1,073
Views
3
CrossRef citations to date
0
Altmetric
Science

Geology of the contact area between the Internal and External Nappe Zone of the Sardinian Variscan Belt (Italy): new insights on the complex polyphase deformation occurring in the hinterland-foreland transition zone of collisional belts

ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon
Pages 472-483 | Received 14 Dec 2021, Accepted 14 Jun 2022, Published online: 11 Jul 2022

ABSTRACT

This study aims to decipher the structural evolution of a sector of the Nappe Zone in the Sardinian Variscan Belt. We investigated the Barbagia Thrust between the Meana Sardo Unit (belonging to the External Nappe Zone) in the footwall and the Barbagia Unit (belonging to the Internal Nappe Zone) in the hanging wall. Combining the geological survey with meso- and microstructural analysis, we realized a 1:10.000 scale geological map highlighting the polyphasic evolution developed under low-grade metamorphic conditions during the Variscan orogeny. Both units preserve (i) an early phase generally observed far from the tectonic contact and mainly in the Meana Sardo Unit (D1), (ii) a syn-nappe ductile deformation linked to the Barbagia Thrust activity and a top-to-the S-SW sense of shear (D2) and (iii) a large-scale nappe refolding (D3). A late extensional stage (D4), with the development of collapse folds, marks the end of the orogenic cycle.

1. Introduction

In collisional belts, the gradual transition from tectonic units occurring in the metamorphic core of the belt to those deformed at higher structural levels and progressively incorporated in the orogenic wedge is identified as the hinterland-foreland transition zone, forming a fold-and-thrust belt (CitationLarson et al., Citation2010; CitationMontomoli et al., Citation2018; CitationSchneider et al., Citation2014; CitationThigpen et al., Citation2010). Generally, these orogenic sectors are characterized by large-scale nappe stacks with different generations of superimposed folds and cleavages that complicate the understanding of the structural architecture (CitationGhosh et al., Citation2020; CitationMontomoli et al., Citation2018; CitationSarkarinejad & Ghanbarian, Citation2014; CitationThigpen et al., Citation2010, Citation2013). Nappe boundaries are often marked by the presence of regional-scale contractional (thrust-sense) ductile shear zones that accommodate large displacements and control the growth of orogenic wedges (CitationGhosh et al., Citation2020; CitationGrasemann et al., Citation1999; CitationRing & Brandon, Citation1999; CitationThigpen et al., Citation2010). Several difficulties in unravelling the tectono-metamorphic evolution of the hinterland-foreland transition zone arise from the overprinting of post-nappe stacking deformation that modifies the original attitude of pre- and syn-nappe stacking elements (e.g. secondary foliations, tectonic contacts, folds vergence). Untangling the geological complexity derived from the geological survey and outcrop observation requires microstructural analysis (e.g. CitationCarosi et al., Citation2016; CitationCaso et al., Citation2021; CitationPapeschi et al., Citation2020).

The Variscan Belt in Sardinia represents an excellent site where it is possible to investigate the tectonic mechanisms active during the evolution of collisional-type orogenesis during the Paleozoic, as this area has not been strongly reworked by the later Alpine orogenic cycle (CitationCarmignani et al., Citation1994, Citation2001). The Sardinian hinterland-foreland transition zone has been subdivided into External and Internal Nappe Zone (CitationCarmignani et al., Citation1994, Citation2001). The Barbagia Thrust (BT; CitationCarosi & Malfatti, Citation1995; CitationCarosi & Pertusati, Citation1990; CitationMontomoli et al., Citation2018; CitationPetroccia et al., Citation2022), a thrust-sense movement, regional ductile to brittle shear zone (CitationCarosi et al., Citation2002, Citation2004; CitationCarosi & Malfatti, Citation1995; CitationConti et al., Citation1998, Citation1999, Citation2001), marks the boundary between both nappes. It plays a critical role during the nappe stacking and the exhumation of the crustal units. At the end of the compressive deformation, the complete nappe pile and relative tectonic contacts have been refolded by post-nappe stacking regional-scale antiforms and synforms that complicate the structural architecture. Whereas the setting of different portions of the hinterland of the Sardinian belt has been well investigated by several authors (CitationCarosi et al., Citation2020, Citation2022; CitationCarosi & Palmeri, Citation2002; CitationCasini et al., Citation2010, Citation2015; CitationCruciani et al., Citation2015), few studies have addressed the understanding of the structural architecture of the low-grade Nappe Zone (CitationCarmignani et al., Citation1994; CitationConti et al., Citation1999, Citation2001; CitationFunedda, Citation2009; CitationMontomoli et al., Citation2018; CitationPetroccia et al., Citation2022). In this work, we present a new 1:10.000 scale geological map (Main Map), supported by unpublished meso- and microstructural data. The excellent exposures of both rocks and structures make this sector of the Sardinian Belt a fundamental area to unravel the complex polyphasic history that occurred during the Variscan pre-, syn- and post-nappe stacking deformation.

2. Geological setting

The Sardinian Paleozoic basement is a segment of the Southern European Variscan belt (CitationMatte, Citation2001). The metamorphic grade increases from very low-grade on the southwestern side of the island up to medium- and high-grade in the northern sector. In particular, the belt is classically described as a collisional-type chain (CitationCarmignani et al., Citation1994, Citation2001, Citation2016; CitationCruciani et al., Citation2015 for a critical review) consisting of ((a)): (i) the External Zone or foreland, (ii) the Nappe Zone or hinterland-foreland transition zone and (iii) the Axial Zone or the hinterland of the belt.

Figure 1. (a) Tectonic sketch map of the Sardinia Island (modified after CitationCarosi et al., Citation2020; CitationMontomoli et al., Citation2018). (b) Schematic geological cross-section (A-A’) along the Nappe Zone (modified after CitationCarmignani et al., Citation1994; CitationConti et al., Citation1999, Citation2001).

Two different images and sketches describing the geographical, geological, and structural setting of the study area.
Figure 1. (a) Tectonic sketch map of the Sardinia Island (modified after CitationCarosi et al., Citation2020; CitationMontomoli et al., Citation2018). (b) Schematic geological cross-section (A-A’) along the Nappe Zone (modified after CitationCarmignani et al., Citation1994; CitationConti et al., Citation1999, Citation2001).

The Nappe Zone has been classically subdivided into External (central to southern Sardinia) and Internal (northern to central Sardinia) Nappe Zone ((a)). The lithostratigraphic succession is similar in both nappes: Lower Cambrian to Ordovician metasandstone, phyllite and quartzite (CitationCarmignani et al., Citation1994; CitationCocco et al., Citation2018; CitationCocco & Funedda, 2019; CitationPertusati et al., Citation2002), Middle Ordovician metaconglomerate and metavolcanic rocks (CitationOggiano et al., Citation2010), Upper Ordovician meta-arkose and metasiltstone, Silurian–Lower Devonian black shale, phyllite, and marble, Middle-Upper Devonian marble, and Lower Carboniferous synorogenic flysch (CitationConti et al., Citation2001). The main difference between Internal and External nappes is the lack or the paucity of Ordovician metavolcanic rocks and Devonian limestones in the Internal Nappe Zone sequence (CitationCarmignani et al., Citation1994, Citation2001).

Meana Sardo, Gerrei, Rio Gruppa, Castello Medusa and Sarrabus units belonging to the External Nappe Zone ((b); CitationCarmignani et al., Citation1994; CitationCarosi et al., Citation1991; CitationCocco & Funedda, Citation2011, Citation2012; CitationFunedda et al., Citation2011, Citation2015) are mainly deformed under syn-tectonic regional greenschist-facies metamorphism (CitationCarmignani et al., Citation1994; CitationCarosi et al., Citation1991, Citation2010; CitationCocco et al., Citation2018; CitationFranceschelli et al., Citation1992). The only exception is the Monte Grighini Unit, where amphibolite-facies conditions have been detected (CitationCruciani et al., Citation2016; CitationMusumeci, Citation1991). Two main metamorphic complexes constitute the Internal Nappe Zone: the Low-Grade Metamorphic Complex (LGMC; Barbagia, Goceano, and southern Nurra units; CitationCarmignani et al., Citation1994; CitationMontomoli, Citation2003) in greenschists-facies and the Medium Grade Metamorphic Complex (MGMC; Baronie, Anglona, and northern Nurra units; CitationCarmignani et al., Citation1982, Citation1994, Citation2001) reaching amphibolite-facies conditions. The Barbagia Thrust (BT; CitationCarosi & Malfatti, Citation1995; CitationCarosi & Pertusati, Citation1990; CitationMontomoli et al., Citation2018; CitationPetroccia et al., Citation2022) is a regional-scale ductile to brittle shear zone that extends from central to southeastern Sardinia for more than 50 km along strike and marks the boundary between the Internal and the External Nappe zones.

Meso- and microstructural observations from different authors (CitationCarmignani et al., Citation1982, Citation1994; CitationCarmignani & Pertusati, Citation1977; CitationCarosi et al., Citation2004; CitationCarosi & Pertusati, Citation1990; CitationConti et al., Citation1998; CitationConti & Patta, Citation1998; CitationDessau et al., Citation1982), point to a complex polyphase evolution. Four deformation phases have been detected and described in both Internal and External Nappe Zone (CitationCarosi et al., Citation2002, Citation2004; CitationConti et al., Citation1998, Citation2001). D1 and D2 are related to continental collision and shortening, responsible for nappe emplacement, followed by later D3 and D4 phases. The D1 tectonic event is generally associated with the development of F1 pluridecametric S-SW-verging folds, better preserved in the External Nappe Zone (CitationCarosi, Citation2004; CitationCarosi et al., Citation2002, Citation2004; CitationCarosi & Malfatti, Citation1995; CitationConti et al., Citation1998, Citation1999). Approaching the BT, F1 folds become much tighter and their wavelength decreases (CitationCarosi et al., Citation2004). The D1 phase was progressively obliterated by the D2 event developed during the main nappe emplacement stage and linked to the BT activity (CitationMontomoli et al., Citation2018 and references therein). A S2 mylonitic foliation with kinematic indicators pointing to a top-to-the S-SW sense of shear is recognizable. The BT is characterized by ductile to brittle deformation, highlighted by cataclastic deformation overprinting mylonite (CitationCarosi et al., Citation1991; CitationCarosi & Malfatti, Citation1995). Greenschist facies deformation is highlighted by syn-tectonic recrystallization of muscovite, quartz, albite and chlorite. After nappe-stacking deformation linked to D1 and D2 phases, the D3 phase is responsible for regional-scale antiforms and synforms (Flumendosa Antiform, Barbagia Synform and Gennargentu Antiform; CitationCarmignani et al., Citation1994; CitationConti et al., Citation1999; (a) and (b)). The entire nappe pile, the BT, older foliations and folds ((b)) were refolded by these kilometre-scale upright folds. The end of the orogenic activity is characterized by the extensional collapse of the belt (D4 phase), associated with open folds, brittle-ductile shear zones and by the emplacement of the Sardo-Corso batholith at ∼320-280 Ma (CitationCasini et al., Citation2012; CitationDel Moro et al., Citation1975). The crystalline basement is unconformably covered by Upper Carboniferous and Permian sequences, followed by Mesozoic carbonates (CitationCostamagna & Barca, Citation2004) and by a volcano-sedimentary succession emplaced during the Oligo-Miocene time (CitationFunedda et al., Citation2011).

The study area, located in the Barbagia region in the southwestern sector of the Lago Alto del Flumendosa ((a)), is represented by the structurally upper Barbagia Unit (BU, Internal Nappe Zone) overthrust on the Meana Sardo Unit (MSU, External Nappe Zone). We focused on a key area of the hinterland-foreland transition zone where the contact between the Internal and External nappes is marked by a well-exposed mylonitic zone. The BT and all previous structural elements (e.g. F2 folds, S2 foliation) were subsequently refolded by an upright kilometre-scale antiform, the Gennargentu Antiform. The study area is located in its southern limb ((a)), where previous F2 folds change their attitude and F2 axial planes dip towards the S-SW and not to the hinterland or N-NE (CitationCarosi et al., Citation2004; CitationMontomoli et al., Citation2018).

3. Methods

The study area is located in the southwestern sector of the Lago Alto del Flumendosa, in central Sardinia, and covers an area of ∼20 km2. New 10 m contour lines have been obtained using the DTM TINITALY/01 (CitationTarquini et al., Citation2007) with 10 m-cell size m grid resolution, integrated into a WGS 84, UTM 32N GIS spatial database to derive a Hillshade map. The geological map (see Main Map) has been realized at 1:10.000 scale. To better document the distribution of the metamorphic basement, we do not report quaternary deposits in the resulting map. The mapping workflow includes (i) mapping of both lithologies and structures; (ii) measurement of structural elements: (S) for syn-metamorphic surfaces or axial plane foliation; (A) for fold axes; (L) for object lineations; (F) for folds; and (D) for the deformation phases; and (iii) microstructural analysis on oriented thin sections, cut parallel to the object lineation and perpendicular to the main foliation (approximating the XZ plane of the finite strain ellipsoid). A numerical progression in the description of deformation phases, besides the primary bedding S0, has been used (e.g. S1, S2). All the collected structural elements have been plotted in equal-area stereo diagrams in the lower hemisphere. Mineral abbreviations are after CitationWhitney and Evans (Citation2010) except for white mica, Wm. Foliations, kinematic indicators and mylonites have been classified according to CitationPasschier and Trouw (Citation2005). Quartz microstructures, indicative of dynamic recrystallization, are defined according to CitationPiazolo and Passchier (Citation2002), CitationStipp et al. (Citation2002) and CitationLaw (Citation2014). Mylonite has been classified considering the percentage of the matrix compared to porphyroclasts, varying from 50-90% for mylonitic rocks to more than 90% for ultramylonite.

4. Lithostratigraphy

The Barbagia Unit succession (also known as ‘Postgotlandiano’; CitationVai & Cocozza, Citation1974) has been mapped as a single metasedimentary formation (Filladi Grigie del Gennargentu Fm., Fg; Main Map). It is constituted by quartzitic and micaceous metasandstone alternating with metasiltstone and quartzite interbedded with grey phyllite. Sometimes, close to the BT, a thin and not mappable layer of porphyroids has been recognized (Main Map). The protolith age was referred from Cambrian to Ordovician (CitationCarmignani et al., Citation1994; CitationPertusati et al., Citation2002). Meana Sardo Unit begins with a basal silicoclastic succession (Arenarie di Solanas-San Vito Fm., Sv; CitationBarca et al., Citation1988; CitationCalvino, Citation1959; Main Map), dated from Middle Cambrian to Lower Ordovician (CitationDi Milia & Tongiorgi, Citation1993). It is made by alternating fine- to coarse-grained metasandstone, grey-green metasiltstone, and meta-argillite, with thickness varying from cm- to m-scale. The succession continues with volcanic rocks belonging to the Monte Santa Vittoria Fm. (Ms; CitationCarmignani et al., Citation2001; Main Map). This formation includes three lithostratigraphic units that were previously informally defined by CitationMinzoni (Citation1975), including the Manixeddu, Monte Corte Cerbos, and Serra Tonnai formations (CitationCarosi et al., Citation2004). It is mainly represented by metaconglomerate, metavolcanoclastite, grey-green metarhyolite, green metabasalt and dacitic metaignimbrite. The age of the protolith is ∼460 Ma (Middle-Late Ordovician) based on radiometric analyses (CitationPavanetto et al., Citation2012), confirming previous stratigraphic ages (CitationCarmignani et al., Citation1994). A change from volcanic to siliciclastic rocks is testified by the presence of Upper Ordovician terrigenous formation (Orroeledu Fm., Or; CitationLoi, Citation1993; CitationLoi & Dabard, Citation1997; Main Map). It consists of metarkose, metagraywacke and phyllite with quartz–sericite–chlorite matrix. These metasediments locally contain interbedded basic to intermediate metaepiclastite. The youngest lithological unit is made by black graphitic shales with interbedded metalimestone layers (Scisti a Graptoliti Fm., Sg; Main Map) that are Silurian to Early Devonian in age (CitationBarca & Jäger, Citation1989; CitationCorradini et al., Citation1998; CitationGnoli, Citation1993). The post-Variscan magmatic rocks are made by northwest-trending rhyolitic to rhyodacitic aplitic dykes (α; Upper Carboniferous-Permian; CitationCarmignani et al., Citation2001). They contain phenocrysts of quartz, feldspar, and plagioclase, as well as aggregates of biotite crystals within a fine-grained groundmass. The Triassic-Jurassic rocks are recognizable in the Perda et Liana inselberg. They are characterized by dolostone, which is typically massive and structureless, alternating shale and dolomitic limestone, quartzitic sandstone and conglomerate (Triassic-Jurassic Genna Selole and Dorgali Fm., C; CitationCostamagna & Barca, Citation2004).

5. Structural analysis

Based on overprinting criteria and structures observed from meso- to microscale, we defined four ductile deformation phases heterogeneously distributed in both units. The original bedding (S0; (a)) is only recognizable as a compositional alternation or dismembered lens-shaped fragments in sectors where a strong lithological contrast, e.g. quartz layers within a metasandstone matrix, is present. Due to the strong transposition related to the D2 phase in the investigated area, it is nearly impossible to identify fold systems associated with the D1 phase. The D1 structural elements are well recognizable, mainly in the MSU far from the BT. S1 foliation can be identified in the hinge zone of the F2 folds or as meso- ((b)) and micro-scale microlithons ((c)). In the thin section, the S1 is marked by syn-kinematic recrystallization of white mica, chlorite, quartz, calcite, opaque minerals and rare albite ((d)). The intensity of strain increases toward the BT, as does the mylonitic foliation. It is worth noting that the frequency of the occurrence of the D1 structures decreases approaching the BT, due to the D2 overprinting. Structures related to the D2 phase are tight to isoclinal, overturned to recumbent folds, and developed from micro- to map-scale. The interlimb angles of F2 folds range from 60° to ∼2° (close to sub-isoclinal; (e)) and vary from gently inclined to recumbent. F2 folds generally show rounded and thickened hinges with stretched limbs (class 2 of CitationRamsay, Citation1967), which can be classified as B5 according to CitationHudleston (Citation1973). F2 folds show an S2 foliation parallel or sub-parallel to the relative fold axial planes, and it generally represents the main foliation at the outcrop-scale. The strike of the S2 foliation is quite constant, ranging from E-W to NW-SE with scattered values ((d)). A2 fold axes show a main E–W trend gently plunging toward both E and W, with quite scattered values ((e)). A N-S or NE-SW trend of the L2 object lineation on the S2 foliation is recognizable. The F2 fold axes are nearly perpendicular to the L2.

Figure 2. Outcrop and microscale view of the main structural features. The structural equal-areal projections (lower hemisphere) of the S2 poles and L2 (d), A2 (e), A3 (f) and A4 (h) axes are shown. (a) Microscopic evidence of S2 gradational crenulation cleavage. S0 bedding and S1 foliation, parallel to S0, are also present. (b, c) S2 spaced foliation at both meso- and microscale (XPL: crossed-polarized light). In the thin section, S1 is marked by Wm + Chl. (d) Fine continuous S2 foliation is defined by Wm + Chl (XPL). (e) F2 fold, deforming the S0//S1 foliation in metasedimentary rocks belonging to the BU. (f,g) F3 folds, deforming the S2 foliation in metasedimentary rocks belonging to the MSU. No clear S3 axial-plane foliation in the F3 hinge zone is recognizable. (h) Late open folds (F4) with sub-horizontal axes and axial planes deforming S2 spaced foliation.

Six images showing the complex polyphase deformation history.
Figure 2. Outcrop and microscale view of the main structural features. The structural equal-areal projections (lower hemisphere) of the S2 poles and L2 (d), A2 (e), A3 (f) and A4 (h) axes are shown. (a) Microscopic evidence of S2 gradational crenulation cleavage. S0 bedding and S1 foliation, parallel to S0, are also present. (b, c) S2 spaced foliation at both meso- and microscale (XPL: crossed-polarized light). In the thin section, S1 is marked by Wm + Chl. (d) Fine continuous S2 foliation is defined by Wm + Chl (XPL). (e) F2 fold, deforming the S0//S1 foliation in metasedimentary rocks belonging to the BU. (f,g) F3 folds, deforming the S2 foliation in metasedimentary rocks belonging to the MSU. No clear S3 axial-plane foliation in the F3 hinge zone is recognizable. (h) Late open folds (F4) with sub-horizontal axes and axial planes deforming S2 spaced foliation.

In the MSU metavolcanic rocks, the principal anisotropy is a continuous and pervasive S2 foliation. In the metapelite and metasandstone, the S2 can be generally classified as a gradational to discrete spaced crenulation cleavage. Far from the tectonic contact, the S2 is mainly associated with pressure solution and reorientation of pre-existing tabular grains ((c)). Moving toward the BT, the S2 becomes a mylonitic continuous foliation ((d)) associated with the blastesis of white mica + chlorite. The D3 phase is characterized by meso- to macroscopic scale, E–W trending F3 folds that overprint and refold all previous structural elements ((f,g)). A3 fold axes ((f)) generally trend parallel to the A2 ones and perpendicular to the L2 object lineation. The F3 structures are gentle to open S-SE verging upright to steeply inclined parallel or asymmetric folds with centimetric ((f)) to plurimetric long wavelength. Locally, kink type folds occur. In the F3 hinge zones and only in less competent rocks, the S3 axial-plane foliation is represented by a spaced, disjunctive cleavage. No metamorphic mineral assemblage related to this phase has been observed, whereas the main deformation mechanism is pressure solution and reorientation of pre-existing grains. D3 folds affect all previous structural elements, determining the variation of the axial plane dipping of the F2 folds. The studied sector is indeed located in the southern limb of one of these regional-scale folds (i.e. Gennargentu Antiform), where F2 folds display S-SW dip of the axial plane and not toward the N-NE as well explained by previous knowledge (CitationCarosi et al., Citation2004; CitationMontomoli et al., Citation2018). Thus, all F2 folds display a S-SW vergence, but they are anticlinal synforms and synclinal antiforms (see the Main Map). The last D4 phase is associated with gentle to open folds, with sub-horizontal axes and axial planes ((h)). Locally, some minor-scale kink-folds occur. No axial plane and metamorphic mineral assemblage related to the D4 phase has been observed. A late brittle deformation with N-S striking faults has been recognized.

6. The Barbagia Thrust

In the mapped area, the BT is characterized by a hm-thick mylonitic zone. The BT-related structures (D2) overprint both metasedimentary and metavolcanic or metavolcanoclastic rocks of the MSU and metasandstone, metasiltstone, and metapsammite of the BU. In both units, a variation in structural style moving toward the BT across the deformation gradient is recognizable. The intensity of deformation increases toward the BT, where folds become tighter and lineation more pervasive; contemporaneously, the spacing between foliation domains decreases. Approaching the high-strain zone, the mylonitic foliation obliterates the previous D1 structures, thus they are rarely detectable. Kinematic indicators, including C-C’-S fabric and asymmetric porphyroclasts of quartz or feldspar indicating a general top-to-the S-SW normal sense of shear with local variation to SE have been recognized ((a)).

Figure 3. (a) Mesoscopic kinematic indicators of mylonites belonging to the Barbagia Thrust indicating a top-to-the S-SW sense of shear. (b,c) Structural variation going toward the Barbagia Thrust in both units. (b) variation from mylonites (on the left) to ultramylonites (on the right) in metavolcanic rocks belonging to the MSU. (c) Progressive variation of the foliation from spaced cleavage to a continuous cleavage moving toward the BT in metasedimentary rocks.

Three images showing the mesostructural evidence of shear deformation along with the BT and six images showing the variation of deformation and structural features going toward the BT from both units.
Figure 3. (a) Mesoscopic kinematic indicators of mylonites belonging to the Barbagia Thrust indicating a top-to-the S-SW sense of shear. (b,c) Structural variation going toward the Barbagia Thrust in both units. (b) variation from mylonites (on the left) to ultramylonites (on the right) in metavolcanic rocks belonging to the MSU. (c) Progressive variation of the foliation from spaced cleavage to a continuous cleavage moving toward the BT in metasedimentary rocks.

Sheared metavolcanic rocks belonging to MSU are characterized by a progressive grain-size reduction and a transition from mylonite to ultramylonite along a deformation gradient towards the centre of the shear zone ((b)). The foliation wraps around rotated plagioclase and feldspar porphyroclasts. Feldspar shows undulose extinction and rare flame perthite. In ultramylonite, quartz recrystallization mechanism could not be detected due to its extremely small grain size and phase-mixing with mica.

In metasedimentary rocks from both nappes, the intensity of deformation increases going toward the BT. The spacing between foliation domains decreases, and at the same time, the foliation varies from a spaced/disjunctive cleavage to a continuous cleavage ((c)). The mylonitic foliation is defined by elongated quartz grains and phyllosilicate-rich levels. Phyllosilicate layers are dominated by chlorite + white mica. In the farthest samples from the BT, quartz shows a strong undulose extinction, deformation bands, small new-grains and rare brittle deformation features. These structures indicate that the bulging mechanism (BLG) is the main dynamic deformation mechanism (CitationLaw, Citation2014; CitationStipp et al., Citation2002). However, quartz often shows elongated sub-grains with new quartz grains of smaller size surrounding larger grains, generating a bimodal grain size, indicating the presence of subgrain rotation recrystallization (SGR) mechanism.

7. Final remarks

The new geological map at 1:10.000 scale is the most detailed, currently available representation of the geo-structural setting of this sector of the Nappe Zone in the Sardinian Variscan Belt. Two juxtaposed units have been distinguished: the lower one represented by the Meana Sardo Unit and the upper one testified by the Barbagia Unit. The contact in between, the Barbagia Thrust (CitationCarosi & Malfatti, Citation1995; CitationCarosi & Pertusati, Citation1990; CitationMontomoli et al., Citation2018; CitationPetroccia et al., Citation2022), is marked by a pluri-m-thick mylonitic zone affecting both units. Detailed mapping, coupled with multi-scale structural observations, allowed the definition of a polyphase evolution (), consisting of three ductile deformation phases developed under a contractional tectonic regime and the fourth one under extensional conditions. The D1 tectonic phase is associated with greenschists-facies paragenesis. In the study area, these structures were obliterated during the D2 phase and, therefore, it is not possible to constrain with precision their kinematics and geometry. According to CitationCarosi et al. (Citation2004), the D1 phase can be related to the early stage of collision. The prominent deformation phase (D2) is responsible for the development of the main foliation ((a)), pervasive at all scales (S2). It becomes more mylonitic moving toward the BT and displays several kinematic indicators with a main top-to-the S-SW normal sense of shear ((a)). Indeed, we detected an increase in the strain gradient moving from the structurally higher parts of the BU and from the structurally lower parts of the MSU toward the BT high-strain zone. The S2 foliation is marked by greenschists-facies mineral assemblage (Wm + Chl). Both BLG and SGR mechanism has been detected (CitationLaw, Citation2014; CitationStipp et al., Citation2002). The presence of MSU and BU highly sheared rocks deformed under non-coaxial mylonitic conditions ((b)) and the emplacing of the MSU and BU with a top-to-the S-SW normal sense of shear during the N-S shortening agree with previous investigations (e.g. CitationCarosi et al., Citation2004; CitationConti et al., Citation2001; CitationMontomoli et al., Citation2018). The D2 phase is linked to the syn-nappe stacking and exhumation of the BU (CitationMontomoli et al., Citation2018). After the crustal thickening and the nappe-stacking evolution, all the previous structural elements are widely deformed by weakly asymmetric to upright D3 folds ((a)). Folds show nearly-vertical axial planes, and therefore no clear transport direction can be associated with this event. However, if we admit that the shortening direction had to be perpendicular to the axial planes and to the trend of fold axes, D3 phase had a similar shortening direction to the D2. Restoring the original pre-Gennargentu Antiform attitude of the D2 fold structures detected in the investigated area, F2 show a SW-S verging anticlinal antiforms and synclinal synforms highlighting a top-to-the S-SW sense of transport. This also agrees with the S-SW sense of shear and the fold polarity detected within the Monte Santa Vittoria area, in the southern limb of the Barbagia Synform (CitationCarosi et al., Citation2004; CitationMontomoli et al., Citation2018). Thus, the presence of the BT mylonitic zone in the different sectors of the Barbagia Synform, with the same structural and kinematic features and the similar shortening direction between the D2 and D3 phase, confirms the post-nappe stacking folded structure. This architecture, in agreement with the model of CitationConti et al. (Citation1999) and CitationCarmignani et al. (Citation1994), emphasizes the presence of large-scale structures acquired in a contractional setting during the late deformation stage of crustal thickening and not during the late orogenic extension. Also, the lack of metamorphic mineral growth during the D3, is consistent with the coupling of BU with MSU under greenschist-facies conditions (D1 and D2) and a subsequent antiformal stacking during the regional horizontal shortening under upper-crustal conditions (D3). During the orogenic extensional phase (D4), folds with sub-horizontal axial planes (F4) are recognizable. This geometry suggests the gravitational collapse of the thickened orogen. Different authors (CitationCarmignani et al., Citation1994; CitationConti et al., Citation1999, Citation2001) highlighted the presence of low- to high-angle normal faults in correspondence of the antiforms limbs displaying an opposite sense of shear and leading to tectonic unroofing of the antiformal hinge zones. However, in the investigated area, no clear BT-parallel extensional structures overprinting BT-related mylonite have been observed.

Figure 4. Synoptic reconstruction of the field and microstructural investigations of the deformation history of the BU and MSU.

Synoptic reconstruction of the deformation history.
Figure 4. Synoptic reconstruction of the field and microstructural investigations of the deformation history of the BU and MSU.

Figure 5. Sketch of the structural evolution derived from the obtained data. (a) Structural features of the pre-, sin- and post-nappe stacking deformation under compressive conditions (D1, D2, D3). (b) Simplified and not to scale structural features of the investigated sector. Far from the mylonitic zone, pre-D2 structures are well-recognizable. The prominence of mylonite and, only in MSU, ultramylonite, characterizes the high-strain zone of the BT. The presence of S-SW dipping of the axial plane of F2 folds due to the D3 folding event has been highlighted.

Summary diagram of deformation history.
Figure 5. Sketch of the structural evolution derived from the obtained data. (a) Structural features of the pre-, sin- and post-nappe stacking deformation under compressive conditions (D1, D2, D3). (b) Simplified and not to scale structural features of the investigated sector. Far from the mylonitic zone, pre-D2 structures are well-recognizable. The prominence of mylonite and, only in MSU, ultramylonite, characterizes the high-strain zone of the BT. The presence of S-SW dipping of the axial plane of F2 folds due to the D3 folding event has been highlighted.

As a whole, the polyphase history recorded in this sector of the Nappe Zone in Sardinia, summarized in this study, highlights the complex tectonic history that occurred during the evolution of an orogenic wedge. In particular, this area records a pre-, syn- to post-nappe stacking evolution developed under contractional conditions and a late-extensional deformation. We also point out the importance of the BT in the architecture of the Sardinian belt, representing a major tectonic boundary characterized by a well-developed non-coaxial deformation and separating two different sectors of the Sardinian orogenic wedge system.

Software

The map has been drawn using QGis 3.16 Hannover, and its final assemblage has been realized with Adobe Illustrator® CC 2018. Structural data have been plotted with the software Stereonet© 10.

Geolocation information

The study area is located in the southwestern sector of the Lago Alto del Flumendosa (Sardinia, Italy). The area is placed between 4417800–1532400 and 4421400–1538400; coordinate system: WGS 1984 / UTM Zone 32 N.

Supplemental material

TJOM_2093660_Supplemental materials

Download PDF (93.2 MB)

Acknowledgements

We want to thank Heike Apps, Giovanni Musumeci and Gaetano Ortolano for their constructive and thorough reviews. We thank Mike J Smith for his efficient editorial work.)

Data Availability Statement

The authors confirm that the data supporting the findings of this study are freely available within the article and its supplementary materials.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Barca, S., Del Rio, M., & Pittau Demelia, P. (1988). New geological and stratigraphical data and discovery of lower Ordovician acritarchs in the San Vito sandstone of the Genn’argiolas unit (Sarrabus, southeastern Sardinia). Rivista Italiana di Paleontologia e Stratigrafa, 94(3), 339–360.
  • Barca, S., & Jäger, H. (1989). New geological and biostratigraphical data on the Silurian in SE Sardinia. Close affinity with Thuringia. Bollettino Della Società Geologica Italiana, 108(4), 565–580.
  • Calvino, F. (1959). Lineamenti strutturali del Sarrabus-Gerrei (Sardegna sud-orientale). Bollettino del Servizio Geologico D’Italia, 81, 489–556.
  • Carmignani, L., Carosi, R., Di Pisa, A., Gattiglio, M., Musumeci, G., Oggiano, G., & Pertusati, P. C. (1994). The hercynian chain in Sardinia (Italy). Geodinamica Acta, 7(1), 31–47. https://doi.org/10.1080/09853111.1994.11105257
  • Carmignani, L., Minzoni, N., Pertusati, P. C., & Gattiglio, M. (1982). Lineamenti geologici principali del Sarcidano-Barbagia di Belvì. In L. Carmignani, T. Cocozza, C. Ghezzo, P. C. Pertusati, & C. A. Ricci (Eds.), Guida alla Geologia del paleozoico sardo (pp. 119–125). Guide geologiche regionali, Società Geologica italiana.
  • Carmignani, L., Oggiano, G., Barca, S., Conti, P., Eltrudis, A., Funedda, A., Pasci, S., & Salvadori, I. (2001). Geologia della Sardegna (Note illustrative della Carta Geologica della Sardegna in scala 1:200.000). Memorie descrittive della Carta Geologica d’Italia, Servizio Geologico Nazionale, Istituto Poligrafico e Zecca dello Stato, Roma.
  • Carmignani, L., Oggiano, G., Funedda, A., Conti, P., & Pasci, S. (2016). The geological map of Sardinia (Italy) at 1:250,000 scale. Journal of Maps, 12(5), 826–835. https://doi.org/10.1080/17445647.2015.1084544
  • Carmignani, L., & Pertusati, P. C. (1977). Analisi strutturale di un segmento della catena ercinica: Il Gerrei (Sardegna sud-orientale). Bollettino Della Società Geologica Italiana, 96, 339–364.
  • Carosi, R. (2004). Carta geologico-strutturale del Monte S. Vittoria (Sarcidano-Barbagia di belvì, Sardegna centrale, italia). In R. Carosi, F. M. Elter, & M. Gattiglio (Eds.), Scala 1:25.000, centrooffset, Siena, 1997. Atti della Società Toscana di Scienze Naturali, serie A, 108 (2002-2003).
  • Carosi, R., D’Addario, E., Mammoliti, E., Montomoli, C., & Simonetti, M. (2016). Geology of the northwestern portion of the Ferriere-Mollieres shear zone, Argentera Massif, Italy. Journal of Maps, 12(sup1), 466–475. https://doi.org/10.1080/17445647.2016.1243491
  • Carosi, R., Iacopini, D., & Montomoli, C. (2004). Asymmetric fold development in the Variscan Nappes of central Sardinia (Italy). Comptes Rendus Geoscience, 336(10), 939–949. https://doi.org/10.1016/j.crte.2004.03.004
  • Carosi, R., Leoni, L., Paolucci, F., Pertusati, P. C., & Trumpy, E. (2010). Deformation and illite crystallinity in metapelitic rocks from the Mandas area, in the Nappe Zone of the Variscan belt of Sardinia. Rendiconti Online Società Geologica Italiana, 11, 393–394.
  • Carosi, R., & Malfatti, G. (1995). Analisi strutturale dell’Unità di Meana Sardo e caratteri della deformazione duttile nel Sarcidano-Barbagia di Seulo (Sardegna centrale, Italia). Atti Della Società Toscana di Scienze Naturali, Serie A, 102, 121–136.
  • Carosi, R., Montomoli, C., Iaccarino, S., Benetti, B., Petroccia, A., & Simonetti, M. (2022). Constraining the timing of evolution of shear zones in two collisional orogens: Fusing structural geology and geochronology. Geosciences, 12(6), 231. https://doi.org/10.3390/geosciences12060231
  • Carosi, R., Montomoli, C., & Iacopini, D. (2002). Le pieghe asimmetriche dell’Unità di Meana Sardo, Sardegna centrale (Italia): evoluzione e meccanismi di piegamento. Atti Della Società Toscana di Scienze Naturali, Serie A, 108, 51–58.
  • Carosi, R., Musumeci, G., & Pertusati, P. C. (1991). Differences in the structural evolution of tectonic units in central-southern Sardinia. Bollettino Della Società Geologica Italiana, 110(3-4), 543–551.
  • Carosi, R., & Palmeri, R. (2002). Orogen-parallel tectonic transport in the Variscan belt of northeastern Sardinia (Italy): implications for the exhumation of medium-pressure metamorphic rocks. Geological Magazine, 139(5), 497–511. https://doi.org/10.1017/S0016756802006763
  • Carosi, R., & Pertusati, P. C. (1990). Evoluzione strutturale delle unità tettoniche erciniche nella Sardegna centro-meridionale. Bollettino Della Società Geologica Italiana, 109, 325–335.
  • Carosi, R., Petroccia, A., Iaccarino, S., Simonetti, M., Langone, A., & Montomoli, C. (2020). Kinematics and timing constraints in a transpressive tectonic regime: The example of the Posada-Asinara shear zone (NE Sardinia, Italy). Geosciences, 10(8), 288. https://doi.org/10.3390/geosciences10080288
  • Casini, L., Cuccuru, S., Maino, M., & Oggiano, G. (2015). Structural map of variscan northern Sardinia (Italy). Journal of Maps, 11(1), 75–84. https://doi.org/10.1080/17445647.2014.936914
  • Casini, L., Cuccuru, S., Maino, M., Oggiano, G., & Tiepolo, M. (2012). Emplacement of the Arzachena Pluton (Corsica-Sardinia batholith) and the geodynamics of incoming Pangaea. Tectonophysics, 544-545, 31–49. https://doi.org/10.1016/j.tecto.2012.03.028
  • Casini, L., Funedda, A., & Oggiano, G. (2010). A balanced foreland-hinterland deformation model for the Southern Variscan belt of Sardinia, Italy. Geological Journal, 45(5–6), 634–649. https://doi.org/10.1002/gj.1208
  • Caso, F., Nerone, S., Petroccia, A., & Bonasera, M. (2021). Geology of the southern gran paradiso massif and lower piedmont zone contact area (middle Ala valley, western Alps, Italy). Journal of Maps, 17(2), 237–246. https://doi.org/10.1080/17445647.2021.1911869
  • Cocco, F., & Funedda, A. (2011). New data on the pre-Middle Ordovician deformation in SE sardinia: A preliminary note. Rendiconti Online Della Società Geologica Italiana, 15, 34–36.
  • Cocco, F., & Funedda, A. (2012). The Variscan basement of the Riu ollastu area (sarrabus, SE sardinia, Italy). Geology of France and Surrounding Areas, 1, 89–91.
  • Cocco, F., & Funedda, A. (2019). The Sardic phase: field evidence of Ordovician tectonics in SE Sardinia, Italy. Geological Magazine, 156(1), 25–38. https://doi.org/10.1017/S0016756817000723
  • Cocco, F., Oggiano, G., Funedda, A., Loi, A., & Casini, L. (2018). Stratigraphic, magmatic and structural features of Ordovician tectonics in Sardinia (Italy): a review. Journal of Iberian Geology, 44(4), 619–639. https://doi.org/10.1007/s41513-018-0075-1
  • Conti, P., Carmignani, L., Cerbai, N., Eltrudis, A., Funedda, A., & Oggiano, G. (1999). From thickening to extension in the Variscan belt - kinematic evidence from Sardinia (Italy). Terra Nova, 11(2/3), 93–99. https://doi.org/10.1046/j.1365-3121.1999.00231.x
  • Conti, P., Carmignani, L., & Funedda, A. (2001). Change of nappe transport direction during the Variscan collisional evolution of central-southern Sardinia (Italy). Tectonophysics, 332(1–2), 255–273. https://doi.org/10.1016/S0040-1951(00)00260-2
  • Conti, P., Funedda, A., & Cerbai, N. (1998). Mylonite development in the Hercynian basement of Sardinia (Italy). Journal of Structural Geology, 20(2/3), 121–133. https://doi.org/10.1016/S0191-8141(97)00091-6
  • Conti, P., & Patta, E. (1998). Large-scale Hercynian W-directed tectonics in southeastern Sardinia (Italy). Geodinamica Acta, 11(5), 217–231. https://www.tandfonline.com/doi/abs/10.1080/09853111.1998.11105321 https://doi.org/10.1080/09853111.1998.11105321
  • Corradini, C., Ferretti, A., & Serpagli, E. (1998). The Silurian and Devonian sequence in SE Sardinia. ECOS VII [special issue]. Giornale di Geologia, 60, 71–74.
  • Costamagna, L., & Barca, S. (2004). Stratigrafia, analisi di facies, paleogeografia e dinquadramento regionale della successione giurassica dell’area dei Tacchi (Sardegna Orientale). Bollettino Della Società Geologica Italiana, 123(3), 477–496.
  • Cruciani, G., Franceschelli, M., Massonne, H.-J., Musumeci, G., & Spano, M. E. (2016). Thermomechanical evolution of the highgrade core in the nappe zone of Variscan Sardinia, Italy: The role of shear deformation and granite emplacement. Journal of Metamorphic Geology, 34(4), 321–342. https://doi.org/10.1111/jmg.12183
  • Cruciani, G., Montomoli, C., Carosi, R., Franceschelli, M., & Puxeddu, M. (2015). Continental collision from two perspectives: A review of Variscan metamorphism and deformation in northern Sardinia. Periodico di Mineralogia, 84(3B), 657–699. https://doi.org/10.2451/2015PM0455
  • Del Moro, A., Di Simplicio, P., Ghezzo, C., Guasparri, G., Rita, F., & Sabatini, G. (1975). Radiometric data and intrusive sequence in the Sardinian batholith. Neues Jahrbuch für Mineralogie Monatshefte, 126, 28–44.
  • Dessau, G., Duchi, G., Moretti, A., & Oggiano, G. (1982). Geologia della zona del Valico di Correboi (Sardegna centro-orientale). rilevamento, tettonica e giacimenti minerari. Bollettino Della Società Geologica Italiana, 101, 223–231.
  • Di Milia, A., & Tongiorgi, M. (1993). Reworked palynomorphs in the Solanas Sandstones (central Sardinia) and their significance for the basin analysis. In L. Carmignani, & F. P. Sassi (Eds.), Contributions to the Geology of Italy with special regards to Paleozoic basement: A volume dedicated to Tommaso Cocozza: IGPC Project no. 276, newsletter, 5 (pp. 461–463.
  • Franceschelli, M., Gattiglio, M., Pannuti, F., & Fadda, S. (1992). Illite crystallinity in pelitic rocks from the external and nappe zones of the Hercynian chain of Sardinia. In L. Carmignani & F. P. Sassi (Eds.), Contributions to the Geology of Italy with special regard to the Paleozoic basements: IGCP Project No. 276, newsletter (pp. 127–135).
  • Funedda, A. (2009). Foreland- and hinterland-verging structures in fold-and-thrust belt: An example from the Variscan foreland of Sardinia. International Journal of Earth Sciences, 98(7), 1625–1642. https://doi.org/10.1007/s00531-008-0327-y
  • Funedda, A., Meloni, M. A., & Loi, A. (2015). Geology of the Variscan basement of the Laconi-Asuni area (central Sardinia, Italy): the core of a regional antiform refolding a tectonic nappe stack. Journal of Maps, 11(1), 146–156. https://doi.org/10.1080/17445647.2014.942396
  • Funedda, A., Naitza, S., Conti, P., Dini, A., Buttau, C., & Tocco, S. (2011). The geological and metallogenic map of the Baccu Locci mine area (Sardinia, Italy). Journal of Maps, 7(1), 103–114. https://doi.org/10.4113/jom.2011.1134
  • Ghosh, P., Bhattacharyya, K., & Parui, C. (2020). Tracking progressive deformation of an orogenic wedge through two successive internal thrusts: Insights from structure, deformation profile, strain, and vorticity of the Main Central Thrust (MCT) and the Pelling-Munsiari thrust (PT), Sikkim Himalayan fold thrust belt. Journal of Structural Geology, 140, 104120. https://doi.org/10.1016/j.jsg.2020.104120
  • Gnoli, M. (1993). Occurrence of middle-late Silurian nautiloids from San Basilio area (Gerrei, SE Sardinia). Bollettino del Museo Della Regione Piemonte, 10(2), 265–269.
  • Grasemann, B., Fritz, H., & Vannay, J. C. (1999). Quantitative kinematic flow analysis from the Main Central Thrust Zone (NW-Himalaya, India): implications for a decelerating strain path and the extrusion of orogenic wedges. Journal of Structural Geology, 21(7), 837–853. https://doi.org/10.1016/S0191-8141(99)00077-2
  • Hudleston, P. (1973). Fold morphology and some geometrical implications of theories of fold development. Tectonophysics, 16(1-2), 1–46. https://doi.org/10.1016/0040-1951(73)90129-7
  • Larson, K. P., Godin, L., Davis, W. J., & Davis, D. W. (2010). Relationships between displacement and distortion in orogens: Linking the Himalayan foreland and hinterland in central Nepal. Geological Society of America Bulletin, 122(7-8), 1116–1134. https://doi.org/10.1130/B30073.1
  • Law, R. D. (2014). Deformation thermometry based on quartz c-axis fabrics and recrystallization microstructures: A review. Journal of Structural Geology, 66, 129–161. https://doi.org/10.1016/j.jsg.2014.05.023
  • Loi, A. (1993). Sedimentological-petrographical study and paleogeographical approach of the Upper Ordovician of the central southern Sardinia. Eur. J. Mineral. ‘Plinius’, 9, 81–86.
  • Loi, A., & Dabard, M. P. (1997). Zircon typology and geochemistry in the palaeogeographic reconstruction of the late Ordovician of Sardinia (Italy). Sedimentary Geology, 112(3–4), 263–279. https://doi.org/10.1016/S0037-0738(97)00038-9
  • Matte, P. (2001). The Variscan collage and orogeny (480–290 Ma) and the tectonic definition of the Armorica microplate: A review. Terra Nova, 13(2), 122–128. https://doi.org/10.1046/j.1365-3121.2001.00327.x
  • Minzoni, N. (1975). La serie delle successioni paleozoiche a sud del Gennargentu. Bollettino Della Società Geologica Italiana, 94, 347–392.
  • Montomoli, C. (2003). Zone di taglio fragili-duttili nel basamento varisico metamorfico di basso grado della Nurra meridionale (Sardegna nord-occidentale). Atti Della Società Toscana di Scienze Naturali, Serie A, 108, 23–29.
  • Montomoli, C., Iaccarino, S., Simonetti, M., Lezzerini, M., & Carosi, R. (2018). Structural setting, kinematics and metamorphism in a km-scale shear zone in the inner nappes of Sardinia (Italy). Italian Journal of Geosciences, 137(2), 294–310. https://doi.org/10.3301/IJG.2018.16
  • Musumeci, G. (1991). Displacement calculation in a ductile shear zone: Monte Grighini shear zone (central-western Sardinia). Bollettino Della Società Geologica Italiana, 110, 771–777.
  • Oggiano, G., Gaggero, L., Funedda, A., Buzzi, L., & Tiepolo, M. (2010). Multiple early Paleozoic volcanic events at the northern Gondwana margin: U–Pb age evidence from the Southern Variscan branch (Sardinia, Italy). Gondwana Research, 17(1), 44–58. https://doi.org/10.1016/j.gr.2009.06.001
  • Papeschi, S., Iaccarino, S., & Montomoli, C. (2020). Underthrusting and exhumation of continent-derived units within orogenic wedge: An example from the Northern Apennines (Italy). Journal of Maps, 16(2), 638–650. https://doi.org/10.1080/17445647.2020.1795736
  • Passchier, C. W., & Trouw, R. A. J. (2005). Microtectonics. Springer, Berlin, Heidelberg. 366. https://doi.org/10.1007/3-540-29359-0
  • Pavanetto, P., Funedda, A., Northrup, C. J., Schmitz, M., Crowley, J., & Loi, A. (2012). Structure and U-Pb zircon geochronology in the Variscan foreland of SW Sardinia, Italy. Geological Journal, 47(4), 426–445. https://doi.org/10.1002/gj.1350
  • Pertusati, P., Sarria, E., Cherchi, G. P., Carmignani, L., Barca, S., Benedetti, M., Chighine, G., Cincotti, F., Oggiano, G., Ulzega, A., Orrù, P., & Pintus, C. (2002). Note illustrative della Carta Geologica d’Italia alla scala 1:50.000 - foglio 541 Jerzu, pp. 143, Servizio Geologico D’Italia.
  • Petroccia, A., Carosi, R., Montomoli, C., Iaccarino, S., & Vitale Brovarone, A. (2022). Deformation and temperature variation along thrust-sense shear zones in the hinterland-foreland transition zone of collisional settings: A case study from the Barbagia Thrust (Sardinia, Italy). Journal of Structural Geology. https://doi.org/10.1016/j.jsg.2022.104640
  • Piazolo, S., & Passchier, C. W. (2002). Controls on lineation development in low to medium grade shear zones: A study from the Cap de Creus peninsula, NE Spain. Journal of Structural Geology, 24(1), 25–44. https://doi.org/10.1016/S0191-8141(01)00045-1
  • Ramsay, J. G. (1967). Folding and fracturing of rocks. Mcgraw-Hill.
  • Ring, U., & Brandon, M. T. (1999). Ductile deformation and mass loss in the Franciscan subduction complex: Implications for exhumation processes in accretionary wedges. In U. Ring, M. T. Brandon, G. S. Lister, & S. D. Willett (Eds.), Exhumation processes: Normal faulting, ductile flow and erosion. Special publications of the Geological Society London 154 (pp. 55–86. https://doi.org/10.1144/gsl.sp.1999.154.01.03
  • Sarkarinejad, K., & Ghanbarian, M. A. (2014). The Zagros hinterland fold-and-thrust belt in-sequence thrusting, Iran. Journal of Asian Earth Sciences, 85, 66–79. https://doi.org/10.1016/j.jseaes.2014.01.017
  • Schneider, J., Corsini, M., Reverso Peila, A., & Lardeaux, J.-M. (2014). Thermal and mechanical evolution of an orogenic wedge during Variscan collision: An example in the Maures-Tanneron Massif (SE France). Geological Society, London, Special Publications, 405(1), 313–331. https://doi.org/10.1144/SP405.4
  • Stipp, M., Stünitz, H., Heilbron, M., & Schmid, D. W. (2002). The eastern Tonale fault zone: A ‘natural laboratory’ for crystal plastic deformation of quartz over a temperature range from 250 to 700°C. Journal of Structural Geology, 24(12), 1861–1884. https://doi.org/10.1016/S0191-8141(02)00035-4
  • Tarquini, S., Isola, I., Favalli, M., Mazzarini, F., Bisson, M., Pareschi, M. T., & Boschi, E. (2007). TINITALY/01: A new Triangular Irregular Network of Italy. Annals of Geophysics, 50, 407–425. https://doi.org/10.4401/ag-4424
  • Thigpen, J. R., Law, R. D., Lloyd, G. E., Brown, S. J., & Cook, B. (2010). Deformation temperatures, vorticity of flow and strain symmetry in the Loch Eriboll mylonites, NW Scotland: Implications for the kinematic and structural evolution of the northernmost Moine Thrust zone. Geological Society, London, Special Publications, 335(1), 623–662. https://doi.org/10.1144/SP335.26
  • Thigpen, J. R., Law, R. D., Loehn, C., Strachan, R. A., Tracy, R., Lloyd, G., Roth, B., & Brown, S. (2013). Thermal structure and tectonic evolution of the Scandian orogenic wedge, Scottish Caledonides: Integrating geothermometry, deformation temperatures and conceptual kinematic-thermal models. Journal of Metamorphic Geology, 31(8), 813–842. https://doi.org/10.1111/jmg.12046
  • Vai, G. B., & Cocozza, T. (1974). Il “postgotlandiano” sardo, unità sinorogenica ercinica. Bollettino Della Società Geologica Italiana, 93, 61–72.
  • Whitney, D., & Evans, B. (2010). Abbreviations for names of rock-forming minerals. American Mineralogist, 95(1), 185–187. https://doi.org/10.2138/am.2010.3371