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.

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

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

Saved successfully!

Ooh no, something went wrong!