foliationmay indicate low differential stress <strong>and</strong> high fluid/meltpressure as suggested by Cosgrove (1997).We conclude that during evolution from type Ib<strong>and</strong>ed orthogneiss to type IV nebulitic migmatite meltwetted a majority <strong>of</strong> grain contacts. The AMS study<strong>and</strong> quartz micr<strong>of</strong>abrics in types II to IV migmatitessuggest that the melt fraction did not exceed the criticalamount to allow free relative movement <strong>of</strong> grainswithout interference, i.e. the melt fraction is below thecritical threshold (e.g. RCMP <strong>of</strong> Arzi, 1978; RPT <strong>of</strong>Vigneresse et al., 1996). Rosenberg & H<strong>and</strong>y (2005)argued that melt fractions <strong>of</strong> only / ¼ 0.07 (meltconnectivity threshold, MCT) will enable the formation<strong>of</strong> interconnected networks <strong>of</strong> melt under dynamicconditions which will lead to a substantial strengthdrop. These authors suggested that weakening at theMCT probably involves localized, inter- <strong>and</strong> intragranularmicrocracking, as well as limited rigid bodyrotation <strong>of</strong> grains, without an important contribution<strong>of</strong> dislocation creep <strong>and</strong> diffusion processes at grainboundaries. However, we do not observe any strainlocalization associated with brittle failure <strong>and</strong> thereforeit is suggested that the deformation has to beaccommodated by mechanisms operating homogeneouslyacross significant rocks volumes. Materialscience experiments (Mabuchi et al., 1997) show thatweakening due to melt-enhanced grain boundary slidingat low melt fraction is an efficient mechanismallowing homogeneous deformation. We suggest thatdeformation <strong>of</strong> both feldspars <strong>and</strong> quartz in the type IIto type IV migmatites occurred by melt-enhanced grainboundary sliding with a contribution to the overalldeformation by dislocation creep. These characteristicsare compatible with granular flow as described byPaterson (2001) accompanied by melt-enhanced diffusion<strong>and</strong>/or direct melt flow.CONCLUSIONSBased on a detail field <strong>and</strong> micro<strong>structural</strong> study, wedistinguish four types <strong>of</strong> gneiss/migmatite in the Gfo¨ hlgneiss complex: (i) b<strong>and</strong>ed orthogneiss (type I), withdistinct layers <strong>of</strong> recrystallized plagioclase, K-feldspar<strong>and</strong> quartz separated by layers <strong>of</strong> biotite; (ii) stromaticmigmatite (type II), composed <strong>of</strong> plagioclase <strong>and</strong> K-feldspar aggregates with subordinate quartz <strong>and</strong>irregular quartz aggregates – the boundaries betweenindividual aggregates are ill defined <strong>and</strong> rather diffuse;(iii) schlieren migmatite (type III), which consists <strong>of</strong>plagioclase–quartz- <strong>and</strong> K-feldspar–quartz-enricheddomains with a foliation marked only by preferredorientation <strong>of</strong> biotite <strong>and</strong> sillimanite dispersed in therock; <strong>and</strong>, (iv) nebulitic migmatite (type IV), with norelicts <strong>of</strong> gneissosity. It is demonstrated that this is acontinuous sequence developed by melt-presentdeformation, in which the type I b<strong>and</strong>ed orthogneisses<strong>and</strong> type IV nebulitic migmatites are end-members.The progressive disintegration <strong>of</strong> the b<strong>and</strong>edmicrostructure <strong>and</strong> the development <strong>of</strong> nebuliticmigmatite is characterized by several systematic texturalchanges. The grain size <strong>of</strong> all felsic phasescontinuously decrease whereas the population density<strong>of</strong> precipitated phases increases. The new phasespreferentially nucleate along high-energy like–likeboundaries, causing the development <strong>of</strong> a regular distribution<strong>of</strong> individual phases. Simultaneously, themodal proportions <strong>of</strong> felsic phases evolve towards aÔgranite minimumÕ composition. Further, this evolutionarytrend is accompanied by a decrease in grainSPO <strong>of</strong> all felsic phases. To explain these textural <strong>and</strong>compositional changes we introduce a model <strong>of</strong> meltinfiltration from an external source in which melt isargued to pass pervasively along grain boundariesthrough the whole-rock volume. It is suggested that theindividual migmatite types represent different degrees<strong>of</strong> equilibration between the host rock <strong>and</strong> migratingmelt during the retrograde metamorphic evolution.The inferred melt topology in type I orthogneissexhibits elongated pockets <strong>of</strong> melt oriented at a highangle to the compositional b<strong>and</strong>ing, indicating that themelt distribution was controlled by deformation thesolid framework. Here, the microstructure exhibitsfeatures compatible with a combination <strong>of</strong> dislocationÓ 2007 Blackwell Publishing Ltd335


50 P. HASALOVÁ ET AL.creep <strong>and</strong> grain boundary sliding deformation mechanisms.The types II–IV microstructures developed bygranular flow accompanied by melt-enhanced diffusion<strong>and</strong>/or melt flow. However, the amount <strong>of</strong> melt presentnever exceeded a critical threshold during the deformationto allow free rotation <strong>of</strong> biotite grains.The model <strong>of</strong> melt infiltration based on <strong>structural</strong><strong>and</strong> micro<strong>structural</strong> observation is supported by thermodynamic(Hasalova´ et al., 2008a) <strong>and</strong> geochemical<strong>modelling</strong> (Hasalova´ et al., 2008b). Although our dataseem to be consistent with such a model, there are stilla number <strong>of</strong> issues to be resolved (e.g. time-scale <strong>of</strong> theprocess, the character <strong>of</strong> the melt <strong>and</strong> the grain-scaledeformation mechanisms enabling pervasive flow <strong>of</strong>viscous melt). Nevertheless, our model has pr<strong>of</strong>oundconsequences for the petrogenesis <strong>of</strong> migmatites, therheology <strong>of</strong> anatectic regions during syn-orogenicexhumation <strong>and</strong> melt transport in the crust.ACKNOWLEDGEMENTSThis work was financially supported by the GrantAgency <strong>of</strong> the Czech Academy <strong>of</strong> Science (Grants No.IAA311140 <strong>and</strong> No. 205/04/2065) by an internalCNRS funding (CGS/EOST) <strong>and</strong> by the FrenchNational Agency (No. 06-1148784). Visits by P.Hasalova´ to ULP Strasbourg were funded by theFrench Government Foundation (BGF). We gratefullyacknowledge A. Langrova´ from the Institute <strong>of</strong>Geology at the Czech Academy <strong>of</strong> Sciences for operatingthe microprobe. We also thank the reviewersC. Rosenberg, R. Weinberg <strong>and</strong> H. Stu¨ nitz for theirconstructive comments <strong>and</strong> suggestions for improvingthis paper, <strong>and</strong> the Journal editor M. Brown for hiscareful h<strong>and</strong>ling <strong>of</strong> the manuscript. R. Powell <strong>and</strong>M. Brown are thanked for helpful discussions.REFERENCESAdams, B. L., Wright, S. I. & Kunze, K., 1993. Orientationimaging: the emergence <strong>of</strong> a new microscopy. MetallurgicalTransactions A, 24, 819–831.Arbaret, L., Fern<strong>and</strong>ez, A., Jezˇek, J., Ildefonse, B., Launeau, P.& Diot, H., 2000. Analogue <strong>and</strong> <strong>numerical</strong> modeling <strong>of</strong> shapefabrics: application to strain <strong>and</strong> flow determination in magma.Transactions <strong>of</strong> the Royal Society <strong>of</strong> Edinburgh: EarthSciences, 90, 97–109.Arzi, A. A., 1978. Critical phenomena in the rheology <strong>of</strong> partiallymelted rocks. Tectonophysics, 44, 173–184.Ashworth, J. R., 1976. Petrogenesis <strong>of</strong> migmatites in the Huntly-Portsoy area, north-east Scotl<strong>and</strong>. Mineralogical Magazine,40, 661–682.Ashworth, J. R., 1979. Comparative petrography <strong>of</strong> deformed<strong>and</strong> undeformed migmatites from the Grampian Highl<strong>and</strong>s <strong>of</strong>Scotl<strong>and</strong>. Geological Magazine, 116, 445–456.Ashworth, J. R. & McLellan, E. L., 1985. Textures. In: Migmatites(ed. Ashworth, J.R.), pp. 180–204, Blackie, Glasgow.Azpiroz, M. D. & Fernández, C., 2003. Characterization <strong>of</strong>tectono-metamorphic events using crystal size distribution(CSD) diagrams. A case study from the Acebuchesmetabasites (SW Spain). Journal <strong>of</strong> Structural Geology, 25,935–947.Bailey, E. H. & Stevens, R. H., 1960. Selecting straining <strong>of</strong> K-feldspar <strong>and</strong> plagioclase on rock lobes <strong>and</strong> thin sections.American Mineralogist, 45, 1020–1025.Baratoux, L., Schulmann, K., Ulrich, S. & Lexa, O., 2005.Contrasting microstructures <strong>and</strong> deformation mechanisms inmetagabbro mylonites contemporaneously deformed underdifferent temperatures (c. 650 °C <strong>and</strong> c. 750 °C). In: DeformationMechanisms, Rheology <strong>and</strong> Tectonics: From Minerals toLithosphere. (eds Gapais, D., Brun, J.P. & Cobbold, P.R.),Geological Society Special Publications, pp. 97–125, GeologicalSociety <strong>of</strong> London, London, UK.Barraud, J., Gardien, V., Allem<strong>and</strong>, P. & Gr<strong>and</strong>jean, P., 2001.Analogue <strong>modelling</strong> <strong>of</strong> melt segregation <strong>and</strong> migration duringdeformation. Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth (A), 26, 317–323.Bascou, J., Barruol, G., Vauchez, A., Mainprice, D. & Egydio-Silva, M., 2001. EBSD-measured lattice-preferred orientations<strong>and</strong> seismic properties <strong>of</strong> eclogites. Tectonophysics, 342, 61–80.Berger, A. & Roselle, G., 2001. Crystallization processes inmigmatites. American Mineralogist, 86, 215–224.Blumenfeld, P. & Bouchez, J.-L., 1988. Shear criteria <strong>and</strong> migmatitedeformed in the magmatic <strong>and</strong> solid states. Journal <strong>of</strong>Structural Geology, 10, 361–372.Bouchez, J.-L., 1997. Granite is never isotropic: an introductionto AMS studies <strong>of</strong> granitic rocks. In: Granite: From Segregation<strong>of</strong> Melt to Emplacement Fabrics (eds Bouchez, J.-L.,Hutton, D.H.W. & Stephens, W.E.), pp. 95–112. Kluwer,Dortrecht.van Breemen, O., Aftalion, M., Bowes, D. R. et al., 1982.Geochronological studies <strong>of</strong> the Bohemian Massif, Czechoslovakia,<strong>and</strong> their significance in the evolution <strong>of</strong> CentralEurope. Transactions <strong>of</strong> the Royal Society <strong>of</strong> Edinburgh, EarthSciences, 73, 89–108.Brown, M., 1994. The generation, segregation, ascent <strong>and</strong>emplacement <strong>of</strong> granite magma: the migmatite-to-crustallyderivedgranite connection in thickened orogens. Earth-Science Reviews, 36, 83–130.Brown, M. & Rushmer, T., 2006. Evolution <strong>and</strong> Differentiation <strong>of</strong>the Continental Crust. Cambridge University Press, Cambridge,553 pp.Brown, M. & Solar, G. S., 1998a. Granite ascent <strong>and</strong> emplacementduring contractional deformation in convergent orogens.Journal <strong>of</strong> Structural Geology, 20, 1365–1393.Brown, M. & Solar, G. S., 1998b. Shear zone systems <strong>and</strong> melts:feedback relations <strong>and</strong> self-organization in orogenic belts.Journal <strong>of</strong> Structural Geology, 20, 211–227.Brown, M., Averkin, Y., McLellan, E. L. & Sawyer, E. W., 1995.Melt segregation in migmatites. Journal <strong>of</strong> Geophysical Research,100, 15655–15679.Brown, M. A., Brown, M., Carlson, W. D. & Denison, C., 1999.Topology <strong>of</strong> syntectonic melt-flow networks in the deep crust:inferences from three-dimensional images <strong>of</strong> leucosomegeometry in migmatites. American Mineralogist, 84, 1793–1818.Bu¨ sch, W., Schneider, G. & Mehnert, K. R., 1974. Initial meltingat grain boundaries. Part II. Melting in rocks <strong>of</strong> granodioritic,quartz dioritic <strong>and</strong> tonalitic composition. Neues Jahrbuch Fu¨rMineralogie Monatshefte, 8, 345–370.Carlson, W. D., 1989. The significance <strong>of</strong> intergranular diffusionto the mechanism <strong>and</strong> kinetics <strong>of</strong> porphyroblast crystallization.Contributions to Mineralogy <strong>and</strong> Petrology, 103, 1–24.Cashman, K. V. & Ferry, J. M., 1988. Crystal size distribution(CSD) in rocks <strong>and</strong> the kinetics <strong>and</strong> dynamics <strong>of</strong> crystallization.3. Metamorphic crystallization. Contributions to Mineralogy<strong>and</strong> Petrology, 99, 401–415.Cashman, K. V. & Marsh, B. D., 1988. Crystal size distribution(CSD) in rocks <strong>and</strong> the kinetics <strong>and</strong> dynamics <strong>of</strong> crystallization.2. Makaopulu lava lake. Contributions to Mineralogy <strong>and</strong>Petrology, 99, 292–305.Ch<strong>and</strong>rasekhar, S., 1961. Hydrodynamic <strong>and</strong> HydromagneticStability. Oxford University Press, London.Ó 2007 Blackwell Publishing Ltd336


ORIGIN OF FELSIC MIGMATITES 51Cmı´ral, M., Fitz Gerald, J. D., Faul, U. H. & Green, D. H.,1998. A close look at dihedral angles <strong>and</strong> melt geometry inolivine-basalt aggregates: a TEM study. Contributions toMineralogy <strong>and</strong> Petrology, 130, 336–345.Collins, W. J. & Sawyer, E. W., 1996. Pervasive granitoidmagma transfer through the lower-middle crust during noncoaxialcompressional deformation. Journal <strong>of</strong> MetamorphicGeology, 14, 565–579.Connolly, J. A. D., Holness, M. B., Rubie, D. C. & Rushmer, T.,1997. Reaction-induced microcracking: an experimentalinvestigation <strong>of</strong> a mechanism for enhancing anatectic meltsextraction. Geology, 25, 591–594.Cosgrove, J. W., 1997. The influence <strong>of</strong> mechanical anisotropyon the behaviour <strong>of</strong> the lower crust. Tectonophysics, 280,1–14.Daines, M. J. & Kohlstedt, D. L., 1997. Influence <strong>of</strong> deformationon melt topology in peridotites. Journal <strong>of</strong> GeophysicalResearch, 102, 10257–10271.Dallain, C., Schulmann, K. & Ledru, P., 1999. Textural evolutionin the transition from subsolidus annealing to meltingprocess, Velay Dome, French Massif Central. Journal <strong>of</strong>Metamorphic Geology, 17, 61–74.DeH<strong>of</strong>f, R. T., 1991. A geometrically general theory <strong>of</strong> diffusioncontrolled coarsening. Acta Metallica <strong>and</strong> Materialia, 39,2349–2360.DellÕ Angelo, L. N. & Tullis, J., 1988. Experimental deformation<strong>of</strong> partially melted granitic aggregates. Journal <strong>of</strong> MetamorphicGeology, 6, 495–515.DellÕ Angelo, L. N., Tullis, J. & Yund, R. A., 1987. Transitionfrom dislocation creep to melt-enhanced diffusion creep infine-grained granitic aggregates. Tectonophysics, 139, 325–332.DeVore, G. W., 1959. Role <strong>of</strong> minimum interfacial free energy indetermining the macroscopic features <strong>of</strong> minerals assemblages.Journal <strong>of</strong> Geology, 67, 211–227.Dougan, T. W., 1983. Textural relations in melanosomes <strong>of</strong> selectedspecimens <strong>of</strong> migmatitic pelitic schists: implications forleucosome-generating process. Contributions to Mineralogy<strong>and</strong> Petrology, 83, 82–98.Dudek, A., Mateˇjovska´, O. & Suk, M., 1974. Gfo¨ hl orthogneissin the Moldanubicum <strong>of</strong> Bohemia <strong>and</strong> Moravia. Krystalinikum,10, 67–78.Ferre´, E. C., Teyssier, C., Jackson, M., Thill, J. W. & Rainey,E. S. G., 2003. Magnetic susceptibility anisotropy: a newpetr<strong>of</strong>abric tool in migmatites. Journal <strong>of</strong> Geophysical Research,108, B2, doi: 10.1029/2002JB001790Ferre´, E. C., Martı´n-Herna´ndez, F., Teyssier, C. & Jackson, M.,2004. Paramagnetic <strong>and</strong> ferromagnetic anisotropy <strong>of</strong> magneticsusceptibility in migmatites: measurements in high <strong>and</strong> lowfields <strong>and</strong> kinematic implications. Geophysical Journal International,157, 1119–1129.Flinn, D., 1969. Grain contacts in crystalline rocks. Lithos, 2,361–370.Friedl, G., Finger, F., Paquette, J., Quadt, A., McNaughton, N.& Fletcher, I., 2004. Pre-Variscan geological events in theAustrian part <strong>of</strong> the Bohemian Massif deduced from U–Pbzircon ages. International Journal <strong>of</strong> Earth Sciences, 93, 802–823.Fyfe, W. S., 1973. The granulite facies, partial melting <strong>and</strong> theArchaean crust. Royal Society <strong>of</strong> London PhilosophicalTransactions, Series, A, 273, 457–461.Gardien, V., Thompson, A. B., Grujic, D. & Ulmer, P., 1995.Experimental melting <strong>of</strong> biotite + plagioclase + quartz +muscovite assemblage <strong>and</strong> implications for crustal melting.Journal <strong>of</strong> Geophysical Research, 100, 15581–15591.Garlick, S. R. & Gromet, L. P., 2004. Diffusion creep <strong>and</strong> partialmelting in high temperature mylonitic gneisses, Hope Valleyshear zone, New Engl<strong>and</strong> Appalachians, USA. Journal <strong>of</strong>Metamorphic Geology, 22, 45–62.Gleason, G. C., Bruce, V. & Green, H. W., 1999. Experimentalinvestigation <strong>of</strong> melt topology in partially moltenquartzo-feldspathic aggregates under hydrostatic <strong>and</strong> nonhydrostaticstress. Journal <strong>of</strong> Metamorphic Geology, 17, 705–722.Greenfield, J. E., Clarke, G. L. & White, R. W., 1998. A sequence<strong>of</strong> partial melting reactions at Mt. Stafford, centralAustralia. Journal <strong>of</strong> Metamorphic Geology, 16, 363–378.Hasalova´, P., Sˇtı´pska´, P., Powell, R., Schulmann, K., Janousˇek,V. & Lexa, O., 2008a. Transforming mylonitic metagranite byopen-system interactions during melt flow. Journal <strong>of</strong> MetamorphicGeology, doi: 10.1111/j.1525-1314.2007.00744.xHasalova´, P., Janousˇek, V., Schulmann, K., Sˇtı´pska´, P. & Erban,V., 2008b. From orthogneiss to migmatite: geochemicalassessment <strong>of</strong> the melt infiltration model in the Gfo¨ hl Unit(Moldanubian Zone, Bohemian Massif). Lithos, doi: 10.1016/j.lithos.2007.07.021Hickey, K. A. & Bell, T. H., 1996. Syn-deformational graingrowth: matrix coarsening during foliation development <strong>and</strong>regional metamorphism rather than by static annealing.European Journal <strong>of</strong> Mineralogy, 8, 1351–1373.Higgins, M. D., 1998. Origin <strong>of</strong> anorthosite by textural coarsening:quantitative measurements <strong>of</strong> a natural sequence <strong>of</strong>textural development. Journal <strong>of</strong> Petrology, 39, 1307–1323.Holub, F. V., Cocherie, A. & Rossi, P., 1997. Radiometric dating<strong>of</strong> granitic rocks from the Central Bohemian Plutonic Complex(Czech Republic): constraints on the chronology <strong>of</strong>thermal <strong>and</strong> tectonic events along the Moldanubian–Barr<strong>and</strong>ianboundary. Earth <strong>and</strong> Planetary Sciences, 325, 19–26.Hrouda, F., Jelínek, V. & Hrusˇkova´, L., 1990. A package <strong>of</strong>programs for statistical evaluation <strong>of</strong> magnetic anisotropydata using IBM-PC computers. EOS, Transactions, AmericanGeophysical Union, 71, 1289.Hrouda, F., Melka, R. & Schulmann, K., 1994. Periodicalchanges in fabric intensity during simple shear deformation<strong>and</strong> its implications for magnetic susceptibility anisotropy <strong>of</strong>sedimentary <strong>and</strong> volcanic rocks. Acta Universitatis CarolinaeGeologica, 38, 37–56.Jelínek, V., 1978. Statistical processing <strong>of</strong> anisotropy <strong>of</strong> magneticsusceptibility measured on groups <strong>of</strong> specimens. Studia Geophysicaet Geodaetica, 22, 50–62.Jelínek, V., 1981. Characterization <strong>of</strong> magnetic fabric <strong>of</strong> rocks.Tectonophysics, 79, 63–67.Jin, Z., Green, H. W. & Zhou, Y., 1994. Melt topology in partiallymolten mantle peridotite during ductile deformation.Nature, 372, 164–167.John, B. E. & Stu¨ nitz, H., 1997. Evidence for magmatic fracturing<strong>and</strong> small-scale melt segregation during plutonemplacement. In: Processes <strong>of</strong> Melt Segregation (eds Bouchez,J.-L. & Clemens, J.), pp. 55–74, Kluwer Press, Amsterdam.Jordan, P., 1988. The rheology <strong>of</strong> polymineralic rocks: an approach.Geologische Rundschau, 77, 285–294.Jurewicz, S. R. & Watson, E. B., 1984. Distribution <strong>of</strong> partialmelt in a felsic system – the importance <strong>of</strong> surface energy.Contributions to Mineralogy <strong>and</strong> Petrology, 85, 25–29.Kim, C.-S. & Rohrer, G. S., 2004. Geometric <strong>and</strong> crystallographiccharacterization <strong>of</strong> WC surfaces <strong>and</strong> grain boundariesin WC–Co composites. Interface Science, 12, 19–27.Kretz, R., 1966. Grain-size distribution for certain metamorphicminerals in relation to nucleation <strong>and</strong> growth. Journal <strong>of</strong>Geology, 74, 147–173.Kretz, R., 1969. On the spatial distribution <strong>of</strong> crystals in rocks.Lithos, 2, 39–66.Kretz, R., 1994. Metamorphic Crystallization. Wiley, New York.Kro¨ ner, A., Wendt, I., Liew, T. C. et al., 1988. U–Pb zircon <strong>and</strong>Sm–Nd model ages <strong>of</strong> high grade Moldanubian metasediments,Bohemian Massif, Czechoslovakia. Contributions toMineralogy <strong>and</strong> Petrology, 99, 257–266.Kruse, R. & Stu¨ nitz, H., 1999. Deformation mechanisms <strong>and</strong>phase distribution in mafic high-temperature mylonites fromthe Jotun Nappe, southern Norway. Tectonophysics, 303, 223–249.Kruse, R., Stu¨ nitz, H. & Kunze, A., 2001. Dynamic recrystallizationprocesses in plagioclase porphyroclasts. Journal <strong>of</strong>Structural Geology, 24, 587–589.Ó 2007 Blackwell Publishing Ltd337


52 P. HASALOVÁ ET AL.Lafeber, D., 1963. On the spatial distribution <strong>of</strong> fabric elementsin rock <strong>and</strong> soil fabric. Proceedings <strong>of</strong> the 4th Australia-NewZeal<strong>and</strong> Conference on Soil Mechanics <strong>and</strong> Foundation Engineering,University <strong>of</strong> Adelaide, Adelaide, 11, 185–199.Laporte, D. & Watson, E. B., 1995. Experimental <strong>and</strong> theoreticalconstraints on melt distribution in crustal sources: the effect <strong>of</strong>crystalline anisotropy on melt interconnectivity. ChemicalGeology, 124, 161–184.Laporte, D., Rapaille, C. & Provost, A., 1997. Wetting angle,equilibrium melt geometry, <strong>and</strong> the permeability threshold <strong>of</strong>partially molten crustal protoliths. In: Granite: From Segregation<strong>of</strong> Melt to Emplacement Fabrics (eds Bouchez, J.L.,Hutton, D.H.W. & Stephens, W.E.), pp. 31–54. Kluwer,Dordrecht.Le Breton, N. & Thompson, A. B., 1988. Fluid-absent (dehydration)melting <strong>of</strong> biotite in metapelites in the early stages <strong>of</strong>crustal anatexis. Contributions to Mineralogy <strong>and</strong> Petrology,99, 226–237.Lexa, O., 2003. Numerical approaches in <strong>structural</strong> <strong>and</strong> micro<strong>structural</strong><strong>analyses</strong>, Unpublished PhD Thesis. CharlesUniversity, Prague.Lexa, O., Sˇtı´pska´, P., Schulmann, K., Baratoux, L. & Kro¨ ner,A., 2005. Contrasting textural record <strong>of</strong> two distinct metamorphicevents <strong>of</strong> similar P–T conditions <strong>and</strong> different durations.Journal <strong>of</strong> Metamorphic Geology, 23, 649–666.Lifshitz, I. M. & Slyozov, V. V., 1961. The kinetics <strong>of</strong> precipitationfrom supersaturated solid solutions. Journal <strong>of</strong> Physics<strong>and</strong> Chemistry <strong>of</strong> Solids, 19, 35–50.Lister, J. R. & Kerr, R. C., 1991. Fluid-mechanical models <strong>of</strong>crack propagation <strong>and</strong> their application to magma-transportin dykes. Journal <strong>of</strong> Geophysical Research, 96, 10049–10077.Lucas, S. B. & St.Onge, M. R., 1995. Syn-tectonic magmatism<strong>and</strong> the development <strong>of</strong> compositional layering, UngavaOrogen (northern Quebec, Canada). Journal <strong>of</strong> StructuralGeology, 17, 475–491.Mabuchi, M., Jeong, H. G., Hiraga, K. & Higashi, K., 1997.Partial melting at interfaces <strong>and</strong> grain boundaries for highstrain-ratesuperplastic materials. Interface Science, 4, 357–368.Marchildon, N. & Brown, M., 2001. Melt Segregation in latesyn-tectonic anatectic migmatites: an example from the OnawaContact Aureole, Maine, USA. Physics <strong>and</strong> Chemistry <strong>of</strong>the Earth, 26, 225–229.Marchildon, N. & Brown, M., 2002. Grain-scale melt distributionin two contact aureole rocks: implications for controls onmelt localization <strong>and</strong> deformation. Journal <strong>of</strong> MetamorphicGeology, 20, 381–396.Marchildon, N. & Brown, M., 2003. Spatial distribution <strong>of</strong> meltbearingstructures in anatectic rocks from Southern Brittany,France: implications for melt transfer at grain-to orogen-scale.Tectonophysics, 364, 215–235.Marsh, B. D., 1988. Crystal size distribution (CSD) in rocks <strong>and</strong>the kinetics <strong>and</strong> dynamics <strong>of</strong> crystallization; 1, Theory. Contributionsto Mineralogy <strong>and</strong> Petrology, 99, 277–291.Martı´n-Hern<strong>and</strong>e´z, F. & Hirth, A. M., 2003. Paramagneticanisotropy <strong>of</strong> magnetic susceptibility in biotite, muscovite <strong>and</strong>chlorite single crystals. Tectonophysics, 367, 13–28.Mateˇjovska´, O., 1974. Petrographic Characteristic <strong>of</strong> GneissSeries between Mohelno <strong>and</strong> Bı´tov, SW Moravia. MS Ge<strong>of</strong>ond,Prague (in Czech).Matte, P., Maluski, H., Rajlich, P. & Franke, W., 1990. Terraneboundaries in the Bohemian Massif: Result <strong>of</strong> large-scaleVariscan shearing. Tectonophysics, 177, 151–170.McKenzie, D., 1984. The generation <strong>and</strong> compaction <strong>of</strong> partiallymolten rock. Journal <strong>of</strong> Petrology, 25, 713–765.McLellan, E. L., 1983. Contrasting textures in metamorphic <strong>and</strong>anatectic migmatites – an example from the Scottish Caledonides.Journal <strong>of</strong> Metamorphic Geology, 1, 241–262.Mehnert, K. R., 1971. Migmatites <strong>and</strong> the Origin <strong>of</strong> GraniticRocks. Elsevier, Amsterdam.Mehnert, K. R., Bu¨ sch, W. & Schneider, G., 1973. Initial meltingat grain boundary <strong>of</strong> quartz <strong>and</strong> feldspar in gneisses <strong>and</strong>granulites. Neues Jahrbuch Fu¨r Mineralogie Monatshefte, 4,165–183.Moazzen, M. & Modjarrad, M., 2005. Contact metamorphism<strong>and</strong> crystal size distribution studies in the Shivar aureole, NWIran. Geological Journal, 40, 499–517.Olsen, T. S. & Kohlstedt, D. L., 1984. Analysis <strong>of</strong> dislocations insome naturally deformed plagioclase feldspars. Physics <strong>of</strong> TheEarth <strong>and</strong> Planetary Interiors, 11, 153–160.Panozzo, R., 1983. The two-dimensional analysis <strong>of</strong> shape fabricusing projections <strong>of</strong> lines in a plane. Tectonophysics, 95, 279–295.Paterson, M. S., 2001. A granular flow theory for the deformation<strong>of</strong> partially molten rock. Tectonophysics, 335, 51–61.Peterson, T. D., 1996. A refined technique for measuring crystalsize distributions in thin section. Contributions to Mineralogy<strong>and</strong> Petrology, 124, 395–405.Petford, N., 1995. Segregation <strong>of</strong> tonalitic–trondhjemitic melts inthe continental crust: the mantle connection. Journal <strong>of</strong> GeophysicalResearch, 100, 15735–15743.Powell, R., Holl<strong>and</strong>, T. & Worley, B., 1998. Calculating phasediagrams involving solid solutions via non-linear equations,with examples using THERMOCALC. Journal <strong>of</strong> MetamorphicGeology, 16, 577–588.Racek, M., Sˇtı´pska´, P., Pitra, P., Schulmann, K. & Lexa, O.,2006. Metamorphic record <strong>of</strong> burial <strong>and</strong> exhumation <strong>of</strong> orogeniclower <strong>and</strong> middle crust: a new tectonothermal model forthe Drosendorf window (Bohemian Massif, Austria). Mineralogy<strong>and</strong> Petrology, 86, 221–251.Ramberg, H., 1981. Gravity, Deformation <strong>and</strong> the Earth’s Crustin Theory, Experiments <strong>and</strong> Geological Applications. AcademicPress, London.R<strong>and</strong>olph, A. D. & Larson, M. A., 1971. Theory <strong>of</strong> ParticleProcesses. Academic Press, New York.Rosenberg, C., 2001. Deformation <strong>of</strong> partially molten granite: areview <strong>and</strong> comparison <strong>of</strong> experimental <strong>and</strong> natural casestudies. International Journal <strong>of</strong> Earth Sciences, 90, 60–76.Rosenberg, C. & Berger, A., 2001. Syntectonic melt pathways ingranitic gneisses, <strong>and</strong> melt-induced transitions in deformationmechanisms. Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth (A), 26, 287–293.Rosenberg, C. L. & H<strong>and</strong>y, M. R., 2000. Syntectonic meltpathways during simple shearing <strong>of</strong> a partially molten rockanalogue (norcamphor-benzamide). Journal <strong>of</strong> GeophysicalResearch, 105, 3135–3149.Rosenberg, C. & H<strong>and</strong>y, M. R., 2001. Mechanisms <strong>and</strong> orientation<strong>of</strong> melt segregation paths during pure shearing <strong>of</strong> apartially molten rock analogue (norcamphor-benzamide).Journal <strong>of</strong> Structural Geology, 23, 1917–1932.Rosenberg, C. L. & Riller, U., 2000. Partial-melt topology instatically <strong>and</strong> dynamically recrystallized granite. Geology, 28,7–10.Rosenberg, C. L. & H<strong>and</strong>y, M. R., 2005. Experimentaldeformation <strong>of</strong> partially melted granite revisited: implicationsfor the continental crust. Journal <strong>of</strong> Metamorphic Geology, 23,19–28.Sakai, T. & Jonas, J. J., 1984. Dynamic recrystallization:mechanical <strong>and</strong> micro<strong>structural</strong> consideration. Acta Metallurgica,32, 198–209.Sawyer, E. W., 1994. Melt segregation in the continental crust.Geology, 22, 1019–1022.Sawyer, E. W., 1998. Formation <strong>and</strong> evolution <strong>of</strong> granite magmasduring crustal reworking: the significance <strong>of</strong> diatexites.Journal <strong>of</strong> Petrology, 39, 1147–1167.Sawyer, E. W., 1999. Criteria for the recognition <strong>of</strong> partialmelting. Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth, 24, 269–279.Sawyer, E. W., 2001. Melt segregation in the continental crust:distribution <strong>and</strong> movement <strong>of</strong> melt in anatectic rocks. Journal<strong>of</strong> Metamorphic Geology, 19, 291–309.Schulmann, K., Melka, R., Lobkowicz, M., Ledru, P., Lardeaux,J. M. & Autran, A., 1994. Contrasting styles <strong>of</strong>deformation during progressive nappe stacking at the south-Ó 2007 Blackwell Publishing Ltd338


ORIGIN OF FELSIC MIGMATITES 53eastern margin <strong>of</strong> the Bohemian Massif (Thaya Dome).Journal <strong>of</strong> Structural Geology, 16, 355–370.Schulmann, K., Kro¨ ner, A., Hegner, E. et al., 2005. Chronologicalconstraints on the pre-orogenic history, burial <strong>and</strong>exhumation <strong>of</strong> deep-seated rocks along the eastern margin <strong>of</strong>the Variscan orogen, Bohemian Massif, Czech Republic.American Journal <strong>of</strong> Science, 305, 407–448.Seng, H., 1936. Die Migmatitfrage und der Mechanismus parakristallinerPrägung. Geologische Rundschau, 27, 471–492.Spry, A. H., 1969. Metamorphic Textures. Pergamon Press,Oxford.Sˇtı´pska´, P., Schulmann, K. & Kro¨ ner, A., 2004. Vertical extrusion<strong>and</strong> middle crustal spreading <strong>of</strong> omphacite granulite: amodel <strong>of</strong> syn-convergent exhumation (Bohemian Massif,Czech Republic). Journal <strong>of</strong> Metamorphic Geology, 22, 179–198.Stu¨ nitz, H., 1998. Syndeformational recrystallization – dynamicor compositionally induced? Contributions to Mineralogy <strong>and</strong>Petrology, 131, 219–236.Tajcˇmanova´, L., Konopa´sek, J. & Schulmann, K., 2006. Thermalevolution <strong>of</strong> the orogenic lower crust during exhumation withina thickened Moldanubian root <strong>of</strong> the Variscan belt <strong>of</strong> CentralEurope. Journal <strong>of</strong> Metamorphic Geology, 24, 119–134.Tanner, D. C., 1999. The scale-invariant nature <strong>of</strong> migmatitefrom the Oberpfalz, NE Bavaria <strong>and</strong> its significance for melttransport. Tectonophysics, 302, 297–305.Turner, F. J. & Weiss, L., 1963. Structural Analysis <strong>of</strong> MetamorphicTectonites. McGraw Hill, New York.Twiss, R. J., 1977. Theory <strong>and</strong> applicability <strong>of</strong> a recrystallizedgrain size paleopiezometer. Pure <strong>and</strong> Applied Geophysics, 115,227–244.Urban, M., 1992. Kinematics <strong>of</strong> the Variscan thrusting in theEastern Moldanubicum (Bohemian Massif, Czechoslovakia):evidence from the Na´meˇsˇtÕ granulite massif. Tectonophysics,201, 371–391.V<strong>and</strong>erhaeghe, O., 1999. Pervasive melt migration from migmatitesto leucogranite in the Shuswap metamorphic corecomplex, Canada: control <strong>of</strong> regional deformation. Tectonophysics,312, 35–55.Vernon, R. H., 1974. Controls <strong>of</strong> mylonitic compositional layeringduring non-cataclastic ductile deformation. GeologicalMagazine, 111, 121–123.Vernon, R. H., 1976. Metamorphic Processes: Reactions <strong>and</strong>Microstructure Development. Wiley, New York.Vernon, R. H. & Paterson, S. R., 2001. Axial-surface leucosomesin anatectic migmatites. Tectonophysics, 335, 183–192.Vielzeuf, D. & Holloway, J. R., 1988. Experimental determination<strong>of</strong> the fluid-absent melting relations in the pelitic systemconsequencesfor crustal differentiation. Contributions toMineralogy <strong>and</strong> Petrology, 98, 257–276.Vigneresse, J.-L., Barbey, P. & Cuney, M., 1996. Rheologicaltransitions during partial melting <strong>and</strong> crystallization withapplication to felsic magma segregation <strong>and</strong> transfer. Journal<strong>of</strong> Petrology, 37, 1579–1600.Walte, N. P., Bons, P. D., Passchier, C. W. & Koehn, D., 2003.Disequilibrium melt distribution during static recrystallization.Geology, 31, 1009–1012.Walte, N. P., Bons, P. D. & Passchier, C. W., 2005. Deformation<strong>of</strong> melt-bearing systems – insight from in situ grain-scaleanalogue experiments. Journal <strong>of</strong> Structural Geology, 27,1666–1679.Waters, D. J. & Lovegrove, D. P., 2002. Assessing the extent <strong>of</strong>disequilibrium <strong>and</strong> overstepping <strong>of</strong> prograde metamorphicreactions in metapelites from the Bushveld aureole. Journal <strong>of</strong>Metamorphic Geology, 20, 135–149.Weinberg, R. F. & Searle, M. P., 1998. The Pangong InjectionComplex, Indian Karakoram: a case <strong>of</strong> pervasive granite flowthrough hot viscous crust. Journal <strong>of</strong> Geological Society, 155,883–891.Wickham, S. M., 1987. The segregation <strong>and</strong> emplacement <strong>of</strong>granitic magmas. Journal <strong>of</strong> the Geological Society <strong>of</strong> London,144, 281–297.Willaime, C. & G<strong>and</strong>ais, M., 1977. Electron microscope study <strong>of</strong>plastic defects in experimentally deformed alkali feldspars.Bulletin de la Societe Francaise de Mineralogie et de Cristallographie,100, 263–271.Willaime, C., Christie, J. M. & Kovacs, M.-P., 1979. Experimentaldeformation <strong>of</strong> K-feldspar single crystals. Bulletin deMineralogie, 102, 168–177.Zapletal, K., 1990. Low-field susceptibility anisotropy <strong>of</strong> somebiotite crystals. Physics <strong>of</strong> the Earth <strong>and</strong> Planetary Interiors,63, 85–97.Za´vada, P., Schulmann, K., Konopa´sek, J., Ulrich, S. & Lexa,O., 2007. Extreme ductility <strong>of</strong> feldspar aggregates – meltenhanced grain boundary sliding (GBS) <strong>and</strong> creep failure:rheological implications for felsic lower crust. Journal <strong>of</strong>Geophysical Research, doi: 10.1029/2006JB004820.2007Received 14 February 2006; revision accepted 25 September 2007.SUPPLEMENTARY MATERIALThe following material is available online at http://www.blackwell-synergy.com:Appendix S1 Digitalized thin sections (separatedplagioclase <strong>and</strong> K-feldspar domains) that have beenused for the quantitative <strong>analyses</strong>.Ó 2007 Blackwell Publishing Ltd339


J. metamorphic Geol., 2011, 29, 103–130 doi:10.1111/j.1525-1314.2010.00911.xOrigin <strong>of</strong> felsic granulite microstructure by heterogeneousdecomposition <strong>of</strong> alkali feldspar <strong>and</strong> extreme weakening <strong>of</strong>orogenic lower crust during the Variscan orogenyJ. FRANĚK, 1,2 K. SCHULMANN, 2 O. LEXA, 1,3 S. ULRICH, 4 P. ŠTÍPSKÁ, 2 J. HALODA 1 AND P. TÝCOVÁ 11 Czech Geological Survey, Klárov 3, 118 21 Prague, Czech Republic (honzaf2@seznam.cz)2 Institut de Physique du Globe de Strasbourg, IPGS – UMR 7516, CNRS et Université de Strasbourg (EOST), 1 Rue Blessig,67084 Strasbourg, France3 Institute <strong>of</strong> Petrology <strong>and</strong> Structural Geology, Charles University, Albertov 6, 128 43 Prague, Czech Republic4 Institute <strong>of</strong> Geophysics, Czech Academy <strong>of</strong> Sciences, Boční II⁄ 1401, 141 31 Prague, Czech RepublicABSTRACTThis study answers the question <strong>of</strong> origin <strong>and</strong> evolution <strong>of</strong> a granulitic microstructure typicallydeveloped in felsic granulites <strong>of</strong> the European Variscan belt. It shows that the precursor <strong>of</strong> the Variscanfelsic granulites was a high-pressure alkali feldspar-rich coarse-grained layered orthogneiss. Its S1subhorizontal layering is defined by the alignment <strong>of</strong> alkali feldspar porphyroclasts alternating withmonomineralic b<strong>and</strong>s <strong>of</strong> quartz <strong>and</strong> b<strong>and</strong>s rich in plagioclase <strong>and</strong> garnet. The alkali feldsparporphyroclasts contain inclusions <strong>of</strong> quartz, garnet, kyanite, biotite <strong>and</strong> rutile, reflecting peak P–Tconditions <strong>of</strong> 1.6–1.8 GPa <strong>and</strong> 850 °C during S1 formation. Superimposed steep folds <strong>and</strong> steepcleavage, S2, are associated with recrystallization <strong>of</strong> alkali feldspar, plagioclase <strong>and</strong> quartz, <strong>and</strong> garnetchemistry modifications that correspond to 0.9–1.0 GPa <strong>and</strong> 800 °C. During exhumation, involving0.8 GPa decompression <strong>and</strong> cooling, the probably perthitic alkali feldspar underwent an unusualprocess <strong>of</strong> heterogeneous decomposition along irregular reaction fronts forming a fine-grained matrixcomposed <strong>of</strong> plagioclase <strong>and</strong> K-feldspar grains. Regular grain distributions in the matrix, nucleationdominatedcrystal size distribution <strong>and</strong> preservation <strong>of</strong> lattice orientation <strong>of</strong> the parental perthitecrystals are all explained by a discontinuous precipitation process. This heterogeneous decomposition <strong>of</strong>alkali feldspar solid solution is controlled by chemically <strong>and</strong> strain induced grain-boundary migration.During exhumation <strong>and</strong> decompression, the fine-grained matrix underwent viscous deformation,forming the typical microstructure <strong>of</strong> the Variscan granulites. R<strong>and</strong>om phase distributions, minorcoarsening <strong>and</strong> feldspar textures are interpreted as a result <strong>of</strong> strain s<strong>of</strong>tening due to diffusion creepaccommodatedgrain-boundary sliding. Subordinate large quartz ribbons were rheologically strongerthan the feldspar-dominated matrix due to the activity <strong>of</strong> different deformational mechanisms. Finally,in mid-crustal levels, the subvertical structure was overprinted by a perpendicular steep fabric associatedwith the growth <strong>of</strong> sillimanite, heterogeneous hydration <strong>and</strong> local partial melting, development <strong>of</strong>aggregate phase distributions <strong>and</strong> significant coarsening. This evolution is accompanied with thedevelopment <strong>of</strong> a strong lattice preferred orientation <strong>of</strong> quartz, K-feldspar <strong>and</strong> plagioclase, reflecting aswitch to dislocation creep mechanism <strong>and</strong> a general hardening <strong>of</strong> the granulites under amphibolitefacies conditions.Key words: felsic granulites; Moldanubian domain; quantitative micro<strong>structural</strong> analysis; quartz <strong>and</strong>feldspar rheology; Variscan belt.INTRODUCTIONFelsic high-pressure (HP) granulites represent the mostabundant deep crustal rocks exposed in the PalaeozoicVariscan belt in Europe <strong>and</strong> form large accumulations<strong>of</strong> the orogenic lower crust, e.g. in the Vosges Massif(Gayk & Kleinschrodt, 2000) <strong>and</strong> the BohemianMassif (OÕBrien & Carswell, 1993), where the classicexample is the Saxony Granulite Massif (Behr, 1961).These rocks are characterized by a fine-grainedrecrystallized matrix <strong>of</strong> K-feldspar <strong>and</strong> plagioclasecontaining kyanite, garnet <strong>and</strong> quartz ribbons. Thesehighly deformed rocks have attracted the interest<strong>of</strong> <strong>structural</strong> geologists for decades (e.g. Behr, 1961;Lister & Dornsiepen, 1982), who have studied thequartz textures <strong>and</strong> considered the granulite microstructurea testimony to deep crustal flow. However,despite the generally accepted opinion that granulitesrepresent deep-seated tectonites reflecting deformationin the deep crust, the parental rocks as well as physicalconditions <strong>and</strong> mechanism <strong>of</strong> deformation remainmatters <strong>of</strong> discussion. Therefore, the key issue <strong>of</strong> thiswork is underst<strong>and</strong>ing the formation <strong>of</strong> the granulitemicrostructures <strong>and</strong> related deformation mechanismsÓ 2010 Blackwell Publishing Ltd 103341


104 J. FRANĚK ET AL.in terms <strong>of</strong> thermal conditions <strong>and</strong> deformationalprocesses. Unravelling the complex deformationalbehaviour <strong>of</strong> the felsic granulites should assist broaderconsiderations <strong>of</strong> rheology <strong>and</strong> mechanical behaviour<strong>of</strong> orogenic lower crust during various stages <strong>of</strong>deformation <strong>and</strong> exhumation.The deformation mechanisms governing rheology <strong>of</strong>lower crustal rocks are <strong>of</strong>ten hard to determine due tochanges <strong>of</strong> the microstructures during exhumation,recrystallization in later deformations or retrogression(e.g. Rutter & Brodie, 1992). At granulite faciesconditions, the ductile strain is usually accommodatedby dislocation creep, or diffusion creep that may becomplemented by grain-boundary sliding (GBS)(Martelat et al., 1999; Garlick & Gromet, 2004),each mechanism having variable importance. Grainboundarydiffusion or the presence <strong>of</strong> silicate meltsmay favour the diffusion creep or GBS according tonew results <strong>of</strong> Schulmann et al. (2008) <strong>and</strong> Za´vadaet al. (2007). The GBS may then operate at the expense<strong>of</strong> other mechanisms <strong>and</strong> promote granular flow,particularly in fine-grained rocks (Za´vada et al., 2007;Schulmann et al., 2008).Feldspar, the main constituent <strong>of</strong> the felsic granulites,is an essential component <strong>of</strong> the EarthÕs crust.Nevertheless, its deformational behaviour is not yetfully understood, mainly due to complex solid solutionmixing <strong>and</strong> variation in crystallographic structure withcooling (e.g. Ribbe, 1983; Putnis et al., 2003; Abartet al., 2009). A common process is exsolution withcooling, which modifies the rheological behaviour <strong>of</strong>alkali feldspar <strong>and</strong> may lead to drastic weakening <strong>of</strong>the orogenic lower crust (Schulmann et al., 2008).Feldspar recrystallization, driven mainly by chemicaldisequilibrium <strong>of</strong> the Or–Ab–An solid solution,has been studied in natural examples (Stu¨ nitz, 1998;Putnis, 2002) or experimentally (Stu¨ nitz & Tullis,2001), but these studies have focused mainly onmedium-temperature water-assisted processes below500 °C. At these conditions the original chemicallyunstable feldspar undergoes dissolution <strong>and</strong> precipitatesas two separate feldspars <strong>of</strong> different composition.Such results cannot be easily extrapolated to thegranulite water under-saturated HT conditions <strong>of</strong> atleast 850 °C (OÕBrien & Ro¨ tzler, 2003; Sˇtı´pska´ &Powell, 2005), where the available fluid is representedby silicate melt, rather than water.In order to address the above-mentioned aspects, wepresent a micro<strong>structural</strong> analysis <strong>of</strong> exceptionallypreserved samples <strong>of</strong> lower-crustal felsic granulites froman 8.5 · 2.5 km domain (Franeˇk et al., 2006, 2011) withwell-preserved granulite facies fabrics. These rocks formpart <strong>of</strong> the Blansky´ les Granulite Massif (BLG) inthe southern Bohemian Moldanubian domain, whichbelongs to the Variscan collisional chain in centralEurope. The combined micro<strong>structural</strong> <strong>and</strong> petrologicalanalysis shows evidence <strong>of</strong> a complex evolution <strong>of</strong>alkali feldspar rheology during granulite formation <strong>and</strong>exhumation. Changes in ductility are ascribed tochemically <strong>and</strong> deformationally driven recrystallization,variations in grain size as well as changingtemperature <strong>and</strong> the amount <strong>of</strong> interstitial melt.Geological settingThe Bohemian Massif (Fig. 1a,b) represents the easternexposure <strong>of</strong> the Variscan orogen in Europe. Duringthe Variscan orogenesis (380–300 Ma), involvingSaxothuringian oceanic subduction <strong>and</strong> subsequentcontinental underthrusting, a 300-km wide orogenicchain evolved. From the NW to the SE, the followingtectonic sequence is developed (Schulmann et al.,2009): the Saxothuringian domain represented byNeoproterozoic basement covered by Palaeozoicsedimentary rocks, the Tepla´ suture zone <strong>and</strong> thesupra-crustal Tepla´–Barr<strong>and</strong>ian Unit. Further tothe SE, the arc-related granitoid plutons separate theTepla´–Barr<strong>and</strong>ian folded sedimentary rocks from thehigh-grade Moldanubian Zone, which shows widespreadanatexis <strong>and</strong> contains slices <strong>of</strong> lower-crustal <strong>and</strong>mantle rocks. This pervasively deformed root domain isfurther to the east bounded by the Brunia microplate(e.g. Schulmann et al., 2005), which is only marginallyaffected by Variscan tectonometamorphic processes.The Moldanubian Zone consists <strong>of</strong> middle- <strong>and</strong>lower-crustal segments, <strong>of</strong>fering an excellent opportunityto examine evidence <strong>of</strong> the exhumation processesoperating in a collisional setting. The exhumed lowercrust, designated as the Gfo¨ hl Unit (Fuchs, 1976), isrepresented by felsic granulites <strong>and</strong> anatectic gneissesthat enclose small bodies <strong>of</strong> mafic granulites, mantlerocks <strong>and</strong> eclogites. The mid-crustal paragneiss-dominatedlevel has been divided according to the prevailinglithology into the Monotonous Group, with only limitedcontent <strong>of</strong> intercalations, such as amphibolites orquartzites, <strong>and</strong> the Varied Group, bearing a largeproportion <strong>of</strong> intercalated amphibolites, quartzites<strong>and</strong> marbles (Fuchs, 1976; Matte et al., 1990). Thestudied BLG is the largest granulite body in SouthernBohemia, <strong>and</strong> belongs to the Gfo¨ hl Unit, which islocated in a complex stack between the Monotonous<strong>and</strong> Varied groups, being accompanied by severalneighbouring granulite bodies (Fig. 2).Previous studies <strong>of</strong> South Bohemian granulitesP–T estimates (Fig. 3) <strong>of</strong> peak metamorphic conditionshave been calculated by various authors using eitherconventional thermobarometry yielding 1000 °C ⁄1.6 GPa (e.g. Vra´na, 1989; OÕBrien & Seifert, 1992;Carswell & OÕBrien, 1993; Cooke, 2000), thermodynamic<strong>modelling</strong> in THERMOCALC s<strong>of</strong>tware that yieldsa maximum <strong>of</strong> 850 °C ⁄ 1.6–1.8 GPa (Sˇtípska´ & Powell,2005) or TWEEQU yielding 970–1000 °C ⁄ 1.6–1.7 GPa(Kro¨ ner et al., 2000). The conditions for the amphibolitefacies overprint are estimated to 700–800 °C <strong>and</strong>0.5–0.8 GPa (Kro¨ ner et al., 2000; Sˇtı´pska´ & Powell,2005; Verner et al., 2007).Ó 2010 Blackwell Publishing Ltd342


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 105(a)(b)Fig. 1. (a) Position <strong>of</strong> the Bohemian Massif in the Variscan chain in Europe. (b) Simplified geological architecture <strong>of</strong> the BohemianMassif. Rectangle marks the study area depicted in Fig. 2. Modified after Franke (2000).Radiometric ages ascribed to protolith crystallization<strong>of</strong> the South Bohemian felsic granulites give between469 ± 4 <strong>and</strong> 357 ± 2 Ma using the U–Pb method onzircon (Wendt et al., 1994; Kro¨ ner et al., 2000). Subsequentpeak HP metamorphism took place between351 ± 6 Ma (Wendt et al., 1994) <strong>and</strong> 341 ± 3Ma(Kro¨ ner et al., 2000). Retrogression under amphibolitefacies conditions proceeded immediately after exhumationto mid-crustal levels, between 340 ± 3 <strong>and</strong>338 ± 3 Ma (both U–Pb on zircon, Kro¨ ner et al.,2000). Cooling below 500 °C is constrained by the331 ± 1 Ma 40 Ar– 39 Ar age <strong>of</strong> hornblende from anamphibolite adjacent to the BLG (Kosˇler et al., 1999).Protolith to the felsic granulites is still debated, butthe majority <strong>of</strong> studies consider the protolith to havebeen a granitic igneous rock (e.g. Fiala et al., 1987;Jakesˇ, 1997; Kotkova´ & Harley, 1999; Finger et al.,2003; Janousˇek et al., 2004, 2006; Tropper et al., 2005;Janousˇek & Holub, 2007). Janousˇek et al. (2004) suggestedan Ordovician granitic protolith with a modelage <strong>of</strong> c. 450 Ma, which is supported by the radiometricU–Pb zircon ages <strong>of</strong> Kro¨ ner et al. (2000) <strong>and</strong>Friedl et al. (2003).Franeˇk et al. (2006) reported a succession <strong>of</strong> threeductile fabrics in the BLG (Fig. 2). The oldestsubhorizontal fabric S1 is preserved in an 8.5 km ·2.5 km elliptical area <strong>of</strong> the BLG <strong>and</strong> macroscopicallyis defined by an alternation <strong>of</strong> up to 1-cm thickwhite b<strong>and</strong>s formed by recrystallized alkali feldspar,up to 1-cm thick quartz-rich b<strong>and</strong>s <strong>and</strong> 1- to 3-mmthick darker plagioclase–garnet-dominated b<strong>and</strong>s(Fig. 4a,b). Large perthite porphyroclasts, up to acentimetre in diameter, are preserved in the feldsparrichb<strong>and</strong>s. The S1 foliation is folded by metre-scalepassive F2 folds with steeply inclined axial planes<strong>and</strong> pervasively developed penetrative cleavage S2,which is macroscopically characterized by strongsubhorizontal elongation <strong>of</strong> quartz ribbons <strong>and</strong>biotite aggregates. Within the limbs, the S1 compositionallayering is stretched <strong>and</strong> progressively rotatedtowards the N–S striking steep S2 cleavage, whichcontains the subhorizontal L2 stretching <strong>and</strong> minerallineation. In strongly reworked areas, the only relicts<strong>of</strong> the S1 fabric are highly attenuated remnants <strong>of</strong> theS1 compositional layering parallel to S2. Macroscopically,the D2 structures are best seen in the formÓ 2010 Blackwell Publishing Ltd343


106 J. FRANĚK ET AL.Fig. 2. Structural map <strong>of</strong> the Blansky´ les Granulite (modified after Franeˇk et al., 2006). Stereographic projections depict contoureddensities <strong>of</strong> foliation poles <strong>and</strong> dots mark corresponding lineations: S2 – 55 foliations <strong>and</strong> 24 lineations; S3 – west 51 foliations<strong>and</strong> 35 lineations; S3 – east 51 foliations <strong>and</strong> 54 lineations. Note that in all cases the foliation poles are distributed along great circle(s)whereas the lineations plunge subparallel to p-axes corresponding to these great circles.Ó 2010 Blackwell Publishing Ltd344


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 107Fig. 3. P–T estimates from Moldanubiangranulites accompanied by relevantradiometric ages suggest almost isothermalrapid exhumation from peak metamorphicconditions to mid-crustal levels. Resultsfrom south Bohemian bodies emphasized bythick lines. For references, see fig. 3 inFraneˇk et al. (2011).<strong>of</strong> the elongated quartz ribbons <strong>and</strong> biotite aggregatesthat show a strong shape-preferred orientation. Boththe S1 <strong>and</strong> S2 foliations show a mineral assemblage<strong>of</strong> Grt–Ky–Bt–Kfs–Pl–Qtz, indicating HP granulitefacies conditions. Mafic boudins enclosed in S2 arecrosscut by Grt–Ky–Kfs–Pl–Qtz felsic dykes. At themarginal part <strong>of</strong> the elliptical <strong>structural</strong> relict domain,the S2 foliation is folded by outcrop-scale F3 foldswith the formation <strong>of</strong> axial gneissosity S3 (Fig. 4c).This gneissosity penetratively transposes all previousfabrics in the majority <strong>of</strong> the BLG <strong>and</strong> showssynkinematic retrogressive breakdown <strong>of</strong> garnet tobiotite <strong>and</strong> kyanite to sillimanite (Fig. 5f), constrainingthe deformation to mid-crustal conditions. Thedetailed tectonic history <strong>and</strong> exhumation in a formsimilar to a forced diapir are described in Franeˇket al. (2011) <strong>and</strong> Lexa et al. (2011).PETROLOGYAnalytical proceduresThe minerals were analysed using the scanning electronmicroscopes Tescan VEGA\\ XMU at 15 kV <strong>and</strong>0.4 nA at the Strasbourg University <strong>and</strong> CamScan CS3200 at 15 kV <strong>and</strong> 3 nA at the LAREM laboratory <strong>of</strong>the Czech Geological Survey in Prague. Compositionalmaps <strong>of</strong> the feldspar were complemented by spotor area <strong>analyses</strong> <strong>of</strong> representative places. Mineralformulae <strong>and</strong> end-member proportions were calculatedusing the NORM s<strong>of</strong>tware (Ulmer, 1986). Micro<strong>structural</strong>types are qualitatively described accordingto the deformational fabrics (S1–S3) <strong>and</strong> subsequentlythey were quantitatively evaluated using theprogram PolyLX (Lexa et al., 2005) <strong>and</strong> by electronÓ 2010 Blackwell Publishing Ltd345


108 J. FRANĚK ET AL.(a)perthite porphyroclasts were oriented with respect tothe penetrative S2 foliation <strong>and</strong> L2 lineation, representingXY, XZ <strong>and</strong> YZ sections.(b)(c)Petrography <strong>of</strong> the granulites with the S1 fabricRocks at outcrop H296 are white-grey, fine-grainedgranulites composed <strong>of</strong> alkali feldspar, quartz,plagioclase <strong>and</strong> garnet (0.2 mm), with minor biotite,kyanite (0.3 mm) <strong>and</strong> porphyroclasts (up to 17 mm)<strong>of</strong> perthitic alkali feldspar. The rocks record evidence<strong>of</strong> the complete granulite facies <strong>structural</strong> evolutiondescribed above. Microscopically, S1 contains discontinuousb<strong>and</strong>s or lenses dominated by plagioclase thatcontain numerous garnet, kyanite, some quartz <strong>and</strong>biotite (Fig. 5b,d). The almost monomineralic S1quartz b<strong>and</strong>s are recrystallized into elongated S2ribbons <strong>and</strong> only quartz accumulation into stripesindicates the original S1 layering (Fig. 5c). Less elongatedquartz grains are rarely preserved in pressureshadows <strong>of</strong> perthite porphyroclasts. Large perthiteporphyroclasts with numerous lensoidal to lamellaroligoclase exsolutions are recrystallized at their grainboundaries to a mixture <strong>of</strong> small K-feldspar grains(0.063 mm) with rare perthitic exsolution lamellae<strong>and</strong> oligoclase grains (0.047 mm). The feldspardominatedb<strong>and</strong>s with rare garnet are predominantlycomposed <strong>of</strong> this K-feldspar–plagioclase mixture withminor quartz (0.055 mm), ascribed to the D2recrystallization process (Figs 5a & 6a,b). In therecrystallized matrix composed <strong>of</strong> K-feldspar, plagioclase<strong>and</strong> quartz, minor garnet, kyanite, biotite, rutile,ilmenite, zircon, monazite <strong>and</strong> apatite also occur(Fig. 6g,h).The perthitic porphyroclasts (up to 17 mm across)contain inclusions <strong>of</strong> quartz, garnet <strong>and</strong> kyanite, <strong>and</strong>more rarely biotite, rutile, ilmenite, zircon, monazite,apatite <strong>and</strong> Fe-sulphide (Fig. 6a–f). Quartz inclusions(up to 1 mm across) commonly consist <strong>of</strong> a singlecrystal with oval or euhedral shape. Garnet (up to1.2 mm across) enclosed in perthite is euhedral <strong>and</strong> iscommonly surrounded by a thin corona <strong>of</strong> plagioclase.Subhedral kyanite inclusions are also separated fromperthite by a thick plagioclase corona. Biotite inclusionsin perthite have short prismatic habits, <strong>and</strong> are inplaces partially retrogressed to chlorite.Fig. 4. Field photographs <strong>of</strong> (a <strong>and</strong> b) passive F2 folds depictingpenetrative development <strong>of</strong> S2 axial cleavage across the foldedS1 compositional layering. (c) F3 folds reworking at amphibolitefacies conditions the granulite facies S2 mylonite.back-scattered diffraction (EBSD). To study the transitionfrom the S1 fabric into the S2 cleavage <strong>and</strong>the associated P–T path, the rocks were collected fromone locality within the elliptical relict <strong>structural</strong>domain (outcrop H296; Fig. 2; 48°51¢52.343¢¢N, 14°19¢14.135¢¢E). Sixty thin sections containing 350 largeMineral chemistryTo specify the P–T path for the S1 <strong>and</strong> S2 fabrics, onesample (H296-S1A) with the S1 layering affected bythe S2 cleavage, as described above, was analysed indetail. It contains garnet, kyanite, perthitic K-feldspar,plagioclase, quartz, biotite, rutile, ilmenite, apatite <strong>and</strong>zircon. The composition <strong>of</strong> the minerals in the individualb<strong>and</strong>s is similar. Large perthite grains includequartz, garnet, kyanite, rutile, ilmenite, apatite <strong>and</strong>zircon. Garnet included in large perthite <strong>and</strong> garnetfrom the matrix are zoned from core to rim withÓ 2010 Blackwell Publishing Ltd346


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 109(a)(b)(c)(d)(e)(f)Fig. 5. BSE images characterizing the three examined fabrics. (a) S1 compositional layering. (b) S1 rotated subparallel to S2 in afold limb; note quartz shapes in perthite <strong>and</strong> in the matrix. (c) Quartz-rich S1 b<strong>and</strong> recrystallized to S2 ribbons. (d) Grt–Ky–Pl-rich S1b<strong>and</strong> affected by the D2 deformation. (e) Penetratively developed S2 granulitic mylonite; note plagioclase corona around kyanite.(f) Amphibolite facies S3 fabric; note biotite with sillimanite growing around garnet <strong>and</strong> a relict kyanite.Ó 2010 Blackwell Publishing Ltd347


110 J. FRANĚK ET AL.(a)(b)(c) (d) (f)(e)(g)(h)Fig. 6. BSE images <strong>of</strong> S1 layering. (a–f) Inclusions in the large perthite grains reflect peak metamorphic assemblage. A microchemicalpr<strong>of</strong>ile across garnet in (f) is presented in Fig. 7. (g) Grt–Ky–Pl-rich S1 b<strong>and</strong> containing biotite, which grows parallel to S2. (h)K-feldspar-rich b<strong>and</strong> enclosing kyanite with typical well-developed plagioclase corona.decreasing grossular, increasing pyrope <strong>and</strong> alm<strong>and</strong>ine,<strong>and</strong> flat X Fe (Alm 0.52Þ0.61 Grs 0.23Þ0.04 Prp 0.22Þ0.28Sps 0.00–0.01 ; X Fe =Fe⁄ (Fe + Mg) = 0.70, Fig. 7b).The X Fe <strong>of</strong> biotite in the matrix ranges from 0.33 to 0.35with Ti = 0.21)0.26 (pfu based on 22 oxygen).Recrystallized K-feldspar <strong>and</strong> plagioclase are zonedfrom core to rim, with Or 77 fi 97 Ab 23 fi 03 An 0.0 <strong>and</strong>Or 02 fi 01 Ab 81 fi 77 An 18 fi 21 , respectively. Kyanite,Ó 2010 Blackwell Publishing Ltd348


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 111(a)(b)Fig. 7. (a) Ternary plot <strong>of</strong> feldspar compositions acquired from microchemical <strong>analyses</strong> <strong>of</strong> perthite recrystallization <strong>and</strong> S2 fabric.Similar bulk compositions <strong>of</strong> parental perthites <strong>and</strong> shielded matrix are contrasted against areal <strong>analyses</strong> <strong>of</strong> deformed matrix. Zoning<strong>of</strong> plagioclase <strong>and</strong> to a lesser extent also <strong>of</strong> K-feldspar are highlighted in the description <strong>of</strong> data clusters. (b) Microchemical pr<strong>of</strong>ileacross garnet enclosed in large perthite showing a wide flat grossular-rich core, gradual compositional zoning at both rims <strong>and</strong> almostconstant X Fe across the whole crystal.both included in perthite <strong>and</strong> in the matrix, issurrounded by a plagioclase corona.The studied sample H296-S1A lacks biotite inclusionsin the perthite porphyroclasts, <strong>and</strong> because biotite<strong>and</strong> garnet inclusions in perthite are generally rare,the other five samples from the same macroscopicallyhomogeneous outcrop were analysed to test whetherthere is a difference in composition between biotite <strong>and</strong>garnet inclusions in perthite <strong>and</strong> those in the matrix.The biotite X Fe ranges from 0.23 to 0.40 <strong>and</strong>Ti = 0.15)0.34 pfu, <strong>and</strong> there is no clear differencebetween biotite inclusions <strong>and</strong> biotite in the matrix(Table 1). Garnet in other thin sections shows the samezoning as that from the sample H296-S1A, but additionallya decrease in X Fe to 0.68 at some garnet rimswas detected.The composition <strong>of</strong> the large unzoned perthitegrains was acquired in several samples from the sameoutcrop by areal <strong>analyses</strong> performed by scanning <strong>of</strong>regions containing 100 plagioclase exsolution lamellae.It corresponds on average to 68.2% Or, 27.3% Ab<strong>and</strong> 4.5% An (Fig. 7a). The K-feldspar domains <strong>of</strong>perthite porphyroclasts contain on average 87.1% Or,12.3% Ab <strong>and</strong> 0.6% An. The relict perthite porphyroclastsreveal two generations <strong>of</strong> plagioclase exsolutionlamellae (e.g. Fig. 6c). The coarser exsolution lamellae,presumably older, exhibit elongated braided shapes<strong>and</strong> are 0.8% Or, 76.7% Ab <strong>and</strong> 22.5% An in composition.The second micro- to crypto-perthitic generation<strong>of</strong> exsolution lamellae was not analysed because<strong>of</strong> its small thickness. Detailed feldspar compositionsstudied in relation to the recrystallization mechanismsfor individual deformation episodes are given below.P–T path for S1 <strong>and</strong> S2 fabricsA pseudosection was calculated using THERMOCALC 3.30(Powell & Holl<strong>and</strong>, 1988) <strong>and</strong> DATASET 5.5 (Holl<strong>and</strong> &Powell, 1998; November 2003 upgrade), in the systemNCKFMASHTO (Na 2 O–CaO–K 2 O–FeO–MgO–Al 2 O 3 –SiO 2 –H 2 O–TiO 2 –O) with biotite <strong>and</strong> meltmodels from White et al. (2007), garnet from Dieneret al. (2008), ilmenite from White et al. (2000), feldsparfrom Holl<strong>and</strong> & Powell (2003), white mica fromCoggon & Holl<strong>and</strong> (2002) <strong>and</strong> cordierite from theTHERMOCALC documentation (Powell & Holl<strong>and</strong>, 2004).Mineral composition isopleths x(g, bi) = Fe ⁄ (Fe +Mg), z(g) = Ca ⁄ (Ca + Fe + Mg) <strong>and</strong> t(bi) =XTi(M1) were plotted for garnet <strong>and</strong> biotite. Theanalysed composition <strong>of</strong> the sample H296-S1A (inwt% SiO 2 = 71.98, TiO 2 = 0.42, Al 2 O 3 = 13.53,FeO = 2.1, MnO = 0.03, MgO = 0.73, CaO = 1.93,Ó 2010 Blackwell Publishing Ltd349


112 J. FRANĚK ET AL.Table 1. Representative microchemical <strong>analyses</strong> <strong>of</strong> principal minerals constituting the sample H296-S1 used for calculation <strong>of</strong> the P–Tpseudosection in Fig. 8.Mineral Grt Grt Bt Kfs Kfs Pl Pl PlSample H296-S1A H296-S1A H296-S1A H296-S1A H296-S1A H296-S1A H296-S1A H296-S1AAnalysis 1 g-c 1 g-r 43 44 45 46 5 47Position core matrix rim matrix matrix core rim core rim very rimSiO 2 38.67 38.79 39.40 65.54 64.95 63.46 63.02 67.40TiO2 na na 4.37 0.00 0.00 0.00 0.00 0.00Al 2 O 3 21.51 21.48 16.88 18.66 18.34 22.66 23.06 20.41FeO 24.93 29.93 12.37 0.00 0.00 0.35 0.00 0.00MnO 0.40 0.44 )0.13 0.00 0.00 0.00 0.00 0.00MgO 5.96 7.99 12.95 0.00 0.00 0.00 0.00 0.00CaO 8.49 1.41 0.00 0.00 0.00 3.76 4.47 2.09Na 2 O na na 0.34 2.48 0.32 9.41 9.17 10.31K 2 O na na 9.73 12.87 15.80 0.16 0.32 0.22Total 99.95 100.05 95.91 99.55 99.42 99.79 100.04 100.43Si 3.00 3.01 3.00 3.01 3.03 2.81 2.79 2.96Ti 0.00 0.00 0.25 0.00 0.00 0.00 0.00 0.00Al 1.97 1.97 1.51 1.01 1.01 1.18 1.20 1.06Fe 3+ 0.04 0.00 0.00 0.00 0.00 0.01 0.00 0.00Fe 2+ 1.58 1.94 0.79 0.00 0.00 0.00 0.00 0.00Mn 0.03 0.03 )0.01 0.00 0.00 0.00 0.00 0.00Mg 0.69 0.93 1.47 0.00 0.00 0.00 0.00 0.00Ca 0.70 0.12 0.00 0.00 0.00 0.18 0.21 0.10Na 0.00 0.00 0.05 0.22 0.03 0.81 0.79 0.88K 0.00 0.00 0.94 0.75 0.94 0.01 0.02 0.01Total 8.00 8.00 8.00 5.00 5.00 5.00 5.01 5.00Prp ⁄ An 0.23 0.31 0.00 0.00 0.18 0.21 0.10Alm ⁄ Ab 0.53 0.64 0.23 0.03 0.81 0.77 0.89Grs ⁄ Or 0.23 0.04 0.77 0.97 0.01 0.02 0.01Sps 0.01 0.01XFe (Fe 2+ ) 0.70 0.68 0.35Na 2 O = 2.76, K 2 O = 4.01, P 2 O 5 = 0.15, H 2 O ) =0.22, H 2 O + = 0.56, CO 2 = 0.03) was modified for<strong>modelling</strong> by adding 1 mol.% <strong>of</strong> kyanite to enablea small amount <strong>of</strong> aluminosilicate to be stable atthe estimated peak metamorphic conditions, as it isobserved in the thin section. The amount <strong>of</strong> H 2 O=0.33 mol.% was chosen after the construction <strong>of</strong>T–M H2O sections, as it allows the garnet X Fe = 0.70 tobe stable in the pseudosection (not shown; see Hasalova´et al., 2008a for approach). The major features <strong>of</strong> thepseudosection involve melt being stable above 810 °C,biotite stable up to 1.6 GPa <strong>and</strong> 860 °C, ilmenite stablebelow 1.26 GPa, cordierite stable below 0.7 GPa, <strong>and</strong>a muscovite-out line heading from 750 °C to 810 °Cat 1.35 GPa <strong>and</strong> then continuing to 880 °C <strong>and</strong>2.0 GPa (Fig. 8).The textural features indicate that perthite <strong>and</strong>minerals included in perthite (kyanite, garnet, biotite,rutile <strong>and</strong> quartz) belong to the early assemblage,whereas recrystallized plagioclase <strong>and</strong> K-feldspar,matrix biotite, rutile, ilmenite <strong>and</strong> quartz belong to alater assemblage. Because <strong>of</strong> intense D2 recrystallization,the original plagioclase composition is notknown, but the recrystallized monomineralic plagioclaseb<strong>and</strong>s <strong>and</strong> lenses indicate the stability <strong>of</strong> plagioclasewith alkali feldspar within the S1 structure.Chloritized biotite <strong>and</strong> probably ilmenite in perthiteporphyroclasts are considered as resulting fromre-equilibration. The flat pr<strong>of</strong>ile <strong>of</strong> garnet X Fe isinterpreted as having equilibrated during the laterstage <strong>of</strong> metamorphism, even for the garnet included inperthite porphyroclasts, whereas strong grossularzoning is consistent with the core composition havingbeen preserved from the early metamorphic stage, <strong>and</strong>the rim having equilibrated during the later stage.In the pseudosection, the grossular compositionalisopleth <strong>of</strong> garnet z(g) = 23 occurs outside the biotitestability field (Fig. 8), whereas biotite also belongs tothe observed peak assemblage. However, the biotitepresentfield is only 0.15 GPa <strong>and</strong> 30 °C apart from thecompositional isopleth z(g) = 23, <strong>and</strong> the peak istherefore estimated as the area between these lines,to 1.6)1.8 GPa <strong>and</strong> 850)880 °C (area 1 in Fig. 8).The X Fe <strong>of</strong> the garnet (=70) <strong>and</strong> the grossular contentat the garnet rim (grs = 4) indicate conditions <strong>of</strong>0.9)1.0 GPa <strong>and</strong> 825 °C for the second stage (area2 in Fig. 8). An attempt was made to correlate thebiotite composition with the two metamorphic stages,but even if the compositions fit approximately thecalculated values in the pseudosection, it is likelythat the measured range (X Fe = 0.29)0.36, <strong>and</strong> Ti =0.19)0.26 pfu) reflects partial re-equilibration ondecompression, as does the X Fe <strong>of</strong> garnet, <strong>and</strong> thereforecannot be used for more precise estimation <strong>of</strong>P)T conditions. The feldspar from the pseudosectioncalculations is not used for thermobarometricÓ 2010 Blackwell Publishing Ltd350


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 113Fig. 8. P–T pseudosection calculated in THERMOCALC s<strong>of</strong>tware for a felsic granulite sample H296-S1. The indicated P–T path reflectscoarse-grained alkali feldspar growth at peak conditions followed by decompression during S2 fabric development.estimations, because recent models use significantextrapolations from relatively low pressure <strong>and</strong> temperaturewith respect to the conditions <strong>of</strong> formation <strong>of</strong>the ternary feldspar <strong>of</strong> interest (see discussion inSˇtípska´ & Powell, 2005). The lack <strong>of</strong> a clear differencebetween the chemistry <strong>of</strong> biotite inclusions <strong>and</strong> matrixbiotite in the other studied thin sections is interpretedas being due to partial chemical re-equilibration ondecompression, even within perthite porphyroclasts.A decrease in X Fe to 0.68 at some garnet rims in theother studied thin sections (Fig. 7b) may be explainedby further decompression in liquid-absent conditionswhere x(g) decreases, or by heating. Plagioclase coronasaround kyanite reflect decompression (Tajcˇmanova´et al., 2007; Sˇtípska´ et al., 2010).In summary, the peak P–T conditions 1.6)1.8 GPa<strong>and</strong> 850)880 °C are inferred for the stage when alkalifeldspar porphyroclasts were stable with garnet, kyanite,biotite, rutile, plagioclase <strong>and</strong> quartz, presumablywithin the S1 layering. The S2 fabric started to developÓ 2010 Blackwell Publishing Ltd351


114 J. FRANĚK ET AL.at these P–T conditions <strong>and</strong> re-equilibrated to theassemblage corresponding to the garnet rim composition,recrystallized K-feldspar, plagioclase, biotite,rutile, ilmenite <strong>and</strong> quartz, whereas kyanite separatedby plagioclase from the equilibrated matrix wasmetastable (see also Sˇtípska´ et al., 2010). The P–Testimate for the second stage is 0.9)1.0 GPa <strong>and</strong>825 °C, showing 0.8 GPa <strong>of</strong> decompression associatedwith cooling within the S2 fabric.Granular microstructuresProgressive recrystallization <strong>of</strong> the original coarsegrainedmicrostructure into the fine-grained granularmatrix is recorded in finger-like granular matrixdomains preserved in large perthite crystals shieldedfrom D2 deformation (Fig. 9a,b,e). In this context, theS2 matrix originated by growth <strong>and</strong> coalescence <strong>of</strong> thegranular domains preserved in the perthites (Fig. 9a,b)<strong>and</strong> the perthite grains, as well as the plagioclaseaggregates, represent precursors <strong>of</strong> the fine-grainedgranular matrix (Fig. 9c,d). A penetratively developedS2 mylonitic foliation is characterized by a granularmatrix composed <strong>of</strong> fine-grained K-feldspar, plagioclase<strong>and</strong> quartz, <strong>and</strong> elongated coarse-grained ÔplatyÕquartz ribbons (Figs 5e & 9e). The S2–S3 transitionis marked by a substantial change <strong>of</strong> the granulitemylonitic microstructure towards orthogneiss-likerocks mainly by phase redistribution <strong>and</strong> grain coarsening(Franeˇk et al., 2006).Type I: microstructure <strong>of</strong> granular domains within perthitesDevelopment <strong>of</strong> the granular microstructure startedwith the formation <strong>and</strong> growth <strong>of</strong> finger-like, finegraineddomains inside the perthite porphyroclasts(Type I microstructure, Fig. 9a,b). These domainscommonly show a lack <strong>of</strong> preferred orientation in allthe studied orthogonal sections; only in XZ sectionsare they locally elongated subparallel to the X direction<strong>of</strong> the D2 strain. The Type I microstructuredeveloped dominantly inside perthite porphyroclastsor along two adjacent perthite crystals <strong>and</strong> garnet–perthite grain boundaries (Fig. 6f), whereas alongquartz–perthite or kyanite–perthite boundaries itdevelops rarely.The K-feldspar grains in the Type I microstructureshow mainly oval shapes <strong>of</strong> small axial ratio comparedwith plagioclase. The plagioclase grains range fromcircular at K-feldspar triple junctions to highly elongatedwhere it coats boundaries between either newK-feldspar crystals or between relict perthite <strong>and</strong> newK-feldspar (Figs 9b,f & 10a). At the recrystallizationfront, the phase boundaries <strong>of</strong> K-feldspar <strong>and</strong> plagioclaseare curved <strong>and</strong> irregular, whereas in the recrystallizedgranular matrix the feldspar boundaries arestraighter (Fig. 10a,b).Areal microchemical <strong>analyses</strong> <strong>of</strong> the Type I microstructure(regions containing 100 grains) yield anaverage composition <strong>of</strong> 67.3% Or, 29.1% Ab <strong>and</strong>3.6% An, similar to that <strong>of</strong> neighbouring parentalperthite grains. On average, the K-feldspar compositionis 87.6% Or, 11.8% Ab <strong>and</strong> 0.6% An, <strong>and</strong> plagioclasecores consist <strong>of</strong> 1.5% Or, 78.1% Ab <strong>and</strong>20.4% An. Compositional maps reveal a sharp reactionfront separating parental perthite from the newlyformed Type I microstructure (Fig. 10a). The boundaryis easily distinguishable by crypto-perthitic exsolutions,which are abundant in the parental perthitebut absent from K-feldspar <strong>of</strong> the Type I microstructure.New K-feldspar grains reveal weak gradual zoningfrom Or 82 in cores to Or 93 at grain boundaries. Thematrix plagioclase grains show strong zoning markedby An 18–23 cores surrounded by An 11–12 rims. Locally,irregular patches <strong>of</strong> pure albite occur at the edges <strong>of</strong>plagioclase grains.At the edges <strong>of</strong> the perthite grains, the new matrix isweakly deformed, showing still the Type I microstructure.Nevertheless, increasing elongation <strong>of</strong> quartzinclusions with weakly recrystallized Type I microstructure(Fig. 9a) documents the influence <strong>of</strong> D2deformation. This Transitional type II microstructurediffers by the occurrence <strong>of</strong> quartz grains in the matrix,which suggests exchange <strong>of</strong> chemical components withthe perthite surroundings. The feldspar composition <strong>of</strong>this transitional type also shows significant differences(Fig. 7a). Area <strong>analyses</strong> yield an average feldsparcomposition <strong>of</strong> 46.5% Or, 42.0% Ab <strong>and</strong> 11.5% An,which is significantly different from both the parentalperthites <strong>and</strong> the shielded recrystallized domains. TheK-feldspar grains are 90.9% Or, 8.1% Ab <strong>and</strong> 1.0%An, <strong>and</strong> the plagioclase is on average 1.6% Or, 76.5%Ab <strong>and</strong> 21.9% An.The compositional map acquired at the outer edge<strong>of</strong> a decomposing perthite also reflects the chemicalchanges. Plagioclase zoning exhibits a distinct pattern,in which An 24 homogeneous grains contain onlyisolated remnants <strong>of</strong> An 11–13 outer domains. Weakgradual zoning <strong>of</strong> K-feldspar, in which the rims aredepleted in albite, resembles K-feldspar in the matrixshielded inside perthite.Plagioclase aggregates at this stage are completelyrecrystallized to an equi-dimensional mosaic markedby equigranular grains with straight boundaries,commonly meeting in triple point junctions (Fig. 9d).Minor quartz <strong>and</strong> K-feldspar locally occur at triplejunctions. Quartz forms weakly elongated aggregatessurrounded by the feldspar matrix.Type II: microstructure <strong>of</strong> penetrative fabric S2The Type II microstructure is linked to D2 deformation<strong>and</strong> it is the dominant micro<strong>structural</strong>type throughout the 8.5-km wide relict granulitefacies domain. The Type II microstructure is definedby a feldspar-dominated granular matrix enclosinglarge quartz ribbons. The feldspar aggregate is amixture <strong>of</strong> plagioclase <strong>and</strong> K-feldspar, forming anÓ 2010 Blackwell Publishing Ltd352


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 115(a)(b)(c)(d)(e)(f)Fig. 9. BSE images documenting recrystallization <strong>of</strong> large perthite <strong>and</strong> formation <strong>of</strong> the fine-grained granular matrix. (a <strong>and</strong> b)Heterogeneous recrystallization <strong>of</strong> the parental perthite at a sharp reaction front in domains partly parallel to trace <strong>of</strong> the S2 fabric.Note the elongation <strong>of</strong> a quartz inclusion released from the perthite documenting D2 strain. (c <strong>and</strong> d) Recrystallized Pl–Grt-richb<strong>and</strong> with interstitial quartz <strong>and</strong> cuspate geometry <strong>of</strong> several K-feldspar grains. (e) Recrystallization <strong>of</strong> a perthite grain partly exposedto the D2 strain (top <strong>of</strong> the porphyroclast) <strong>and</strong> partly in a pressure shadow (right side). (f) Detail <strong>of</strong> a shape-preferred orientation<strong>of</strong> plagioclase in a newly formed granular matrix adjacent to a parental perthite.Ó 2010 Blackwell Publishing Ltd353


116 J. FRANĚK ET AL.(a)(b)Fig. 10. Compositional maps <strong>of</strong> Ca <strong>and</strong> K from (a) matrix shielded inside parental perthite <strong>and</strong> (b) progressively deformed matrix inthe S2 fabric. For Ca, the K-feldspar <strong>and</strong> quartz grains are masked by white <strong>and</strong> black colour to emphasize the Ca zoning inplagioclase. Similarly, the plagioclase <strong>and</strong> quartz grains are masked in the K distribution maps to demonstrate the K zoning inK-feldspar. The BSE images are <strong>of</strong>fered for better orientation in phase distribution.Ó 2010 Blackwell Publishing Ltd354


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 117equi-dimensional mosaic (Fig. 5e). Transposition <strong>of</strong> S1compositional layering into the S2 fabric via rotation<strong>and</strong> attenuation suggests intense deformation duringD2 (Fig. 9a). The quartz b<strong>and</strong>s inherited from S1disintegrated to form the ribbons parallel to S2 <strong>and</strong>were further dynamically recrystallized along ribbonboundaries, leading to significant reduction <strong>of</strong> quartzgrain size (Fig. 11a). Numerous new small quartzgrains nucleated in triple junctions <strong>of</strong> feldspar grains inthe matrix. The garnet grains became dispersed in thematrix <strong>and</strong> kyanite crystals deformed by kinking.Isolated biotite flakes lie parallel to S2 <strong>and</strong>, togetherwith quartz ribbons, define a strong L > S fabric.Areal analysis <strong>of</strong> a Type II matrix region covering100 feldspar grains yields an average feldspar composition<strong>of</strong> 56.4% Or, 35.9% Ab <strong>and</strong> 7.7% An(Fig. 7a). The average composition <strong>of</strong> the K-feldspar is84.7% Or, 13.5% Ab <strong>and</strong> 1.8% An, <strong>and</strong> the plagioclaseis composed <strong>of</strong> 1.8% Or, 74.6% Ab <strong>and</strong> 23.6%An.Two compositional maps acquired from the feldspar-dominatedmatrix (one depicted in Fig. 10b)reveal typical mild zoning <strong>of</strong> K-feldspar, reflectingthe loss <strong>of</strong> Na around grain boundaries. Plagioclaseexhibits uniform An 23 composition with a weakincrease <strong>of</strong> 2–3% anorthite content towards rims.Locally, thin films <strong>of</strong> albite occur along plagioclase–K-feldspar <strong>and</strong> K-feldspar–K-feldspar boundaries;they are significantly thinner than rims in the Type Imicrostructure.Type III: microstructure <strong>of</strong> S3 amphibolite facies fabricThe S3 fabrics prevailing in the felsic granulites <strong>of</strong> theBLG reveal highly variable degrees <strong>of</strong> retrogression<strong>of</strong> the previous mineral assemblage. Contemporaneoushydration <strong>of</strong> granulites heterogeneously increasestowards the boundaries <strong>of</strong> the granulite massif, whereit is accompanied by widespread partial melting <strong>and</strong>segregation <strong>of</strong> the melt into mm–dm thick b<strong>and</strong>sparallel to S3 (Kodym, 1972; Franeˇk et al., 2006). Thequartz, K-feldspar <strong>and</strong> plagioclase show irregulargrain boundaries <strong>and</strong> develop into monomineralicaggregates (Fig. 12), whereas the quartz ribbons typicalfor the S2 entirely disappear. The aspect ratios <strong>of</strong>quartz grains significantly decrease, compared with theS2 fabric, in conjunction with coarsening <strong>of</strong> feldsparmosaic (Figs 5f & 11b). The abundant biotite flakeslie parallel to the foliation planes <strong>and</strong> sillimanite insome instances defines the lineation. The resulting S3microstructure resembles a common Bt + Sil ± Grtorthogneiss more than a retrograde granulite.In order to eliminate the effect <strong>of</strong> strain localizationinto melt b<strong>and</strong>s or biotite stripes that developed duringD3, only macroscopically homogeneous S3 sampleswith dispersed biotite were studied. The petrology <strong>and</strong>microchemistry <strong>of</strong> a steep fabric from a neighbouringKrˇisˇťanov Granulite Massif, analogous to the S3 in theBLG, are given in Verner et al. (2007).<strong>Quantitative</strong> micro<strong>structural</strong> analysisThe previous section defined three types <strong>of</strong> microstructuresdeveloped from the coarse-grained precursor<strong>of</strong> the felsic granulites: Type I corresponds togranular microstructure inside coarse perthite <strong>and</strong> atransitional type in the vicinity <strong>of</strong> coarse perthitegrains; Type II microstructure corresponds to thepervasively developed granulitic S2 fabric; <strong>and</strong> TypeIII microstructure corresponds to the S3 fabric developedunder amphibolite facies conditions. Accordingto the principle <strong>of</strong> fabric superposition (Wilson, 1961),the succession <strong>of</strong> micro<strong>structural</strong> types is seen as anevolutionary micro<strong>structural</strong> trend reflecting deformationstages along the exhumation path <strong>of</strong> thegranulites. The coarse-grained orthogneiss precursormicrostructure cannot be sufficiently describedusing quantitative methods. In order to comparethe three micro<strong>structural</strong> stages, we have quantifiedseveral parameters characterizing the microstructuresusing the PolyLX toolbox (Lexa et al., 2005) forthe MATLAB TMs<strong>of</strong>tware package <strong>and</strong> CSD-correctionprogram (e.g. Higgins, 1998). Four thin sections orientedparallel to the lineation <strong>and</strong> perpendicular to thefoliation (XZ sections) from representative samples <strong>of</strong>the S2 (956 & 1252 grains) <strong>and</strong> S3 fabrics (848 & 886grains) were digitalized in an ArcView GIS environment(Fig. 11a,b) <strong>and</strong> analysed by the PolyLX <strong>and</strong>CSD-correction s<strong>of</strong>tware. In order to obtain statisticallysignificant results, we have merged four areas <strong>of</strong>the Type I microstructure yielding 801 grains. Beforemerging, the individual areas were rotated with respectto the trace <strong>of</strong> S2. In addition, the transitional matrixarea at the boundary between the perthite <strong>and</strong> D2matrix was digitalized, covering 868 grains.Different micro<strong>structural</strong> types have been quantifiedin plots <strong>of</strong> bulk grain-boundary preferred orientation(GBPO) against contact frequencies (Fig. 12) <strong>and</strong>slope <strong>of</strong> linear regression from crystal size distribution(CSD) analysis v. regression intercept (Fig. 13). Thegrain size as well as other quantitative characteristicsare statistically evaluated in Table 2, more detaileddescription <strong>of</strong> the quantitative data is given inAppendix S1.The contact frequency method (Kretz, 1969)compares the observed (O) count <strong>of</strong> contacts betweentwo minerals with the value expected (E) for a perfectlyr<strong>and</strong>om distribution. The calculation involves modalproportions <strong>of</strong> phases, only a scatter in grain size maycause a limited uncertainty <strong>of</strong> the results. The boundariesare designated as Ôlike–likeÕ for contacts betweengrains <strong>of</strong> one phase or ÔunlikeÕ for the case <strong>of</strong> boundarybetween two different phases. The resulting v 2 -value(O ) E) ⁄ sqrt(E) is a measure <strong>of</strong> the deviation <strong>of</strong>mineral spatial distribution from r<strong>and</strong>om. It is positivefor like–like contacts <strong>and</strong> negative for unlike contactsif the corresponding microstructure tends to formmonomineralic aggregates (Lexa et al., 2005). On theother h<strong>and</strong>, the unlike contacts exhibit positive values,Ó 2010 Blackwell Publishing Ltd355


118 J. FRANĚK ET AL.(a)(b)Fig. 11. Micro<strong>structural</strong> character <strong>of</strong> XZ sections (X is parallel to lineation, Z normal to foliation) <strong>of</strong> the two fabrics quantified inthe felsic granulites: (a) the granulite facies S2 <strong>and</strong> (b) the amphibolite facies S3. The left column combines optical image undercrossed polarizers with its vectorized form used for statistical analysis in Fig. 12. The right column presents BSE images pointing outphase distribution.Ó 2010 Blackwell Publishing Ltd356


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 119Fig. 12. <strong>Quantitative</strong> micro<strong>structural</strong>analysis <strong>of</strong> the degree <strong>of</strong> grain boundarypreferredorientation v. contact frequenciesfor like <strong>and</strong> unlike boundaries <strong>of</strong> feldspars.Numbers at individual data points refer tothe evolution stages defined in the sectionGranular microstructures.Fig. 13. Crystal size distribution curves for both feldspars. Values <strong>of</strong> slope <strong>and</strong> upper intercept <strong>of</strong> linear regression are plotted togetherin the inset showing an evolutionary trend for both the feldspars. Numbers at individual data points refer to the evolution stagesdefined in the section Granular microstructures.whereas the like–like contacts are negative where thephases tend to be regularly distributed.Crystal size distributions measured in metamorphicrocks yield quantitative information about crystalnucleation <strong>and</strong> growth rates, growth times <strong>and</strong> thedegree <strong>of</strong> overstepping <strong>of</strong> reactions during metamorphism(Cashman & Ferry, 1988). The CSDs aredescribed as the population density function <strong>of</strong> thecumulative number <strong>of</strong> crystals per unit volume perlinear crystal size. The data are plotted in log-normalspace <strong>and</strong> a linear regression <strong>of</strong> points is the characteristic.The intercept values (N0) are proportional tothe nucleation density <strong>of</strong> the examined phase at theonset <strong>of</strong> nucleation (size 0), <strong>and</strong> the slope (Gt) is proportionalto the modification <strong>of</strong> original grain-sizedistribution by a variety <strong>of</strong> coarsening processes suchÓ 2010 Blackwell Publishing Ltd357


120 J. FRANĚK ET AL.Table 2. <strong>Quantitative</strong> characteristics <strong>of</strong> the micro<strong>structural</strong>evolution.EAD (lm) Bulk SPO AxialratioQ1 Median Q3as Ostwald ripening or Ôcommunicating neighboursÕtheory (Marsh, 1988). Although the individual absolutevalues are meaningless without knowledge <strong>of</strong>kinetic parameters, they can be used to compare individualstages <strong>of</strong> the textural maturation history in asingle rock type (Lexa et al., 2005; Hasalova´ et al.,2008b; Schulmann et al., 2008). The SPO <strong>and</strong> GBPOare calculated as ratios <strong>of</strong> eigenvalues <strong>of</strong> the orientationtensors <strong>of</strong> long axes <strong>of</strong> minerals <strong>and</strong> grainboundaries, respectively (Lexa et al., 2005), <strong>and</strong> theirvalues usually depend on types <strong>of</strong> active recrystallization<strong>and</strong> ⁄ or deformation mechanisms. The aspectratio, SPO <strong>and</strong> GBPO are relatively high for dynamicallyrecrystallized grains deformed by dislocationcreep (Kruse et al., 2001; Ulrich et al., 2002; Baratouxet al., 2005), whereas low values commonly characterizeGBS-controlled diffusion creep associated withmutual rotation <strong>of</strong> individual grains (Boullier &Gueguen, 1975; Behrmann & Mainprice, 1987).Grain contact frequency evolutionCSDinterceptCSDslope1-shielded granular Kfs 33.7 56.3 84.4 1.21 1.53 11.57 )16.2matrixPl 26.8 42.3 59.4 1.22 1.51 12.78 )22.62T-weakly deformed Qtz 18.0 26.6 40.9 2.90 1.67matrixKfs 44.8 67.1 90.8 1.40 1.49 11.57 )15.7Pl 29.2 40.3 59.0 1.20 1.53 13.54 )26.02-S2 granulitic matrix Qtz 14.7 22.5 45.2 2.12 1.49Kfs 56.5 83.9 126.7 1.30 1.43 10.40 )11.6Pl 46.3 64.7 90.6 1.14 1.48 11.73 )16.73-S3 retrograde fabric Qtz 39.5 59.8 104.2 1.50 1.46Kfs 42.7 67.4 109.2 1.40 1.54 10.07 )11.4Pl 40.9 67.2 111.5 1.27 1.43 10.50 )12.5Equal area diameter (EAD) statistical characteristics depicting K-feldspar, plagioclase <strong>and</strong>quartz grain-size evolution from the perthite recrystallization process to amphibolite faciessyntectonic retrogression. Q1 <strong>and</strong> Q3 refer to 1st <strong>and</strong> 3rd quartiles. The bulk SPO iscalculated as the ratio <strong>of</strong> eigenvalues <strong>of</strong> bulk orientation tensor according to Lexa et al.(2005), the axial ratio values represent modus <strong>of</strong> their lognormal distribution.The grain contact frequency method yields a welldefinedevolutionary trend <strong>of</strong> like–like contacts forboth Kfs <strong>and</strong> Pl. The two minerals reveal fairly similar<strong>and</strong> highly negative values <strong>of</strong> (O ) E) ⁄ sqrt(E) ratio()6) for Type I microstructure (Fig. 12), suggestinghighly regular spatial distribution <strong>of</strong> both feldspars.Both feldspars also reveal distinct GBPO (1.07–1.15).The Type II microstructure shows an important increase<strong>of</strong> like–like grain contact frequency in Fig. 12,reaching zero, which suggests an almost perfectlyr<strong>and</strong>om distribution <strong>of</strong> both feldspars, coupled withalmost complete loss <strong>of</strong> GBPO. Finally, the Type IIImicrostructure reveals a further increase <strong>of</strong>(O ) E) ⁄ sqrt(E) values, suggesting a progressivelydeveloping aggregate distribution (higher for plagioclasecompared to K-feldspar) associated with anincrease <strong>of</strong> GBPO values. The evolution <strong>of</strong> K-feldspar–plagioclase grain boundary frequencies mirrors thetrend <strong>of</strong> like–like contacts, but the GBPO remainsmore developed for all stages, compared with like–likeboundaries, with peaks in Types I <strong>and</strong> III microstructures.The progressive evolution <strong>of</strong> rock structurefrom a regular distribution <strong>of</strong> grains through r<strong>and</strong>omto an aggregate distribution indicates a process <strong>of</strong>solid-state differentiation <strong>and</strong> development <strong>of</strong> minerallayering. The fluctuation <strong>of</strong> GBPO suggests destruction<strong>of</strong> the original preferred orientation <strong>of</strong> grainboundaries by D2, whereas D3 was responsible for thedevelopment <strong>of</strong> a new GBPO.Grain-size evolutionThe grain-size statistics (Table 2, Appendix S1) has arather constant <strong>and</strong> low value (55–70 lm) for recrystallizedK-feldspar <strong>of</strong> the Type I microstructure <strong>and</strong>transitional microstructures adjacent to perthite crystals,<strong>and</strong> a slightly larger grain size for the Types II <strong>and</strong>III microstructures (84 & 67 lm, respectively). Similarly,fine-grained plagioclase has a small <strong>and</strong> constantgrain size for the Type I microstructure (42 lm) <strong>and</strong> alarger grain size for Types II <strong>and</strong> III (66 lm). Thequartz grain size first decreases from the transitionalmicrostructure (26 lm) to the Type II microstructure(23 lm), <strong>and</strong> increases in the Type III microstructure(60 lm). In general, the K-feldspar recrystallized grainsize is the largest <strong>of</strong> all the studied minerals (70 lm).The CSD <strong>analyses</strong> reveal a two stage grain-sizeevolution for both plagioclase <strong>and</strong> K-feldspar(Fig. 13). Both minerals show decrease <strong>of</strong> N0 values<strong>and</strong> increase <strong>of</strong> Gt values from the Type I to Types II<strong>and</strong> III microstructures. For plagioclase the evolutionis marked by gradual decrease in the N0 value, coupledwith an increase in the Gt value from the Types II toIII microstructures. Importantly, although the evolution<strong>of</strong> the plagioclase <strong>and</strong> K-feldspar grain size startsfrom very different positions in the Gt–N0 plot (higherN0 <strong>and</strong> smaller Gt in plagioclase, compared withK-feldspar), they end up with nearly the same values <strong>of</strong>both parameters for the Type III microstructure. Bothminerals reach the greatest difference in grain-sizeparameters for the Type II microstructure, for whichplagioclase shows significantly higher N0 values <strong>and</strong>smaller Gt values compared with K-feldspar.Shape-preferred orientation <strong>and</strong> grain elongationThe aspect ratio remains constant for both K-feldspar<strong>and</strong> plagioclase during the whole deformationsequence <strong>and</strong> the bulk shape-preferred orientation(SPO) <strong>of</strong> plagioclase remains similarly low, whereasK-feldspar SPO is slightly higher for Types II <strong>and</strong>III microstructures (Table 2). Quartz reaches thehighest degree <strong>of</strong> SPO (mean value 2–3) for Type II,together with a relatively high aspect ratio. The TypeIII microstructure is characterized by a decrease <strong>of</strong> thequartz SPO in conjunction with a significant decreasein the aspect ratio (Table 2).Ó 2010 Blackwell Publishing Ltd358


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 121Lattice-preferred orientationLattice-preferred orientation (LPO) measurements <strong>of</strong>quartz, K-feldspar <strong>and</strong> plagioclase grains from all thedescribed micro<strong>structural</strong> types were additionallycarried out in order to evaluate operative deformationmechanisms. A strong LPO usually originates duringplastic deformation via dislocation creep, whereas thediffusion-accommodated GBS generally weakensthe LPO <strong>of</strong> deforming grains (e.g. Jiang et al., 2000).The slip systems accommodating dislocation creep arewell known for quartz (e.g. Schmid & Casey, 1986) <strong>and</strong>plagioclase (e.g. Tullis, 1983; Montardi & Mainprice,1987; Kruse et al., 2001; Stu¨ nitz et al., 2003), but lessknown for K-feldspar (Tullis, 1983). To ensure highquality <strong>of</strong> measurements, the crystal lattice orientationswere collected manually via EBSD in theLAREM laboratory <strong>of</strong> the Czech Geological Survey<strong>and</strong> at the Institute <strong>of</strong> Petrology <strong>and</strong> StructuralGeology at Charles University in Prague. LPO data foreach sample (XZ thin section) were plotted separatelyas non-polar projections on a lower hemisphere <strong>and</strong>important slip planes <strong>and</strong> directions for quartz,K-feldspar <strong>and</strong> plagioclase were projected.Type I microstructureAt first, the EBSD study was carried out in domainsconsisting <strong>of</strong> perthite <strong>and</strong> the newly formed Type Imatrix enclosed within the perthite, therefore protectedfrom D2 <strong>and</strong> D3 deformation. The Type I K-feldspar<strong>and</strong> plagioclase grains reveal roughly the same latticeorientation as the K-feldspar host <strong>and</strong> plagioclaseexsolutions in parental perthite. When passing outsidethe protected matrix into the Transitional type IImicrostructure, a continuous increase <strong>of</strong> LPO scatteringwith respect to the parental perthite orientationoccurs in the matrix without any tendency to develop anew LPO (Fig. 14).Type II microstructureThe pole figures <strong>of</strong> matrix quartz show incomplete highangleType II cross-girdle (Lister & Price, 1978) <strong>of</strong> c-axes(Fig. 15a). High opening angle <strong>and</strong> central maximumare consistent with dominance <strong>of</strong> prism <strong>and</strong> slip systems, being variably accompanied by rhomb + slip (e.g. Lister & Dornsiepen, 1982).The large quartz ribbons have an irregular size <strong>and</strong>distribution <strong>of</strong> subgrains, with inclined subgrainboundaries <strong>and</strong> a strong central maximum sometimesaccompanied with minor submaxima oriented closeto the lineation. This typical granulite pattern (Behr,1961) results from predominance <strong>of</strong> the prism slip, only rarely accompanied by subordinate prism slip (Schmid & Casey, 1986).In plagioclase, the LPO shows an incomplete girdle<strong>of</strong> (010) poles normal to the lineation direction <strong>and</strong>maxima <strong>of</strong> [201], [101] <strong>and</strong> [001] subparallel to thelineation. It indicates the activity <strong>of</strong> primary [001] (010)Fig. 14. Electron back-scattered diffraction measurements <strong>of</strong> a perthite domain undergoing recrystallization to granular matrix <strong>and</strong>initial D2 deformation. The crystal lattice orientation <strong>of</strong> both K-feldspar background <strong>and</strong> plagioclase exsolutions in parental perthiteare compared to both feldspars in the granulitic matrix located inside the perthite, at the edge <strong>of</strong> the perthite <strong>and</strong> in the deformedmatrix adjacent to the perthite grain. The elongation <strong>of</strong> quartz grains reflects D2 strain intensity <strong>and</strong> depicts orientation <strong>of</strong> S2 fabric.Lower hemisphere non-polar equal-area projections.Ó 2010 Blackwell Publishing Ltd359


122 J. FRANĚK ET AL.(a)(b)Fig. 15. Polar diagrams <strong>of</strong> lattice preferred orientation <strong>of</strong> quartz, K-feldspar <strong>and</strong> plagioclase from XZ sections <strong>of</strong> felsic granulitesexhibiting penetrative (a) S2 <strong>and</strong> (b) S3 fabrics. Only the slip planes significant for each mineral are presented; lower hemispherenon-polar equal-area projections.as well as secondary [101] (010) <strong>and</strong> [201] (010) slipsystems according to Kruse et al. (2001). The LPO <strong>of</strong>K-feldspar is similar to that <strong>of</strong> plagioclase suggestingactivity <strong>of</strong> the same slip systems. Additional slip system[100] (010) in K-feldspar makes the only differencebetween both <strong>of</strong> the feldspars. It is noteworthy thatpole figures <strong>of</strong> both feldspars show double maxima <strong>of</strong>single crystal orientations close to each other.Ó 2010 Blackwell Publishing Ltd360


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 123Type III microstructureThe quartz LPO shows a high angle Type II crossgirdle<strong>of</strong> c-axes distribution with a maximum either inthe centre or near edges <strong>of</strong> the girdle. The LPO issimilar to those measured from the S2 matrix microstructure<strong>and</strong> implies a combination <strong>of</strong> prism <strong>and</strong> rhomb + slip (Fig. 15b). Plagioclaseshows the (010) <strong>and</strong> (001) planes subparallel to foliation<strong>and</strong> a girdle distribution <strong>of</strong> [100] <strong>and</strong> [201] alongthe foliation with single maxima parallel to the lineation.This pattern points to activity <strong>of</strong> three slip systems,namely [100] (010), [201] (010) <strong>and</strong> [100] (001). Inthe case <strong>of</strong> K-feldspar, all samples show a pronouncedmaximum <strong>of</strong> (010) planes parallel to foliation, whereasother common slip planes do not show such relationship.A girdle distribution is observed for the [001] <strong>and</strong>[100] directions, with a maximum <strong>of</strong> the [001] directionoriented parallel to the stretching lineation. Theseattributes indicate that the activity <strong>of</strong> [100](010) slipsystem is the characteristic for both feldspars.DISCUSSIONWe discuss the origin <strong>and</strong> petrological significance <strong>of</strong>the S1 fabric as a precursor for development <strong>of</strong> characteristicgranular microstructure <strong>of</strong> Variscan felsicgranulites. Subsequently, an attempt is made to discussthe origin <strong>of</strong> granular fabric, its specific quantitativemicro<strong>structural</strong> characteristics <strong>and</strong> its importance forunderst<strong>and</strong>ing deep crustal flow processes. Finally, thepetrology, microstructure <strong>and</strong> LPO <strong>of</strong> various types <strong>of</strong>granulites are discussed in terms <strong>of</strong> the rheologicalevolution <strong>of</strong> orogenic lower crust <strong>and</strong> its extrusion(Franěk et al., 2011) during the Variscan orogeny inEurope. Figure 16 puts the micro<strong>structural</strong> evolutionin context with the tectonic history.Interpretation <strong>of</strong> S1 fabric: HP orthogneissThe observed <strong>structural</strong> succession, field distribution<strong>of</strong> the S1 relicts <strong>and</strong> micro<strong>structural</strong> relations implythat the perthitic alkali feldspar, plagioclase <strong>and</strong> quartzaggregates preserved locally within the S1 compositionallayering represent a remnant <strong>of</strong> the precursor <strong>of</strong>the Blansky´ les felsic granulites. The best preservedrelicts <strong>of</strong> the S1 layering contain a substantial amount<strong>of</strong> large alkali feldspar (presently perthite) <strong>and</strong> largequartz grains in the microstructure. The inclusions <strong>of</strong>kyanite, high-grossular garnet <strong>and</strong> rutile suggest thatthe large alkali feldspar crystallized at HP conditions(Figs 8 & 16). Despite diffusional re-equilibration <strong>of</strong>most minerals, the peak P–T conditions related to thegrowth <strong>of</strong> large alkali feldspar are estimated at1.6)1.8 GPa <strong>and</strong> 850)880 °C. The abundant idiomorphicor oval-shaped quartz inclusions in the largeperthite porphyroclasts point to a lack <strong>of</strong> plasticdeformation during alkali feldspar growth. Suchinclusions may develop either in granites, or duringmelt-assisted recrystallization in migmatites (Mehnert,1968, pp. 111, 192). The lenticular plagioclase-richaggregates are interpreted as completely recrystallizedremnants <strong>of</strong> large plagioclase grains complementary tothe alkali feldspar porphyroclasts (Figs 9c,d & 16, 3Dblock diagram 1). Consequently, the original rock was acoarse-grained granite ⁄ orthogneiss crystallized underHP conditions in the kyanite stability field. Existence <strong>of</strong>two complementary feldspars at peak conditions contrastswith laboratory experiments <strong>of</strong> Tropper et al.(2005) focused on the South-Bohemian felsic granulites,where only a single alkali feldspar existed above 850 °C.The S1 layering <strong>of</strong> initially coarse-grained quartz,K-feldspar <strong>and</strong> plagioclase-rich layers may be interpretedas a result <strong>of</strong> solid-state deformational segregation<strong>of</strong> minerals with different plasticity at hightemperatures, a process common in formation <strong>of</strong>, e.g.layered orthogneisses. The length <strong>of</strong> the b<strong>and</strong>s thenindicates significant strain during the D1 phase.Assuming growth <strong>of</strong> large alkali feldspar during theD1 episode, the overgrowth <strong>of</strong> quartz, garnet <strong>and</strong>biotite by the large feldspar indicates a dominance <strong>of</strong>growth processes resulting in overall coarsening <strong>of</strong> thegneiss (Higgins, 1998; Lexa et al., 2005). Such crystalgrowth indicates that the temperature ⁄ strain rate ratiowas rather high <strong>and</strong> thermodynamic <strong>modelling</strong> (Fig. 8)suggests contemporaneous partial melting. High-Ppeak conditions (1.6–1.8 GPa) indicate that the flowoccurred at the bottom <strong>of</strong> thickened crust as alsosuggested by other studies (e.g. Sˇtípska´ & Powell, 2005;Racek et al., 2006; Tajcˇmanova´ et al., 2006). Therefore,the S1 fabric probably originated during lowercrustal flow at the bottom <strong>of</strong> a thickened crustal root<strong>and</strong> at progressively increasing temperatures (Fig. 16).The inclusions within the perthite porphyroclasts,the S1 compositional layering <strong>and</strong> the S2 myloniticfoliation in the BLG are similar to the oldest structuresdescribed from the Saxonian Granulitgebirge (Behr,1961). In addition, all <strong>of</strong> the Variscan felsic granulitesreveal similar geochemistry <strong>and</strong> geochronology(340 Ma peak). Consequently, both the mentionedmassifs <strong>and</strong> other Variscan felsic granulite massifs aswell may have evolved from such HP coarse-grainedorthogneisses.Origin <strong>of</strong> granular matrix in granulitesThe microchemical <strong>and</strong> textural relations indicate thatthe initial large alkali feldspar underwent heterogeneousstep-by-step recrystallization during the D2event, which resulted in formation <strong>of</strong> a fine-grainedmatrix. This recrystallization probably postdatedexsolution <strong>of</strong> the coarse perthitic plagioclase because(i) the newly formed Type I K-feldspar in the granuliticmatrix is devoid <strong>of</strong> such large braided exsolutions, <strong>and</strong>(ii) the perthite exsolutions reveal 1–2% higheranorthite content than the cores <strong>of</strong> Type I matrixplagioclase indicating a bit higher temperature <strong>of</strong> formation(Fig. 16, 3D block diagram 2). Nevertheless, itÓ 2010 Blackwell Publishing Ltd361


124 J. FRANĚK ET AL.Fig. 16. Interpretative 3D scheme <strong>of</strong> the micro<strong>structural</strong> evolution <strong>of</strong> felsic granulites from the coarse-grained layered HP granuliteprecursor to granulite facies mylonite developed during exhumation in a crustal scale ÔdiapiricÕ dome according to Franeˇk et al. (2011).The block diagram 1 shows the tentative model <strong>of</strong> original layered orthogneiss microstructure with numerous kyanite, quartz <strong>and</strong>garnet crystals in large feldspar porphyroblasts. The model <strong>of</strong> ÔdiapiricÕ exhumation (Franeˇk et al., 2011; Lexa et al., 2011) shows theposition <strong>of</strong> this microstructure at the bottom <strong>of</strong> the crust <strong>and</strong> its P–T conditions (1). The extent <strong>of</strong> preservation <strong>of</strong> the S1 fabric ismarked by light shading at the bottom <strong>of</strong> diapiric extrusion. The block diagram (2) shows the development <strong>of</strong> perthites at the beginning<strong>of</strong> exhumation with decompression <strong>and</strong> limited cooling. The block diagram (3) exhibits the development <strong>of</strong> typical granulite fabric,such as platy quartz <strong>and</strong> granular microstructures <strong>of</strong> Types I <strong>and</strong> II. Relicts <strong>of</strong> kyanite preserved in perthite porphyroclasts <strong>and</strong> matrixwith decompression kelyphitic haloes are shown. This microstructure is developed in the main vertical channel <strong>of</strong> the ÔdiapirÕ <strong>and</strong> ismarked by intermediate shading in the <strong>structural</strong> model. The ÔdiapirÕ head shows the stability <strong>of</strong> Type III granular microstructure (darkshading) <strong>and</strong> rotated relict domains preserving Type II microstructure.Ó 2010 Blackwell Publishing Ltd362


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 125cannot be fully excluded that the large perthitic exsolutionsin alkali feldspar porphyroclasts originatedduring later stages <strong>of</strong> cooling <strong>and</strong> decompression alongthe exhumation path.The recrystallization must have initiated by deformation-inducedconversion <strong>of</strong> the (perthitic) parentalalkali feldspar directly into small K-feldspar grains byan uncommon process, which is discussed in the nextparagraphs. Plagioclase was then gradually redistributedalong the new K-feldspar grain boundaries (e.g.Fig. 9b,f), presumably via grain-boundary diffusion.The sharp compositional boundaries between An 23cores <strong>and</strong> An 12 rims <strong>of</strong> the Type I plagioclase(Fig. 10a) indicate that they did not develop by continuousgrowth <strong>and</strong> that they were not significantlyaffected by later diffusion. We suggest that at first, theAn 23 plagioclase grains coalesced in triple junctions <strong>of</strong>the new K-feldspar grains by grain-boundary diffusionmechanism. They originated by reduction <strong>of</strong> surface<strong>and</strong> coalescence <strong>of</strong> preexisting coarse <strong>and</strong> compositionallysimilar An 24 plagioclase perthitic exsolutions,assuming perthitic nature <strong>of</strong> the parental alkali feldspar.The new grain boundaries were subsequentlyovergrown by An 12 plagioclase, the components forwhich were probably released from surroundingK-feldspar by volume diffusion. This process is indicatedby gradual zoning at rims <strong>of</strong> adjacent K-feldspargrains (Fig. 10a). The CSD patterns (Fig. 13) showthat the formation <strong>of</strong> the Type I K-feldspar–plagioclasemosaic originated by a nucleation-dominatedprocess accompanied by limited crystal growth, comparedwith the later Types II <strong>and</strong> III microstructure(Lexa et al., 2005; Hasalova´ et al., 2008b). The nucleationdensity <strong>of</strong> plagioclase was significantly higherthan that <strong>of</strong> K-feldspar.The most important manifestation <strong>of</strong> the perthiterecrystallization is redistribution <strong>of</strong> Or, Ab <strong>and</strong> Ancomponents on a micro-scale in a chemically closedsystem, resulting in development <strong>of</strong> the fine-grainedgranular matrix (Fig. 16, 3D block diagram 3). Thehigh amount <strong>of</strong> unlike grain boundaries in the granularmatrix preserved inside parental perthite indicates atendency to regular distribution <strong>and</strong> thus active intermixing<strong>of</strong> the K-feldspar <strong>and</strong> plagioclase. The regulardistribution represents the energetically lowest state <strong>of</strong>such a polyphase material (Seng, 1936; DeVore, 1959)<strong>and</strong> progressive straightening <strong>of</strong> grain boundarygeometries <strong>of</strong> K-feldspar <strong>and</strong> plagioclase further minimizesthe surface energy by grain shape simplificationtypical for grain boundary reduction (Passchier et al.,1992). According to this model the recrystallizationthen appears to be substantially driven by a decrease <strong>of</strong>chemical <strong>and</strong> surface energy in the metastable (perthitic)alkali feldspar. The concept <strong>of</strong> surface energyminimization was suggested by Flin (1969), who proposedthat the regular distribution is a consequence <strong>of</strong>a smaller interfacial energy <strong>of</strong> unlike boundaries incomparison with like–like boundaries. However,Ramberg (1952) considered the interfacial energies tobe too small to drive diffusional mass transfer ingranulites.The weak but systematic scatter <strong>of</strong> lattice orientations<strong>of</strong> new plagioclase <strong>and</strong> K-feldspar grains withrespect to the parental perthite (Fig. 14) points tocontribution <strong>of</strong> stress to the bulk recrystallizationprocess. Such a process <strong>of</strong> non-coherent decomposition<strong>of</strong> metastable alkali feldspar may be triggeredeither by water addition or by the application <strong>of</strong>external stress (Brown & Parsons, 1989 <strong>and</strong> referencestherein). As a lack <strong>of</strong> hydrous phases <strong>and</strong> presumablylack <strong>of</strong> water is typical for the granulite facies evolution<strong>of</strong> these rocks, the activity <strong>of</strong> external stressseems to be better c<strong>and</strong>idate to trigger the recrystallizationprocess.The almost identical LPO <strong>of</strong> plagioclase <strong>and</strong>K-feldspar, regular grain distribution <strong>and</strong> high nucleationdensity CSD pattern <strong>of</strong> the polymineralicaggregate developed by recrystallization <strong>of</strong> solidsolution crystal has not been previously described innature. The sharp recrystallization front advancinginto parental perthites represents a moving grainboundary driven by both chemical <strong>and</strong> deformationalprocesses. According to Stu¨ nitz (1998), such recrystallizationcan be considered as a combination <strong>of</strong>strain- <strong>and</strong> chemically induced grain boundarymigration processes. At the temperature <strong>and</strong> chemicalconditions given during the D2, the activation energy<strong>of</strong> the grain boundary migration process was probablylower than that <strong>of</strong> dynamic recrystallization <strong>of</strong> themetastable (perthitic) alkali feldspar. Material scienceliterature (e.g. Saheb et al., 1995; Sennour et al., 2004)has examples <strong>of</strong> sharp recrystallization fronts advancinggradually into intact grains <strong>of</strong> alloys (Fig. 9a,b,f)explained by the chemically induced grain boundarymigration or a similar discontinuous precipitationprocess, which in many aspects resembles the alkalifeldspar recrystallization. According to Yoon (1995),the discontinuous precipitation in metastable alloysrepresents largely autocatalytic recrystallization drivenmainly by coherency strain energy. Activation <strong>of</strong> thisprocess depends upon composition, temperature, stress<strong>and</strong> strain, elastic anisotropy, crystallographic relations<strong>and</strong> boundary curvature <strong>of</strong> involved grains(Yoon, 1995). Once triggered, e.g. by external strain, itcauses very efficient heterogeneous decomposition <strong>of</strong>oversaturated solid solutions (like alkali feldspar) via(mainly chemically driven) grain boundary migration(Hay & Evans, 1987; Saheb et al., 1995; Sennour et al.,2004).In conclusion, the formation <strong>of</strong> the granular matrixresembles the discontinuous precipitation triggered bystraining <strong>of</strong> the probably perthitic alkali feldspargrains during D2 deformation (Fig. 16). The processstarts preferably on pre-existing grain boundaries withother feldspar porphyroclasts or with garnet. Themigration <strong>of</strong> the recrystallization front is driven mainlyby the metastability <strong>of</strong> the (perthitic) alkali feldspar. Inturn, the metastability <strong>of</strong> parental feldspar controls theÓ 2010 Blackwell Publishing Ltd363


126 J. FRANĚK ET AL.development <strong>of</strong> the irregular to amoeboid domains(e.g. Fig. 9a,b) <strong>of</strong> fine-grained regularly distributedK-feldspar <strong>and</strong> plagioclase matrix.Lower crustal ascent <strong>and</strong> strength evolution <strong>of</strong> felsicgranulitesThe onset <strong>of</strong> the D2 deformation is marked by highlynon-cylindrical folding <strong>of</strong> the S1 anisotropy prior tothe S2 cleavage development (Figs 4a,b & 16, blockdiagram 3). The formation <strong>of</strong> a typical Variscan felsicgranulite fine-grained microstructure with highlyelongated quartz ribbons is the main result <strong>of</strong> the D2deformation (Figs 5e & 9e) that is contemporaneouswith the development <strong>of</strong> the granular Type I microstructure.Petrological <strong>modelling</strong> <strong>and</strong> mineral zoning,particularly the decrease <strong>of</strong> the grossular componentin garnet, suggest that between the perthite growth<strong>and</strong> the end <strong>of</strong> the D2 there was significant decompression<strong>of</strong> 0.8 GPa at slightly decreasing temperatureover a time span <strong>of</strong> several million years (Figs 8& 16). That means the S2 microstructure developedat slightly lower temperature compared with the S1<strong>and</strong> under continuously decreasing pressure as indicatedby a number <strong>of</strong> decompression microstructuressuch as plagioclase rims around kyanite <strong>and</strong> garnet(Figs 5e & 6h). This conclusion agrees with theobservations <strong>of</strong> spinel–plagioclase <strong>and</strong> garnet coronasdeveloped around kyanite, which form characteristicdecompression microstructures in S2 (Sˇtípska´ et al.,2010). The common denominator <strong>of</strong> these coronas isthat they are elongated parallel to S2 foliation(Fig. 6h) <strong>and</strong> therefore developed at the end <strong>of</strong> theD2 deformation.The small interstitial quartz grains, discontinuousalbite films at plagioclase or K-feldspar boundaries<strong>and</strong> cuspate feldspar grain shapes (Fig. 9d) are consistentwith the presence <strong>of</strong> syn-deformational intergranularpartial melt in the granulitic S2 matrix, aconclusion previously suggested by Franeˇk et al.(2006) <strong>and</strong> supported also by dykes <strong>of</strong> felsic granulitecrosscutting mafic boudins embedded in S2. Tajčmanova´et al. (2006, 2007) inferred a low content <strong>of</strong>partial melt (


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 127D3 switch in rheology <strong>of</strong> feldspar v. quartzThe micro<strong>structural</strong> characteristics <strong>of</strong> the S3 micr<strong>of</strong>abric(Type III microstructure) differ significantlycompared with those <strong>of</strong> the S2. The high number <strong>of</strong>like–like boundaries in the feldspar-dominated matrix(Fig. 12) indicates significant coalescence <strong>of</strong> individualphases into monomineralic aggregates, <strong>and</strong> the SPO,GBPO <strong>and</strong> axial ratio <strong>of</strong> both feldspars also increase.In contrast, quartz grains exhibit a decrease in SPO aswell as in axial ratio. The Type III microstructure isclearly characterized by a growth-dominated process,especially in the case <strong>of</strong> K-feldspar. The N0–Gt values<strong>of</strong> Types I, II <strong>and</strong> III microstructures project on acurve indicating simultaneous change <strong>of</strong> N0 <strong>and</strong> Gtdue to temperature <strong>and</strong> strain variations <strong>and</strong> suggestthat strain rate change probably plays a key role in theresulting CSD shape. Such an evolutionary trend is<strong>of</strong>ten interpreted as a result <strong>of</strong> grain coarsening, whichunusually occured at lower temperature related toupper amphibolite facies metamorphism during D3(700–800 °C) compared with the D2 granulite faciesconditions (800–850 °C) (Figs 3 & 8). Therefore, thestarting D2 recrystallized grain size was not in equilibriumwith the temperature-corrected strain rate(Zener & Holomon, 1944, parameter Z), such that themicrostructure was located in the grain-coarseningfield (fig. 1 <strong>of</strong> Ulrich et al., 2006 modified after Sakai &Jonas, 1984). In conclusion, only a significant strainrate decrease can explain the observed S3 graincoarsening at decreased temperature.The S3 microstructure tendency for monomineralicaggregate distribution, high SPO <strong>and</strong> the generalprevalence <strong>of</strong> grain growth over nucleation comparedwith S2 are in agreement with the strong LPO, suggestingthat the quartz, K-feldspar <strong>and</strong> plagioclasedeformed predominantly by dislocation creep. Theirregular grain boundaries in the S3 microstructure arealso typical <strong>of</strong> dislocation creep rather than GBS ordiffusional mechanisms. The dominance <strong>of</strong> dislocationcreep indicates hardening <strong>of</strong> the felsic granulites, wherethe less plastic feldspar forms a low viscosity contrastload-bearing framework (LBF, H<strong>and</strong>y, 1990) that resultedin high bulk strength <strong>of</strong> the retrograde felsicrock in mid-crustal levels.The precursor <strong>of</strong> the Blansky´ les felsic granulites wasan HP alkali feldspar–plagioclase–quartz–garnet–biotite–kyanite bearing coarse-grained layered orthogneissthat developed via deformational solid-statesegregation <strong>of</strong> quartz <strong>and</strong> feldspars. It records peakP–T conditions <strong>of</strong> 1.6)1.8 GPa <strong>and</strong> 850)880 °C(Fig. 16).Granular matrix (Type I microstructure) formed byrecrystallization <strong>of</strong> 10 mm large <strong>and</strong> probablyperthitic strong alkali feldspar porphyroclasts intoa weak fine-grained (0.055 mm) plagioclase–K-feldsparmatrix. This transition occurred via a processresembling discontinuous precipitation mechanism inmetal alloys. It was driven mainly by chemically <strong>and</strong>strain induced grain boundary migration, but triggeredby stress build up during early stages <strong>of</strong> D2. Newlyformed matrix grains maintained roughly the latticeorientation <strong>of</strong> the parental alkali feldspar crystal. Theyhave a sufficiently small grain size to allow flow byGBS.The fine-grained S2 granulitic matrix (Type IImicrostructure) with a very low content <strong>of</strong> silicatemelt deformed predominantly via diffusion creepcontrolledGBS. The large quartz ribbons were rheologicallystronger than the feldspar-dominatedmatrix due to the activity <strong>of</strong> different deformationalmechanisms. During this stage, in only several millionyears, the linearly viscous granulite matrix containinglow amount <strong>of</strong> interstitial melt was extruded upwardsthrough a vertical channel (Franeˇk et al., 2011) frompressures <strong>of</strong> 1.7 GPa at 850 °C to 1.0 GPa at800 °C (Fig. 16).After cooling <strong>and</strong> crystallization <strong>of</strong> the syn-D2partial melt along grain boundaries, the granulitesdeformed at lower P–T conditions in the middle crustduring D3 episode by less-efficient dislocation creep.Local hydration produced partial melt that segregatedfrom solid matrix into isolated b<strong>and</strong>s instead <strong>of</strong>coating grain boundaries.ACKNOWLEDGEMENTSThe work was supported by Grants <strong>of</strong> the Czech ScienceFoundation (GACR 205 ⁄ 05 ⁄ 2187 <strong>and</strong>205 ⁄ 09 ⁄ 0539) <strong>and</strong> the internal project <strong>of</strong> the CzechGeological Survey (No. 326700). Stays <strong>of</strong> J. Franěkat Strasbourg University were funded by theFrench Government Foundation (BGF). The FrenchNational Grant Agency (No.06-1148784) <strong>and</strong> GrantMSM0021620855 <strong>of</strong> the Ministry <strong>of</strong> Education <strong>of</strong> theCzech Republic are acknowledged for financial support<strong>of</strong> K. Schulmann <strong>and</strong> O. Lexa. We are grateful toJ.-E. Martelat for constructive suggestions. We alsothank R. Vernon, H. Stu¨ nitz <strong>and</strong> J. Tullis for theirconstructive revisions. We also thank M. Brown <strong>and</strong>R. White for careful editorial work.CONCLUSIONSREFERENCESAbart, R., Petrishcheva, E., Rhede, D. & Wirth, R., 2009.Exsolution by spinodal decomposition: II: perthite formationduring slow cooling <strong>of</strong> anatexites from Ngornghoro, Tanzania.American Journal <strong>of</strong> Science, 309, 450–475; doi: 10.2475/06.2009.02.Baratoux, L., Lexa, O., Cosgrove, J.W. & Schulmann, K., 2005.The quantitative link between fold geometry, mineral fabric<strong>and</strong> mechanical anisotropy: as exemplified by the deformation<strong>of</strong> amphibolites across a regional metamorphic gradient.Journal <strong>of</strong> Structural Geology, 27, 707–730.Behr, H.J., 1961. Beitrage zur petrographischen und tektonischenAnalyse des sachsischen Granulitgebirges. FrebiergerForschungshefte C, 119, 5–118.Ó 2010 Blackwell Publishing Ltd365


128 J. FRANĚK ET AL.Behrmann, J.H. & Mainprice, D., 1987. Deformation mechanismsin a high-temperature quartz-feldspar mylonite: evidencefor super-plastic flow in the lower crust. Tectonophysics,140, 297–305.Boullier, A.M. & Gueguen, Y., 1975. SP-mylonites; origin <strong>of</strong>some mylonites by superplastic flow. Contributions to Mineralogy<strong>and</strong> Petrology, 50, 93–104.Brown, W.L. & Parsons, I., 1989. Alkali feldspars: orderingrates, phase transformations <strong>and</strong> behaviour diagrams forigneous rocks. Mineralogical Magazine, 53, 25–42.Carswell, D.A. & OÕBrien, P.J., 1993. Thermobarometry <strong>and</strong>geotectonic significance <strong>of</strong> high-pressure granulites – examplesfrom the Moldanubian Zone <strong>of</strong> the Bohemian Massif inLower Austria. Journal <strong>of</strong> Petrology, 34, 427–459.Casey, M., Kunze, K. & Olgaard, D.L., 1998. Texture <strong>of</strong> Solnh<strong>of</strong>enlimestone deformed to high strains in torsion. Journal<strong>of</strong> Structural Geology, 20, 255–267.Cashman, K.V. & Ferry, J.M., 1988. Crystal size distribution(CSD) in rocks <strong>and</strong> the kinetics <strong>and</strong> dynamics <strong>of</strong> crystallization– III. Metamorphic crystallization. Contributions toMineralogy <strong>and</strong> Petrology, 99, 401–415.Coggon, R. & Holl<strong>and</strong>, T., 2002. Mixing properties <strong>of</strong> phengiticmicas <strong>and</strong> revised garnet-phengite thermobarometers. Journal<strong>of</strong> Metamorphic Geology, 20, 683–696.Cooke, R.A., 2000. High-pressure ⁄ temperature metamorphismin the St. Leonhard Granulite Massif, Austria: evidence fromintermediate pyroxene-bearing granulites. International Journal<strong>of</strong> Earth Sciences, 89, 647–651.DellÕAngelo, L.N., Tullis, J. & Yund, R.A., 1987. Transitionfrom dislocation creep to melt-enhanced diffusion creepin fine-grained granitic aggregates. Tectonophysics, 139, 325–332.DeVore, G.W., 1959. Role <strong>of</strong> minimum interfacial free energy indetermining the macroscopic features <strong>of</strong> mineral assemblages.I. The model. Journal <strong>of</strong> Geology, 67, 211–227.Diener, J.F.A., White, R.W. & Powell, R., 2008. Granulite faciesmetamorphism <strong>and</strong> subsolidus fluid-absent reworking,Strangways Range, Arunta Block, central Australia. Journal<strong>of</strong> Metamorphic Geology, 26, 603–622.Fiala, J., Mateˇjovska´, O.&Vanˇkova´, V., 1987. Moldanubiangranulites: source material <strong>and</strong> petrogenetic considerations.Neues Jahrbuch fur Mineralogie, Abh<strong>and</strong>lungen, 157, 133–165.Finger, F., Cooke, R., Janousˇek, V. et al., 2003. Petrogenesis <strong>of</strong>the south Bohemian granulites: the importance <strong>of</strong> crystal-meltrelationships. Journal <strong>of</strong> the Czech Geological Society, 48, 44–45.Flin, D., 1969. Grain contacts in crystalline rocks. Lithos, 2, 361–370.Franeˇk, J., Schulman, K. & Lexa, O., 2006. Kinematic <strong>and</strong>rheological model <strong>of</strong> exhumation <strong>of</strong> high pressure granulites inthe Variscan orogenic root: example <strong>of</strong> the Blansky´ les granulite,Bohemian Massif, Czech Republic. Mineralogy <strong>and</strong>Petrology, 86, 253–276.Franeˇk, J., Schulman, K., Lexa, O., Tomek, C. & Edel, J.B.,2011. Model <strong>of</strong> syn-convergent extrusion <strong>of</strong> orogenic lowercrust in the core <strong>of</strong> the Variscan belt: implications forexhumation <strong>of</strong> high-pressure rocks in large hot orogens.Journal <strong>of</strong> Metamorphic Geology, 29, 53–78.Franke, W., 2000. The mid-European segment <strong>of</strong> the Variscides:tectonostratigraphic units, terrane boundaries <strong>and</strong> plate tectonicevolution. Geological Society Special Publication, 179,35–56.Friedl, G., Cooke, R., Finger, F., McNaughton, N.J. & Fletcher,I., 2003. U-Pb SHRIMP dating <strong>and</strong> trace element investigationson multiple zoned zircons from a South-Bohemiangranulite. Journal <strong>of</strong> the Czech Geological Society, 48, 51.Fuchs, G., 1976. Zur Entwicklung der Bo¨ hmischen Masse.Jahrbuch des Geologischen Bundesanstalts, 119, 45–61.Garlick, S.R. & Gromet, L.P., 2004. Diffusion creep <strong>and</strong> partialmelting in high temperature mylonitic gneisses, Hope Valleyshear zone, New Engl<strong>and</strong> Appalachians, USA. Journal <strong>of</strong>Metamorphic Geology, 22, 45–62.Gayk, T. & Kleinschrodt, R., 2000. Hot contacts <strong>of</strong> garnetperidotites in middle ⁄ upper crustal levels: new constraints onthe nature <strong>of</strong> the late Variscan high-T ⁄ low-P event in theMoldanubian (Central Vosges ⁄ NE France). Journal <strong>of</strong> MetamorphicGeology, 18, 293–305.H<strong>and</strong>y, M., 1990. The solid state flow <strong>of</strong> polymineralic rocks.Journal <strong>of</strong> Geophysical Research, 95, 8647–8661.Hasalová, P., Sˇtı´pska´, P., Powell, R., Schulmann, K., Janousˇek,V. & Lexa, O., 2008a. Transforming mylonitic metagranite byopen-system interactions during melt flow. Journal <strong>of</strong> MetamorphicGeology, 26, 55–80.Hasalová, P., Schulmann, K., Lexa, O. et al., 2008b. Origin <strong>of</strong>migmatites by deformation-enhanced melt infiltration <strong>of</strong>orthogneiss: a new model based on quantitative micro<strong>structural</strong>analysis. Journal <strong>of</strong> Metamorphic Geology, 26, 29–53.Hay, R.S. & Evans, B., 1987. Chemically induced grain boundarymigration in calcite: temperature dependence, phenomenology,<strong>and</strong> possible applications to geologic systems.Contributions to Mineralogy <strong>and</strong> Petrology, 97, 127–141.Higgins, M.D., 1998. Origin <strong>of</strong> anorthosite by textural coarsening:quantitative measurements <strong>of</strong> a natural sequence <strong>of</strong>textural development. Journal <strong>of</strong> Petrology, 39, 1307–1323.Holl<strong>and</strong>, T. & Powell, R., 1998. An internally consistent thermodynamicdata set for phases <strong>of</strong> petrological interest. Journal<strong>of</strong> Metamorphic Geology, 16, 309–343.Holl<strong>and</strong>, T. & Powell, R., 2003. Activity–composition relationsfor phases in petrological calculations: an asymmetric multicomponentformulation. Contributions to Mineralogy <strong>and</strong>Petrology, 145, 492–501.Jakesˇ, P., 1997. Melting in high-P region – case <strong>of</strong> Bohemiangranulites. Acta Universitatis Carolinae, Geologica, 41, 113–125.Janousˇek, V. & Holub, F.V., 2007. The causal link between HP-HT metamorphism <strong>and</strong> ultrapotassic magmatism in collisionalorogens: case study from the Moldanubian Zone <strong>of</strong> theBohemian Massif. Proceedings <strong>of</strong> the Geologists Association,118, 75–86.Janousˇek, V., Finger, F., Roberts, M., Fry´da, J., Pin, C. &Dolejsˇ, D., 2004. Deciphering the petrogenesis <strong>of</strong> deeplyburied granites: whole-rock geochemical constraints onthe origin <strong>of</strong> largely undepleted felsic granulites from theMoldanubian Zone <strong>of</strong> the Bohemian Massif. Transactions <strong>of</strong>the Royal Society <strong>of</strong> Edinburgh – Earth Sciences, 95, 141–159.Janousˇek, V., Gerdes, A., Vrána, S. et al., 2006. Low-pressuregranulites <strong>of</strong> the Lisˇov Massif, Southern Bohemia: viseanmetamorphism <strong>of</strong> Late Devonian plutonic arc rocks. Journal<strong>of</strong> Petrology, 47, 705–744.Jiang, Z., Prior, D.J. & Wheeler, J., 2000. Albite crystallographicpreferred orientation <strong>and</strong> grain misorientation distribution ina low-grade mylonite: implications for granular flow. Journal<strong>of</strong> Structural Geology, 22, 1663–1674.Knipe, R.J., 1989. Deformation mechanisms – recognitionfrom natural tectonites. Journal <strong>of</strong> Structural Geology, 11,127–146.Kodym, O., 1972. Multiphase deformation in the Blanský lesgranulite massif (South Bohemia). Krystalinikum, 9, 91–105.Kosˇler, J., Kelley, S.P., Vance, D. & Svojtka, M., 1999. Independentdating <strong>of</strong> cooling <strong>and</strong> decompression <strong>of</strong> high graderocks in the southern Bohemian Massif with Ar-Ar, Sm-Nd<strong>and</strong> U-Pb techniques. Journal <strong>of</strong> Conference Abstracts, 4, 39.Kotkova´, J. & Harley, S.L., 1999. Formation <strong>and</strong> evolution <strong>of</strong>high-pressure leucogranulites: experimental constraints <strong>and</strong>unresolved issues. Physics <strong>and</strong> Chemistry <strong>of</strong> the Earth Part A –Solid Earth <strong>and</strong> Geodesy, 24, 299–304.Kretz, R., 1969. On the spatial distribution <strong>of</strong> crystals in rocks.Lithos, 2, 39–65.Kro¨ ner, A., OÕBrien, P.J., Nemchin, A.A. & Pidgeon, R.T.,2000. Zircon ages for high pressure granulites from SouthBohemia, Czech Republic, <strong>and</strong> their connection to Carboniferoushigh temperature processes. Contributions to Mineralogy<strong>and</strong> Petrology, 138, 127–142.Ó 2010 Blackwell Publishing Ltd366


PRECURSOR AND RHEOLOGY OF VARISCAN GRANULITES 129Kruse, R., Stunitz, H. & Kunze, K., 2001. Dynamic recrystallizationprocesses in plagioclase porphyroclasts. Journal <strong>of</strong>Structural Geology, 23, 1781–1802.Lexa, O., Sˇtı´pska´, P., Schulmann, K., Baratoux, L. & Kro¨ ner,A., 2005. Contrasting textural record <strong>of</strong> two distinct metamorphicevents <strong>of</strong> similar P–T conditions <strong>and</strong> different durations.Journal <strong>of</strong> Metamorphic Geology, 23, 649–666.Lexa, O., Schulmann, K., Janousˇek, V., Sˇtı´pska´, P., Guy, A. &Racek, M., 2011. Heat sources <strong>and</strong> trigger mechanisms <strong>of</strong>exhumation <strong>of</strong> HP granulites in Variscan orogenic root.Journal <strong>of</strong> Metamorphic Geology, 29, 79–102.Lister, G.S. & Dornsiepen, U.F., 1982. Fabric transition in theSaxony granulite terrain. Journal <strong>of</strong> Structural Geology, 4, 81–92.Lister, G.S. & Price, G.P., 1978. Fabric development in aquartz–feldspar mylonite. Tectonophysics, 49, 37–78.Marsh, B.D., 1988. Crystal size distributions (CSD) in rocks <strong>and</strong>the kinetics <strong>and</strong> dynamics <strong>of</strong> crystallization – I. Theory.Contributions to Mineralogy <strong>and</strong> Petrology, 99, 277–291.Martelat, J.E., Schulmann, K., Lardeaux, J.M., Nicollet, C. &Cardon, H., 1999. Granulite micr<strong>of</strong>abrics <strong>and</strong> deformationmechanisms in southern Madagascar. Journal <strong>of</strong> StructuralGeology, 21, 671–687.Matte, P., Maluski, H., Rajlich, P. & Franke, W., 1990. Terraneboundaries in the Bohemian Massif: result <strong>of</strong> the large–scaleVariscan shearing. Tectonophysics, 177, 151–170.Mehnert, K.R., 1968. Migmatites <strong>and</strong> the Origin <strong>of</strong> GraniticRocks. Elsevier, Amsterdam, 393 pp.Montardi, Y. & Mainprice, D., 1987. A transmission electronmicroscopy study <strong>of</strong> the natural plastic deformation <strong>of</strong>calcic plagioclases (An 68–70). Bulletin de Mineralogie, 110,1–14.OÕBrien, P.J. & Carswell, D.A., 1993. Tectonometamorphicevolution <strong>of</strong> the Bohemian Massif – evidence from highpressuremetamorphic rocks. Geologische Rundschau, 82, 531–555.OÕBrien, P.J. & Ro¨ tzler, J., 2003. High-pressure granulites: formation,recovery <strong>of</strong> peak conditions <strong>and</strong> implications fortectonics. Journal <strong>of</strong> Metamorphic Geology, 21, 3–20.OÕBrien, P.J. & Seifert, K., 1992. P–T–t paths as records <strong>of</strong>orogenic processes: examples <strong>and</strong> problems from the crystalline<strong>of</strong> the Bohemian Massif. Terra Nostra, 1992, 58–59.Passchier, C.W., Trouw, R.A.J., Zwart, H.J. & Vissers, R.L.M.,1992. Porphyroblast rotation – Eppur-Si-Muove. Journal <strong>of</strong>Metamorphic Geology, 10, 283–294.Powell, R. & Holl<strong>and</strong>, T., 1988. An internally consistent datasetwith uncertainties <strong>and</strong> correlations. 3. Applications to geobarometry,worked examples <strong>and</strong> a computer-program.Journal <strong>of</strong> Metamorphic Geology, 6, 173–204.Powell, R. & Holl<strong>and</strong>, T., 2004. Course notes for Ô THERMO-CALC workshop 2004: calculating metamorphic phase equilibriaÕ(ETH Zurich, Switzerl<strong>and</strong>). CD-ROM.Putnis, A., 2002. Mineral replacement reactions: from macroscopicobservations to microscopic mechanisms. MineralogicalMagazine, 66, 689–708.Putnis, A., Pina, C.M., Astilleros, J.M., Fernández-Díaz, L. &Prieto, M., 2003. Nucleation <strong>of</strong> solid solutions crystallisingfrom aqueous solutions. Philosophical Transactions <strong>of</strong> theRoyal Society A: Mathematical, Physical <strong>and</strong> EngineeringSciences, 361, 615–632.Racek, M., Sˇtı´pska´, P., Pitra, P., Schulmann, K. & Lexa, O.,2006. Metamorphic record <strong>of</strong> burial <strong>and</strong> exhumation <strong>of</strong> orogeniclower <strong>and</strong> middle crust: a new tectonothermal model forthe Drosendorf window (Bohemian Massif, Austria). Mineralogy<strong>and</strong> Petrology, 86, 221–251.Ramberg, H., 1952. The Origin <strong>of</strong> Metamorphic <strong>and</strong> MetasomaticRocks. University <strong>of</strong> Chicago Press, Chicago, 317 pp.Ribbe, P.H. (ed.), 1983. Feldspar Mineralogy, 2nd edn. Mineralogicalsociety <strong>of</strong> America, Washington, 362 pp.Rosenberg, C.L. & H<strong>and</strong>y, M.R., 2005. Experimental deformation<strong>of</strong> partially melted granite revisited: implications forthe continental crust. Journal <strong>of</strong> Metamorphic Geology, 23,19–28.Rutter, E.H. & Brodie, K.H., 1988. The role <strong>of</strong> tectonic grainsize reduction in the rheological stratification <strong>of</strong> the lithosphere.Geologische Rundschau, 77, 295–307.Rutter, E.H. & Brodie, K.H., 1992. Rheology <strong>of</strong> the lower crust.In: Continental Lower Crust (eds Fountain, D.M., Arculus, R.& Kay, R.W.), pp. 201–267. Elsevier, Amsterdam.Saheb, N., Boumerzoug, Z., Hamana, D., Laoui, T. & Van DerBiest, O., 1995. Different types <strong>of</strong> discontinuous precipitationin Cu-15 Wt% in alloy. Scripta Metallurgica et Materiala, 32,1453–1458.Sakai, T. & Jonas, J.J., 1984. Dynamic recrystallization:mechanical <strong>and</strong> micro<strong>structural</strong> consideration. Acta Metallurgica,32, 198–209.Schmid, S.M. & Casey, M., 1986. Complete fabric analysis <strong>of</strong>some commonly observed quartz c-axis patterns. In: Mineral<strong>and</strong> Rock Deformation: Laboratory Studies (eds Hobbs, B.E. &Heard, H.C.), pp. 263–286. American Geophysical UnionMonograph, Vol. 36, Washington, DC.Schulmann, K., Kro¨ ner, A., Hegner, E. et al., 2005. Chronologicalconstraints on the pre-orogenic history, burial <strong>and</strong>exhumation <strong>of</strong> deep-seated rocks along the eastern margin <strong>of</strong>the Variscan Orogen, Bohemian Massif, Czech Republic.American Journal <strong>of</strong> Science, 305, 407–448.Schulmann, K., Martelat, J.-E., Ulrich, S., Lexa, O., Sˇtı´pska´, P.& Becker, J.K., 2008. Evolution <strong>of</strong> microstructure <strong>and</strong> melttopology in partially molten granitic mylonite: implicationsfor rheology <strong>of</strong> felsic middle crust. Journal <strong>of</strong> GeophysicalResearch, 113, B10406.Schulmann, K., Konopa´sek, J., Janousˇek, V. et al., 2009. AnAndean type Palaeozoic convergence in the Bohemian Massif.Comptes Rendus – Geoscience, 341, 266–286.Seng, H., 1936. Die Migmatitfrage und der Mechanismusparakristalliner Pragung. Geologische Rundschau, 27, 471–492.Sennour, M., Jouneau, P.H. & Esnouf, C., 2004. TEM <strong>and</strong>EBSD investigation <strong>of</strong> continuous <strong>and</strong> discontinuous precipitation<strong>of</strong> CrN in nitrided pure Fe-Cr alloys. Journal <strong>of</strong>Materials Science, 39, 4521–4531.Sˇtı´pská, P. & Powell, R., 2005. Does ternary feldspar constrainthe metamorphic conditions <strong>of</strong> high-grade metaigneousrocks? Evidence from orthopyroxene granulites,Bohemian Massif. Journal <strong>of</strong> Metamorphic Geology, 23, 627–647.Sˇtı´pská, P., Powell, R., White, R.W. & Baldwin, J.A., 2010.Using calculated chemical potential relationships to accountfor coronas around kyanite: an example from the BohemianMassif. Journal <strong>of</strong> Metamorphic Geology, 28, 97–116.Stu¨ nitz, H., 1998. Syndeformational recrystallization – dynamicor compositionally induced? Contributions to Mineralogy <strong>and</strong>Petrology, 131, 219–236.Stu¨ nitz, H. & Tullis, J., 2001. Weakening <strong>and</strong> strain localizationproduced by syn-deformational reaction <strong>of</strong> plagioclase.International Journal <strong>of</strong> Earth Sciences, 90, 136–148.Stu¨ nitz, H., Fitz Gerald, J.D. & Tullis, J., 2003. Dislocationgeneration, slip systems, <strong>and</strong> dynamic recrystallization inexperimentally deformed plagioclase single crystals. Tectonophysics,372, 215–233.Tajčmanova´, L., Konopa´sek, J. & Schulmann, K., 2006. Thermalevolution <strong>of</strong> the orogenic lower crust during exhumationwithin a thickened Moldanubian root <strong>of</strong> the Variscan belt<strong>of</strong> Central Europe. Journal <strong>of</strong> Metamorphic Geology, 24, 119–134.Tajčmanova´, L., Konopa´sek, J. & Connolly, J.A.D., 2007.Diffusion-controlled development <strong>of</strong> silica-undersaturateddomains in felsic granulites <strong>of</strong> the Bohemian Massif (Variscanbelt <strong>of</strong> Central Europe). Contributions to Mineralogy <strong>and</strong>Petrology, 153, 237–250.Tropper, P., Konzett, J. & Finger, F., 2005. Experimental constraintson the formation <strong>of</strong> high-P ⁄ high-T granulites in theÓ 2010 Blackwell Publishing Ltd367


130 J. FRANĚK ET AL.Southern Bohemian Massif. European Journal <strong>of</strong> Mineralogy,17, 343–356.Tullis, J., 1983. Deformation <strong>of</strong> feldspars. In: Feldspar Mineralogy(ed. Ribbe, P.H.), pp. 297–323. Mineralogical Society <strong>of</strong>America, Washington.Ulmer, P., 1986. Norm-Program for Cation <strong>and</strong> Oxygen MineralNorms. Computer Library IKP-ETH, Zu¨ rich.Ulrich, S., Schulmann, K. & Casey, M., 2002. Micro<strong>structural</strong>evolution <strong>and</strong> rheological behaviour <strong>of</strong> marbles deformed atdifferent crustal levels. Journal <strong>of</strong> Structural Geology, 24, 979–995.Ulrich, S., Thompson, A.B., Schulmann, K. & Casey, M., 2006.Microstructure mechanism map <strong>of</strong> dynamically recrystallizedmarble. Tectonophysics, 412, 173–182.Verner, K., Zˇák, J., Nahodilová, R. & Holub, F.V., 2007.Magmatic fabrics <strong>and</strong> emplacement <strong>of</strong> the cone-sheet-bearingKnízˇecı´ Stolec durbachitic pluton (Moldanubian Unit, BohemianMassif): implications for mid-crustal reworking <strong>of</strong>granulitic lower crust in the Central European Variscides.International Journal <strong>of</strong> Earth Sciences, 97, 19–33.Vrána, S., 1989. Perpotassic granulites from Southern Bohemia –a new rock-type derived from partial melting <strong>of</strong> crustal rocksunder upper mantle conditions. Contributions to Mineralogy<strong>and</strong> Petrology, 103, 510–522.Wendt, J.I., Kro¨ ner, A., Fiala, J. & Todt, W., 1994. U-Pb zircon<strong>and</strong> Sm-Nd dating <strong>of</strong> Moldanubian HP ⁄ HT granulites fromsouth Bohemia, Czech Republic. Journal <strong>of</strong> the GeologicalSociety London, 151, 83–90.White, R.W., Powell, R., Holl<strong>and</strong>, T. & Worley, B.A., 2000. Theeffect <strong>of</strong> TiO 2 <strong>and</strong> Fe 2 O 3 on metapelitic assemblages atgreenschist <strong>and</strong> amphibolite facies conditions: mineral equilibriacalculations in the system K 2 O–FeO–MgO–Al 2 O 3 –SiO 2 –H 2 O–TiO 2 –Fe 2 O 3 . Journal <strong>of</strong> Metamorphic Geology, 18,497–511.White, R.W., Powell, R. & Holl<strong>and</strong>, T., 2007. Progress relatingto calculation <strong>of</strong> partial melting equilibria for metapelites.Journal <strong>of</strong> Metamorphic Geology, 25, 511–527.Wilson, G., 1961. The tectonic significance <strong>of</strong> small–scalestructures, <strong>and</strong> their importance to the geologist in the field.Annales de la Socie´te´ Ge´ologique de Belgique, 84, 423–595.Yoon, D.Y., 1995. Theories <strong>and</strong> observations <strong>of</strong> chemicallyinduced interface migration. International Materials Reviews,40, 149–179.Za´vada, P., Schulmann, K., Konopa´sek, J., Ulrich, S. & Lexa,O., 2007. Extreme ductility <strong>of</strong> feldspar aggregates – meltenhancedgrain boundary sliding <strong>and</strong> creep failure: rheologicalimplications for felsic lower crust. Journal <strong>of</strong> GeophysicalResearch B: Solid Earth, 112, B10210.Zener, C. & Holomon, J.H., 1944. Effect <strong>of</strong> strain rate uponplastic flow <strong>of</strong> steel. Journal <strong>of</strong> Applied Physics, 15, 22–32.SUPPORTING INFORMATIONAdditional Supporting Information may be found inthe online version <strong>of</strong> this article:Appendix S1. <strong>Quantitative</strong> characteristics <strong>of</strong> theindividual micro<strong>structural</strong> evolutionary stages.Please note: Wiley-Blackwell are not responsible forthe content or functionality <strong>of</strong> any supporting materialssupplied by the authors. Any queries (other thanmissing material) should be directed to the correspondingauthor for the article.Received 26 March 2010; revision accepted 6 September 2010.Ó 2010 Blackwell Publishing Ltd368

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!