Hostname: page-component-7f64f4797f-d87pz Total loading time: 0 Render date: 2025-11-12T03:38:16.165Z Has data issue: false hasContentIssue false

The Lower Paleozoic to Permian evolution of the Dora-Maira Massif basement (Western Alps) revealed by zircon geochronology

Published online by Cambridge University Press:  03 November 2025

Davide Dana*
Affiliation:
Dipartimento di Scienze Della Terra, Università di Torino, Turin, Italy
Francesco De Cesari
Affiliation:
Dipartimento di Scienze Della Terra, Università di Torino, Turin, Italy
Chiara Montomoli
Affiliation:
Dipartimento di Scienze Della Terra, Università di Torino, Turin, Italy
Salvatore Iaccarino
Affiliation:
Dipartimento di Scienze Della Terra, Università di Torino, Turin, Italy
Alberto Corno
Affiliation:
Dipartimento di Scienze Della Terra, Università di Torino, Turin, Italy
Daniela Rubatto
Affiliation:
Institut für Geology, Bern University, Bern, Switzerland
Rodolfo Carosi
Affiliation:
Dipartimento di Scienze Della Terra, Università di Torino, Turin, Italy
*
Corresponding author: Davide Dana; Email: davide.dana@unito.it
Rights & Permissions [Opens in a new window]

Abstract

The Dora-Maira Massif offers a unique window into the pre-Alpine and Alpine tectonic history of continental crust involved in the collision that formed the Western Alps. Whereas high-pressure Alpine metamorphism has strongly affected the rocks, zircon crystals still retain age domains from earlier magmatic events, offering a reliable tool for reconstructing the evolution of the continental basement. Combining new U–Pb LA-ICP-MS zircon geochronology with whole-rock geochemistry, this study unveils a more complex magmatic and tectonic evolution than previously recognized. New findings reveal that the Dora-Maira basement preserves three distinct magmatic cycles: Ediacaran–Early Cambrian (c. 550–530 Ma), Ordovician–Silurian (c. 460–440 Ma) and Permian (c. 280–260 Ma). These results challenge the overestimated presence of Permian intrusions in the Massif, instead highlighting a significant Lower Paleozoic magmatic component that was previously overlooked. We document the oldest magmatic suite recognized to date in the area (540.0 ± 5.4 and 548.0 ± 5.5 Ma), which sheds light on the Cambrian-to-Precambrian tectonic history of the northern margin of Gondwana. This contribution refines the geodynamic framework of the Dora-Maira Massif. It demonstrates that its polycyclic basement was assembled through successive crust-forming events during the Ediacaran–Early Paleozoic and subsequently reworked during the Variscan and Alpine orogenies. This underlines the key role of Lower Paleozoic tectono-magmatic processes in shaping the continental crust of the Western Alps.

Information

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1. Introduction

Continental basements commonly preserve evidence of crustal recycling and thus provide a valuable archive for deciphering the tectonic events that shaped the continental crust (Cawood et al., Reference Cawood, Hawkesworth and Dhuime2013; Condie, Reference Condie2014; Spencer et al., Reference Spencer, Roberts and Santosh2017). Such records are often difficult to retrieve from units located in the axial zone of orogenic belts, where intense deformation and metamorphism may conceal or reset earlier geological signatures. The Western Alps represent a natural laboratory for investigating the architecture and evolution of recycled continental crust. Over the past five decades, extensive geochronological work has been carried out on the continent-derived units of this orogenic belt, yielding both crystallization ages of magmatic protoliths and depositional ages of (meta)sedimentary sequences (e.g. Hunziker et al., Reference Hunziker, Desmons and Hurford1992; Neubauer et al., Reference Neubauer, Chang, Dong, Genser and Liu2025). These data have revealed a long-lasting and complex pre-Alpine history for the crystalline basements, now incorporated into the Alpine orogen. Nevertheless, despite these efforts, the pre-Alpine evolution of individual basement units remains only partially constrained, especially in regions affected by strong Alpine metamorphic (up to high- and ultra-high-pressure conditions – HP and UHP) and polyphase deformation overprint.

The Dora-Maira Massif, part of the internal Western Alps, comprises HP to UHP metamorphic rocks, including widespread orthogneisses interpreted as the products of pre-Alpine granitic intrusions (e.g. Compagnoni et al., Reference Compagnoni, Rolfo, Groppo, Hirajima and Turello2012). These meta-intrusive rocks offer a unique opportunity to investigate the early magmatic and tectonic evolution of the continental crust involved in the orogen. We approach this challenge using zircon U–Pb geochronology, which is a powerful tool for reconstructing the timing of magmatic and metamorphic events in such polycyclic basement units, as zircon can retain age domains corresponding to multiple geological processes (e.g. Gebauer, Reference Gebauer and von Raumer J.F.1993; Schaltegger & Gebauer, Reference Schaltegger and Gebauer1999; Rubatto, Reference Rubatto2017).

In this study, we present new U–Pb LA-ICP-MS zircon geochronology and whole-rock geochemical data from orthogneisses exposed in the northern-central Dora-Maira Massif (between the Pellice and Chisone Valleys). These data, acquired within the framework of the Geological Map of Italy (CARG Project, sheet 172 ‘Pinerolo’, scale 1:50,000), reveal the presence of distinct magmatic suites, previously differentiated (and in several cases misattributed) based solely on geochemistry. Our results provide new age constraints and bring to light previously unrecognized magmatic cycles, helping to clarify the history of this portion of the pre-Alpine basement. These findings are placed within a broader geodynamic and paleogeographic framework, with particular emphasis on the role played by Lower Paleozoic events in shaping the Dora-Maira basement.

2. Geological setting

2.a. Continental crust in the Western Alps

In the Western Alps, several tectonic units contain slices of continental crust sampling three main continental blocks involved in the Alpine collision (Dal Piaz et al., Reference Dal Piaz, Bistacchi and Massironi2003; Schmid et al., Reference Schmid, Fügenschuh, Kissling and Schuster2004, Reference Schmid, Kissling, Dichl, van Hinsbergen and Molli2017; Dal Piaz, Reference Dal Piaz2010; Handy et al., Reference Handy, Schmid, Bousquet, Kissling and Bernoulli2010, Reference Handy, Schmid, Paffrath and Friederich2021; Agard & Handy, Reference Agard and Handy2021; Brunsmann et al., Reference Brunsmann, Rosenberg and Bellahsen2024). These slices are derived from:

(i) The European passive margin, exposed in the Helvetic-Dauphinois domain, where pre-Mesozoic basement rocks are exposed in the External Crystalline Massifs (Fig. 1), associated with a Meso-Cenozoic cover. In these massifs, Alpine metamorphism reaches only sub-greenschist to greenschist facies conditions (e.g. Bousquet et al., Reference Bousquet, Oberhänsli, Goffé, Wiederker, Koller, Schmid, Schuster, Engi, Berger and Martinotti2008; Bellanger et al., Reference Bellanger, Bellahsen, Jolivet, Baudin, Augier and Boutoux2014), and Variscan tectonics is largely preserved (Simonetti et al., Reference Simonetti, Carosi, Montomoli, Langone, D’Addario and Mammoliti2018; Reference Simonetti, Carosi, Montomoli, Cottle and Law2020, Reference Simonetti, Carosi, Montomoli, Law and Cottle2021).

Figure 1. Tectonic sketch of the Western Alps (modified from Dana, Reference Dana2024). The geolocation of the map is shown in the upper left insert. Magmatic protoliths with ages around 500–525 Ma are reported (1: Liati et al., Reference Liati, Gervers, Froitzheim and Fanning2001; 2: Sartori et al., Reference Sartori, Gouffon and Marthaler2006; 3: Scheiber et al., Reference Scheiber, Berndt, Mezger and Pfiffner2014; 4: Bussy et al., Reference Bussy, Derron, Jacquod, Sartori and Thélin1996; 5: Guillot et al., Reference Guillot, Liégeois and Fabre1991; 6: Bertrand et al., Reference Bertrand, Pidgeon, Leterrier, Guillot, Gasquet and Gattiglio2000b, 7: Bertrand et al., Reference Bertrand, Guillot and Leterrier2000a; 8: Ménot et al., Reference Ménot, Peucat, Scarenzi and Piboule1988; 9: Thiéblemont et al., Reference Thiéblemont, Jacob, Lach, Guerrot and Leguérinel2023; 10: Balestro et al., Reference Balestro, Festa, Cadoppi, Groppo and Roà2022; 11: Filippi et al., Reference Filippi, Jouffray, Lardeaux, Tiepolo and Spalla2024). The tectonic subdivision of the Western Alps is largely based on Schmid et al. (Reference Schmid, Fügenschuh, Kissling and Schuster2004, Reference Schmid, Kissling, Dichl, van Hinsbergen and Molli2017) and Gouffon et al. Reference Gouffon, Bernoulli, Dall’agnolo, Fantoni, Jordan, Madritsch, Mosar, Picotti, Pfiffner, Schenker, Schlunegger and Schmid(2024b). The red box highlights the position of Fig. 2. Abbreviations of the tectonic units cited in the text: AM = Ambin Massif, DM = Dora-Maira, GP = Gran Paradiso, MR = Monte Rosa, R = Ruitor, SM = Siviez-Mischabel, ZH = Zone Houillère.

(ii) The Briançonnais microcontinent (Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022), whose remnants are exposed in the Middle Penninic domain, separated from the Helvetic-Dauphinois by the Penninic Frontal Thrust (Ceriani et al., Reference Ceriani, Fügenschuh and Schmid2001).

(iii) The Adria extensional allochthons and the Adriatic continental plate. The extensional allochthons (Sesia and Dent Blanche Units; Fig. 1) underwent hyper-extension during the Latest Triassic to Early Jurassic rifting (Mohn et al., Reference Mohn, Manatschal, Müntener, Beltrando and Masini2010; Manzotti et al., Reference Manzotti, Ballèvre, Zucali, Robyr and Engi2014) and nowadays are exposed on top of the nappe pile in the Salassic domain (Marthaler et al., Reference Kroner and Romer2020; Gouffon et al., Reference Gouffon, Bernoulli, Dall’agnolo, Fantoni, Jordan, Madritsch, Mosar, Pfiffner, Picotti and Schenker2024a). These units experienced blueschist to eclogitic Alpine metamorphism (Compagnoni, Reference Compagnoni1977; Beltrando et al., Reference Beltrando, Compagnoni and Lombardo2010; Manzotti et al., Reference Manzotti, Ballèvre, Zucali, Robyr and Engi2014; Reference Manzotti, Rubatto, Zucali, Korh, Cenki-Tok, Ballevre and Engi2018). In contrast, an almost complete crustal section of the former Adriatic continental plate is exposed in the South Alpine domain (Fig. 1), where Alpine metamorphism is absent or weak and mostly related to deformation zones (Henk et al., Reference Henk, Franz, Teufel and Oncken1997; Handy et al., Reference Handy, Franz, Heller, Janott and Zurbriggen1999; Schmid et al., Reference Schmid, Kissling, Dichl, van Hinsbergen and Molli2017).

The Briançonnais microcontinent, which was separated from the European margin by the Valaisan basin, likely formed during the Cretaceous (Frisch, Reference Frisch1979; Stampfli, Reference Stampfli1993; De Broucker et al., Reference De Broucker, Siméon, Stampfli, Thiéblemont, Lach and Marthaler2021; Boschetti et al., Reference Boschetti, Mouthereau, Schwartz, Rolland, Bernet, Balvay and Lahfid2025). The Briançonnais-derived units include (i) the Grand Saint Bernard nappe system, comprising detached cover units in the French-Italian Alps and basement units located west of the ophiolite-bearing units (e.g. Desmons, Reference Desmons1992; Sartori et al., Reference Sartori, Gouffon and Marthaler2006; Dumont et al., Reference Dumont, Schwartz, Guillot, Malusà, Jouvent, Monié and Verly2022; Pantet et al., Reference Pantet, Epard, Masson, Baumgartner-Mora, Baumgartner and Baumgartner2023; Dana et al., Reference Dana, Iaccarino, Schmid, Petroccia and Michard2023; Fig. 1), and (ii) the basement-dominated Internal Crystalline Massifs, which lack significant Mesozoic covers (Gasco et al., Reference Gasco, Gattiglio and Borghi2011; Reference Gasco, Gattiglio and Borghi2013; Ballèvre et al., Reference Ballèvre, Camonin, Manzotti and Poujol2020). The paleogeographic attribution of the Internal Crystalline Massifs (Dora-Maira, Gran Paradiso and Monte Rosa, Fig. 1) has been debated (e.g. Froitzheim et al., Reference Froitzheim2001, proposed a sub-Penninic origin, at least for the Monte Rosa Massif), but they are generally considered part of the pre-Triassic basement of the Briançonnais microcontinent (Michard, Reference Michard1967; Ballèvre et al., Reference Ballèvre, Manzotti and Dal Piaz2018, Reference Ballèvre, Camonin, Manzotti and Poujol2020; Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022). The Briançonnais-derived units record peak Alpine metamorphism, varying from eclogite and blueschist facies conditions in the Internal Crystalline Massifs to blueschist, greenschist and sub-greenschist facies in more external Classic Briançonnais units (Bousquet et al., Reference Bousquet, Oberhänsli, Goffé, Wiederker, Koller, Schmid, Schuster, Engi, Berger and Martinotti2008; Beltrando et al., Reference Beltrando, Compagnoni and Lombardo2010). UHP conditions were reached in the Brossasco-Isasca and Rocca Solei (Southern Dora-Maira) and in the Chasteiran, Muret and Serre (Northern Dora-Maira) Units (Chopin, Reference Chopin1984; Manzotti et al., Reference Manzotti, Schiavi, Nosenzo, Pitra and Ballèvre2022, Reference Manzotti, Schiavi, Ballèvre and Nosenzo2025b; Groppo et al., Reference Groppo, Ferrando, Tursi and Rolfo2025).

2.b. Overview of the pre-Alpine evolution of the Western Alps

Intrusive rocks dated between c. 530 and 500 Ma represent the oldest magmatic rocks in the Western Alps (Fig. 1; e.g. Bertrand et al., Reference Bertrand, Pidgeon, Leterrier, Guillot, Gasquet and Gattiglio2000b; Thiéblemont et al., Reference Thiéblemont, Jacob, Lach, Guerrot and Leguérinel2023). Along with large outcrops of metabasic rocks of similar ages (540–500 Ma) found in the Lepontine and Austroalpine units, these rocks have been interpreted as allochthonous fragments of the Cadomian basement (von Raumer et al., Reference Von Raumer, Stampfli, Arenas and Sánchez Martínez2015; Siegesmund et al., Reference Siegesmund, Oriolo, Schulz, Heinrichs, Basei and Lammerer2021, Reference Siegesmund, Oriolo, Broge, Hueck, Lammerer, Basei and Schulz2023; Thiéblemont et al., Reference Thiéblemont, Jacob, Lach, Guerrot and Leguérinel2023). They could represent pieces of former Ediacaran to Cambrian magmatic arcs, which were accreted to the northern, active margin of Gondwana, an active continental margin at the time, involved with the subduction of the Iapetus Ocean (Linnemann et al., Reference Linnemann, Gerdes, Drost and Buschmann2007; Von Raumer & Stampfli, Reference Von Raumer and Stampfli2008; Kroner & Romer, Reference Kroner and Romer2013; Stampfli et al., Reference Stampfli, Hochard, Vérard and Wilhem2013; Couzinié et al., Reference Couzinié, Laurent, Poujol, Mintrone, Chelle-Michou, Moyen and Marko2017).

The future Alpine basements, located along the northern margin of Gondwana (von Raumer & Neubauer, Reference Von Raumer, Neubauer, von Raumer and Neubauer1993), underwent an event of bimodal magmatism during the Middle to Late Ordovician. The tectonic setting of this magmatism remains uncertain (e.g. Kroner & Romer, Reference Kroner and Romer2013; Von Raumer et al., Reference Von Raumer, Bussy, Schaltegger, Schulz and Stampfli2013; Bergomi et al., Reference Bergomi, Dal Piaz, Malusà, Monopoli and Tunesi2017). Some studies have proposed that short-lived (470–450 Ma) bimodal magmatism indicates a transtensional to extensional setting (Ballèvre et al., Reference Ballèvre, Fourcade, Capdevila, Peucat, Cocherie and Fanning2012; Von Raumer et al., Reference Von Raumer, Bussy, Schaltegger, Schulz and Stampfli2013; Villaseca et al., Reference Villaseca, Martínez, Orejana, Andersen and Belousova2016), whereas other models favour an Alaskan-type orogenic setting, known as Cenerian orogeny (Zurbriggen, Reference Zurbriggen2017; Siegesmund et al., Reference Siegesmund, Oriolo, Schulz, Heinrichs, Basei and Lammerer2021, Reference Siegesmund, Oriolo, Broge, Hueck, Lammerer, Basei and Schulz2023; Finger & Riegler, Reference Finger and Riegler2023). Ordovician orthogneisses in the pre-Variscan basement of the Central and Western Alps are closely associated with eclogitic mafic boudins and massive to banded amphibolite sills (Desmons, Reference Desmons1992; Poller, Reference Poller1997; Gaggero et al., Reference Gaggero, Cortesogno and Bertrand2004; Gauthiez et al., Reference Gauthiez, Bussy, Ulianov, Gouffon and Sartori2011). These bimodal magmatic features suggest that all these units were part of the same basement in the northern Gondwana margin during the Middle to Late Ordovician period (Bergomi et al., Reference Bergomi, Dal Piaz, Malusà, Monopoli and Tunesi2017).

The northern margin of Gondwana was later dismembered into different units during the Variscan orogeny (Late Devonian to Early Permian). The Variscan belt was assembled along the southern margin of Laurussia, including the Avalonia and Armorica microcontinents (Matte et al., Reference Matte, Maluski, Rajlich and Franke1990; Stampfli & Borel, Reference Stampfli and Borel2002). Several pulses of magmatism are related to the Variscan belt: an early pre-collisional pulse of medium- to high-K calc-alkaline melts (Late Devonian to Early Carboniferous, 380–345 Ma), followed by synorogenic high-K calc-alkaline to shoshonitic magmatism (345–330 Ma), and later post-collisional plutons intruded between 330 and 300 Ma (Bonin, Reference Bonin1988; Reference Bonin, Brändlein, Bussy, Desmons, Eggenberger, Finger, Graf, Marro, Mercolli, Oberhänsli, Ploquin, von Quadt, von Raumer, Schaltegger, Steyrer, Visonà and Vivier1993; Schaltegger, Reference Schaltegger1997), associated with Barrovian metamorphism and anatectic migmatites (Žák et al., Reference Žák, Verner, Janoušek, Holub, Kachlík, Finger and Trubač2014; Ballèvre et al., Reference Ballèvre, Manzotti and Dal Piaz2018). Evidence of this Variscan magmatic evolution in the Alps is preserved in the External Crystalline Massif and partially in the Briançonnais-derived basements, where the Grand Nomenon pluton provides a Late Devonian magmatic emplacement age (Bergomi et al., Reference Bergomi, Dal Piaz, Malusà, Monopoli and Tunesi2017; Bertrand et al., Reference Bertrand, Pidgeon, Leterrier, Guillot, Gasquet and Gattiglio2000b).

From the Late Carboniferous to the Early Permian, wrench tectonics, post-Variscan transtension and lithospheric thinning were active (Schuster & Stüwe, Reference Schuster and Stüwe2008; Von Raumer et al., Reference Von Raumer, Bussy, Schaltegger, Schulz and Stampfli2013; Spalla et al., Reference Spalla, Zanoni, Marotta, Rebay, Roda, Zucali and Gosso2014; Manzotti et al., Reference Manzotti, Poujol and Ballèvre2015). This transition caused the fragmentation of the Variscan belt through several dextral shear zones (Corsini & Rolland, Reference Corsini and Rolland2009; Guillot et al., Reference Guillot, di Paola, Ménot, Ledru, Spalla, Gosso and Schwartz2009; Carosi et al., Reference Carosi, Montomoli, Tiepolo and Frassi2012; Simonetti et al., Reference Simonetti, Carosi, Montomoli, Cottle and Law2020, Reference Simonetti, Carosi, Montomoli, Law and Cottle2021), with faulting and lithospheric extension associated with igneous intrusions and the opening of intramontane coal-bearing basins (Von Raumer, Reference Von Raumer1998; Ballèvre et al., Reference Ballèvre, Manzotti and Dal Piaz2018).

During the Early Permian, mafic melts intruded at depth in the future Alpine basement (Dal Piaz, Reference Dal Piaz1993; Marotta & Spalla, Reference Marotta and Spalla2007; Manzotti et al., Reference Manzotti, Ballèvre and Dal Piaz2017), associated with bimodal volcanism and granite intrusions in the upper crust (Dal Piaz & Martin, Reference Dal Piaz and Martin1998; Schaltegger & Gebauer, Reference Schaltegger and Gebauer1999; Bergomi et al., Reference Bergomi, Dal Piaz, Malusà, Monopoli and Tunesi2017). This widespread Early Permian magmatism is generally interpreted as a response to lithospheric thinning and strike-slip tectonics, followed by regional extension and K-rich volcanism in the Late Permian (Schaltegger & Brack, Reference Schaltegger and Brack2007; Dallagiovanna et al., Reference Dallagiovanna, Gaggero, Maino, Seno and Tiepolo2009).

The scarce Upper Carboniferous magmatism and the occurrence of Lower Permian intrusions and volcanism suggest that the Briançonnais units derive from the Gondwana foreland, in a more external position within the Variscan belt with respect to the basement of the External Crystalline Massifs (Bergomi et al., Reference Bergomi, Dal Piaz, Malusà, Monopoli and Tunesi2017; Ballèvre et al., Reference Ballèvre, Manzotti and Dal Piaz2018). According to Ballèvre et al. (Reference Ballèvre, Manzotti and Dal Piaz2018), the Permian event likely influenced the subsequent Triassic thermal subsidence, leading to the deposition of thick carbonate platforms in the Briançonnais s.l. domain (Mégard-Galli & Baud, Reference Mégard-Galli and Baud1977; Dumont et al., Reference Dumont, Schwartz, Guillot, Malusà, Jouvent, Monié and Verly2022; Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022). In contrast, in the Helvetic Zone, where no such thermal subsidence occurred, Triassic sediments were deposited in the ‘Germanic Trias’ facies, with thicknesses not exceeding a few metres (von Gümbel, Reference von Gümbel1891; Hauschke & Wilde, Reference Hauschke and Wilde1999; Ziegler, Reference Ziegler, Raumer and Neubauer1993).

2.c. The Dora-Maira Massif

The Dora-Maira Massif consists of tectonic units largely made of pre-Alpine basement and partly detached Permo-Mesozoic covers (Novarese Reference Novarese1895; Vialon, Reference Vialon1966; Michard, Reference Michard1967; Henry, Reference Henry1990; Sandrone et al., Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993). The main tectonic units of the Dora-Maira Massif (Chopin et al., Reference Chopin, Henry and Michard1991; Michard et al., Reference Michard, Henry and Chopin1993, Reference Michard, Henry, Chopin, Coleman and Wang1995, Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022; Bonnet et al., Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022; Manzotti et al., Reference Manzotti, Schiavi, Nosenzo, Pitra and Ballèvre2022Reference Manzotti, Millonig, Gerdes, Whitehouse, Jeon, Poujol and Ballèvre2025a, Reference Manzotti, Schiavi, Ballèvre and Nosenzo2025b; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024) from bottom to top are (Fig. 2):

Figure 2. Tectonic sketch of the Dora-Maira Massif (compiled from Henry, Reference Henry1990; Sandrone et al., Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993; Piana et al., Reference Piana, Fioraso, Irace, Mosca, d’Atri, Barale, Falletti, Monegato, Morelli, Tallone and Vigna2017; Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024; Groppo et al., Reference Groppo, Ferrando, Tursi and Rolfo2025 and our new observations). Magmatic photolith ages from the literature are given (a: Bussy & Cadoppi, Reference Bussy and Cadoppi1996; b: Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024; c: Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022; d: Paquette et al., Reference Paquette, Montel and Chopin1999; e: Chen et al., Reference Chen, Zhou, Zheng and Schertl2017; f: Balestro et al., Reference Balestro, Festa, Cadoppi, Groppo and Roà2022; g: Gebauer et al., Reference Gebauer, Schertl, Brix and Schreyer1997). Tectonic units abbreviations: VG = Val Grana, VMS = Val Maira-Sampeyre, CLSZ = Cima Lubin Shear Zone, DU = Dronero, VSZ = Valmala Shear Zone, RU = Ricordone, SE = Serre, MU = Muret, RS = Rocca Solei, GU = Grimbassa, CU = Chasteiran, SC = San Chiaffredo, BIU = Brossasco-Isasca, PSU = Pinerolo–Sanfront, MV = Monviso, OR = Orsiera-Rocciavrè, SU = Susa, LA = Lanzo. The yellow stars show the location of the studied samples; the black square is the area covered by the Geological Map of Italy sheet 172 ‘Pinerolo’.

(i) The Pinerolo–Sanfront Unit (‘Complesso Grafitico del Pinerolese’ of Franchi & Novarese, Reference Franchi and Novarese1895 and Borghi et al., Reference Borghi, Cadoppi, Porro, Sacchi and Sandrone1984, and ‘Ensemble de Pinerolo’ of Vialon, Reference Vialon1966) mainly consists of monocyclic metasediments (metaconglomerates, meta-arkoses and metapelites characterized by the presence of graphite) of Permo-Carboniferous age (Manzotti et al., Reference Manzotti, Ballèvre and Poujol2016; Carosi et al., Reference Carosi, Montomoli, Iaccarino, Dana, Corno, De Cesari and Spina2025) and Upper Paleozoic magmatic bodies (Fig. 2; Bussy & Cadoppi, Reference Bussy and Cadoppi1996). Alpine metamorphic conditions are generally referred to as the blueschist or eclogite-blueschist transition facies (Borghi et al., Reference Borghi, Cadoppi, Porro and Sacchi1985; Groppo et al., Reference Groppo, Ferrando, Gilio, Botta, Nosenzo, Balestro, Festa and Rolfo2019).

(ii) The ‘Basement Complex’ (Chopin et al., Reference Chopin, Henry and Michard1991; Michard et al., Reference Michard, Henry and Chopin1993, Reference Michard, Henry, Chopin, Coleman and Wang1995; Bonnet et al., Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022), a set of different tectono-metamorphic units with comparable lithologies and metamorphic Alpine evolution at eclogitic facies (in some cases up to UHP conditions, Chopin, Reference Chopin1984; Manzotti et al., Reference Manzotti, Schiavi, Nosenzo, Pitra and Ballèvre2022, Reference Manzotti, Schiavi, Ballèvre and Nosenzo2025b; Groppo et al., Reference Groppo, Ferrando, Tursi and Rolfo2025). These units have been distinguished primarily based on ‘peak’ P-T estimates (Compagnoni et al., Reference Compagnoni, Rolfo, Groppo, Hirajima and Turello2012; Groppo et al., Reference Groppo, Ferrando, Gilio, Botta, Nosenzo, Balestro, Festa and Rolfo2019, Reference Groppo, Ferrando, Tursi and Rolfo2025) and consist of a pre-Alpine basement with rarely preserved Permo-Mesozoic cover (Fig. 2; Franchi, Reference Franchi1898; Sacchi et al., Reference Sacchi, Balestro, Cadoppi, Carraro, Delle Piane, Di Martino, Enrietti, Gallarà, Gattiglio, Martinotti and Perello2004; Caron, Reference Caron1977; Sacchi et al., Reference Sacchi, Balestro, Cadoppi, Carraro, Delle Piane, Di Martino, Enrietti, Gallarà, Gattiglio, Martinotti and Perello2004; Marthaler et al., Reference Marthaler, Fudral, Deville and Rampnoux1986; Gasco et al., Reference Gasco, Gattiglio and Borghi2013; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024). The pre-Alpine basement is mainly made of pre-Carboniferous metasediments (garnet-chloritoid-bearing micaschist, paragneiss and marble; Vialon, Reference Vialon1966; Cadoppi, Reference Cadoppi1990; Wheeler, Reference Wheeler1991) associated with several metabasite and orthogneiss bodies. These meta-intrusive rocks have been regarded as late Variscan (Hunziker et al., Reference Hunziker, Desmons and Hurford1992; Sandrone et al., Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993; Bussy & Cadoppi, Reference Bussy and Cadoppi1996), although few studies have addressed their dating (e.g. Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Paquette et al., Reference Paquette, Montel and Chopin1999; Chen et al., Reference Chen, Schertl, Zheng, Huang, Zhou and Gong2016, Reference Chen, Zhou, Zheng and Schertl2017; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022, Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024).

(iii) The uppermost Dronero Unit (Michard et al., Reference Michard, Henry and Chopin1993), cropping out in the southern portion of the Massif (Fig. 2), consists of polycyclic garnet ± chloritoid-bearing micaschist, metagranitoid and metabasite bodies (Michard & Vialon, Reference Michard and Vialon1966; Michard, Reference Michard1967; Henry, Reference Henry1990), with monocyclic ankerite-chloritoid-bearing micaschist attributed to the Permian-Carboniferous (Michard, Reference Michard1967; Henry, Reference Henry1990). The Dronero Unit reached blueschist facies Alpine metamorphic conditions (Groppo et al., Reference Groppo, Ferrando, Gilio, Botta, Nosenzo, Balestro, Festa and Rolfo2019; Bonnet et al., Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022). This unit is separated by the underlying units by the Ophiolitiferous Band of Henry et al. (Reference Henry, Michard and Chopin1993) or Valmala Shear Zone (Fig. 2; Balestro et al., Reference Balestro, Nosenzo, Cadoppi, Fioraso, Groppo and Festa2020; Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022), consisting of highly deformed tectonic slices of continental and oceanic rocks.

Similar shear zones with both oceanic and continental rocks (Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022) occur above the Dronero Unit: (i) the San Damiano Shear Zone, which separates the Pre-Piemonte Val Maira-Sampeyre Unit from the Dronero Unit (Michard et al.,Reference Michard, Henry and Chopin1993; Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022; Dana, Reference Dana2024), and (ii) Cima Lubin Shear Zone (CLSZ) separating the Val Maira–Sampeyre and Val Grana Units in the south and the continental units of the Dora-Maira Massif from the oceanic-derived Monviso Unit in the north (Fig. 2; Michard, Reference Michard1967; Piana et al., Reference Piana, Fioraso, Irace, Mosca, d’Atri, Barale, Falletti, Monegato, Morelli, Tallone and Vigna2017; Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022).

In Fig. 2, a simplified, updated tectonic sketch of the entire Dora-Maira Massif, compiled from various sources, illustrates the current state of knowledge. It is worthy to emphasize that only the northern and southern portions of the Dora-Maira Massif have been mapped and studied intensively (e.g. Michard et al., Reference Michard, Henry and Chopin1993; Cadoppi et al., Reference Cadoppi, Castelletto, Sacchi, Baggio, Carraro and Giraud2002; Sacchi et al., Reference Sacchi, Balestro, Cadoppi, Carraro, Delle Piane, Di Martino, Enrietti, Gallarà, Gattiglio, Martinotti and Perello2004; Gasco et al., Reference Gasco, Gattiglio and Borghi2011; Compagnoni et al., Reference Compagnoni, Rolfo, Groppo, Hirajima and Turello2012; Groppo et al., Reference Groppo, Ferrando, Gilio, Botta, Nosenzo, Balestro, Festa and Rolfo2019; Reference Groppo, Ferrando, Tursi and Rolfo2025; Bonnet et al., Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022; Manzotti et al., Reference Manzotti, Ballèvre and Poujol2016, Reference Manzotti, Schiavi, Nosenzo, Pitra and Ballèvre2022, Reference Manzotti, Millonig, Gerdes, Whitehouse, Jeon, Poujol and Ballèvre2025a, Reference Manzotti, Schiavi, Ballèvre and Nosenzo2025b; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022, Reference Nosenzo, Manzotti and Robyr2023, Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024; Balestro et al., Reference Balestro, Gattiglio, Roà and Festa2025), whereas the central portion (comprised between the Pellice and Po Valleys, Fig. 2) is still relatively poorly known.

The rock samples studied in this contribution were collected from a large area (Fig. 2), among the Pellice, Germanasca and Chisone Valleys, including the Pinerolo–Sanfront, the ‘Basement Complex’ Units and CLSZ, investigated in the frame of the Geological Map of Italy CARG project – Sheet 172 ‘Pinerolo’.

2.d. The Dora-Maira orthogneisses and metadiorites

Orthogneisses are a very common lithology in the Dora-Maira Massif (Fig. 2). Based on petrographic and geochemical data, and particularly protolith ages, several groups of orthogneisses have been recognized in the past (Sandrone et al., Reference Sandrone, Cordola and Fontan1986; Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993; Borghi et al., Reference Borghi, Cadoppi, Poli, Sacchi and Sandrone1989; Paquette et al., Reference Paquette, Montel and Chopin1999; Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022, Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024). In Table 1, we present an overview of all the different types of orthogneiss, showing a revised nomenclature based on our own original observations, compared and integrated with the existing literature, in order to provide an updated summary of their main features (mineral abbreviations used in the tables, figures and text are after Whitney & Evans (Reference Whitney and Evans2010) except for Wm = white mica).

Table 1. Meta-intrusive rocks of the Dora-Maira Massif and associated regional shear zones (excluding meta-mafic rocks), with their main features (MR = magmatic relics, MM = metamorphic minerals, ACC = accessory minerals) and the proposed or published absolute ages available in the literature prior to this study. Reference: (1) Franchi (Reference Franchi1898), Franchi & Novarese (Reference Franchi and Novarese1895), Novarese (Reference Novarese1895; Reference Novarese1896; Reference Novarese1898) and Stella (Reference Stella1895; Reference Stella1896); (2) Vialon (Reference Vialon1966); (3) Borghi et al. (Reference Borghi, Cadoppi, Porro, Sacchi and Sandrone1984); (4) Sandrone et al. (Reference Sandrone, Cordola and Fontan1986); (5) Sandrone et al. (Reference Sandrone, Sacchi, Cordola, Fontan and Villa1988); (6) Henry (Reference Henry1990); (7) Cadoppi (Reference Cadoppi1990); (8) Wheeler (Reference Wheeler1991); (9) Bussy & Cadoppi (Reference Bussy and Cadoppi1996); (10) Gebauer et al. (Reference Gebauer, Schertl, Brix and Schreyer1997); (11) Paquette et al (Reference Paquette, Montel and Chopin1999); (12) Chen et al. (Reference Chen, Zhou, Zheng and Schertl2017); (13) Balestro et al. (Reference Balestro, Festa, Cadoppi, Groppo and Roà2022); (14) Nosenzo et al. (Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022); (15) Nosenzo et al. (Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024)

In the following, we will describe in detail the different orthogneiss groups recognized in the Dora-Maira Massif and their main features.

2.d.1. Meta-intrusives of the Pinerolo–Sanfront Unit

Three different types of acid to intermediate meta-intrusives have been distinguished in the Pinerolo–Sanfront Unit based on petrography, geochemistry and U–Pb dating (Novarese, Reference Novarese1895; Zanettin-Lorenzoni, Reference Zanettin-Lorenzoni1967; Borghi et al., Reference Borghi, Cadoppi, Porro, Sacchi and Sandrone1984; Cadoppi, Reference Cadoppi1990; Sandrone et al., Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993; Bussy & Cadoppi, Reference Bussy and Cadoppi1996), namely: the Cavour orthogneiss, the Malanaggio metadiorite and the Freidour orthogneiss (Tab. 1, Fig. 2). U–Pb dating on zircons from these three types of meta-intrusives returned Upper Carboniferous to Lower Permian ages (Bussy & Cadoppi, Reference Bussy and Cadoppi1996). Magmatism in the Pinerolo–Sanfront Unit evolved from calc-alkaline to more alkaline melts, starting from the oldest Cavour orthogneiss and progressing to the younger Freidour orthogneiss (Table 1). The Cavour orthogneiss crops out in a small outcrop (<0.08 km2) in the Rocca di Cavour inselberg (Fig. 2; Vialon, Reference Vialon1966; Bussy & Cadoppi, Reference Bussy and Cadoppi1996), whereas the metadioritic to meta-granodioritic Malanaggio metadiorite crops out as several km-scale bodies in the lower Chisone Valley (Fig. 2; Franchi & Novarese, Reference Franchi and Novarese1895; Novarese, Reference Novarese1895; Borghi et al., Reference Borghi, Cadoppi, Porro, Sacchi and Sandrone1984; Sandrone et al., Reference Sandrone, Sacchi, Cordola, Fontan and Villa1988, Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993; Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Vialon, Reference Vialon1966). Original magmatic textures and mineral assemblages (Tab. 1) of the Malanaggio metadiorite are well preserved, mainly in the more acidic ‘quartz-diorite’ facies (Bussy & Cadoppi, Reference Bussy and Cadoppi1996). Two samples have been dated by Bussy and Cadoppi (Reference Bussy and Cadoppi1996), yielding crystallization ages of 290 ± 2 Ma for the ‘quartz-diorite’ facies and 288 ± 2 Ma for the ‘granodioritic’ facies. These ages, slightly older than the one obtained for the Freidour orthogneiss (268–283 Ma; Bussy & Cadoppi, Reference Bussy and Cadoppi1996), are further supported by the field evidence of dykes and sills of the latter cross-cutting the Malanaggio metadiorite (Section 4).

The Freidour orthogneiss, with a granitic composition, makes up a c. 120 km2 body in the study area (Fig. 2). Given the large areal extension, we have studied a sample collected in a different location (Fig. 2; Tab. 2) than the sample studied by Bussy & Cadoppi (Reference Bussy and Cadoppi1996). Orthogneiss bodies in the southern portion of the Pinerolo–Sanfront Unit (Monte Bracco area; Fig. 2) have not been studied in detail, but have been referred to the Freidour orthogneiss (Piana et al., Reference Piana, Fioraso, Irace, Mosca, d’Atri, Barale, Falletti, Monegato, Morelli, Tallone and Vigna2017).

Table 2. Meta-intrusive samples studied in this contribution. Coordinates are given in WGS84 system. Mineral abbreviations after Whitney & Evans (Reference Whitney and Evans2010), except for Wm = white mica

2.d.2. Orthogneisses in the ‘Basement Complex’ Units

Granitic to dioritic meta-intrusives referring to at least two different magmatic cycles are present in these units (Tab. 1, e.g. Vialon, Reference Vialon1966; Compagnoni & Sandrone, Reference Compagnoni and Sandrone1981; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024). In low-strain domains, preserved cross-cutting relationships with host rocks have been described (e.g. Sandrone et al., Reference Sandrone, Cordola and Fontan1986).

The newly defined Ferrera orthogneiss forms a large body of leucocratic augengneiss with interlayered greyish-green felsic gneiss cropping out in the westernmost portion of the Pellice Valley (central Dora-Maira Massif; Fig. 2). This body was originally attributed to the Luserna orthogneiss by Vialon (Reference Vialon1966), whereas Bussy & Cadoppi (Reference Bussy and Cadoppi1996) and Sandrone et al. (Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993) mapped it as part of the Ghiandone orthogneiss. At the map scale, the Ferrera orthogneiss is intruded by apophyses of the Ghiandone orthogneiss (see below). Given these uncertainties, and in the light of our new geochemical and geochronological data (Sections 5 and 6), we introduce here the term Ferrera orthogneiss to distinguish this body from the other orthogneiss groups of the Dora-Maira Massif (Fig. 2).

The Ghiandone, Melle and Granero orthogneisses are attributed to the Ordovician–Early Silurian magmatic cycle, while the remaining lithologies (Brossasco, Luserna, Sangone and Borgone orthogneiss) are generally assigned to the Carboniferous–Permian cycle (Tab. 1). The Ghiandone orthogneiss, originally defined by Novarese (Reference Novarese1895) and later renamed and subdivided (‘Gneiss Amygdalaires’ of Vialon, Reference Vialon1966; ‘polymetamorphic augengneiss’ of Bussy & Cadoppi, Reference Bussy and Cadoppi1996 or ‘Muret orthogneiss’ of Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022), is easily distinguished by its augen-like structure, with large polycrystalline aggregates replacing the original K-feldspar crystals, and frequently presents microgranular mafic enclaves deformed along the main foliation. Two distinct types of metamorphic biotite are described in the Ghiandone orthogneiss: the reddish-brown variety, considered a relict of the Variscan metamorphic event (Borghi et al., Reference Borghi, Cadoppi, Porro, Sacchi and Sandrone1984, Reference Borghi, Cadoppi, Porro and Sacchi1985; Compagnoni & Sandrone, Reference Compagnoni and Sandrone1981; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022), and the greenish biotite, interpreted as an Alpine metamorphic mineral (Sandrone et al., Reference Sandrone, Cordola and Fontan1986). This orthogneiss occurs in several distinct bodies, ranging from decametric to kilometric in size, mainly in the central portion of the Dora-Maira Massif (Fig. 2). One of these bodies, cropping out in the Punta Muret area in the Germanasca Valley, was dated by zircon U–Pb at 457 ± 2 Ma, with older inherited grains around 475 Ma (Bussy & Cadoppi, Reference Bussy and Cadoppi1996). Nosenzo et al. (Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022) for the same body, designated as ‘Muret orthogneiss’ by these authors, suggest a younger emplacement age of 442 ± 2 Ma, based on zircon U–Pb dating. The discrepancy between the two ages is most likely related to the analytical technique employed, that is, ID-TIMS in Bussy & Cadoppi (Reference Bussy and Cadoppi1996) versus LA-ICP-MS in Nosenzo et al. (Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022).

The Melle orthogneiss (Henry, Reference Henry1990; Henry et al., Reference Henry, Michard and Chopin1993), cropping out in the middle Varaita Valley (Fig. 2), has recently been studied petrographically by Groppo et al. (Reference Groppo, Ferrando, Tursi and Rolfo2025). They recognized fine-grained polycrystalline aggregates of polygonal quartz, locally with palisade microstructure, interpreted as indirect evidence of the former occurrence of coesite in the less deformed Melle orthogneiss. Currently, there are no available geochronological or geochemical data for the Melle orthogneiss; however, according to Henry (Reference Henry1990), it is comparable to the Ghiandone orthogneiss (Tab. 1) based on similar petrographic and geochemical features.

The Granero orthogneiss (Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024) occurs as a thin stripe, ranging in thickness from 50 to 200 m, directly overlying the micaschist of the UHP Chasteiran Unit (mapped as part of the ‘Basement Complex’ in Fig. 2). Granero orthogneiss can be followed for at least 10 km, although it crops out discontinuously and is often strongly boudinaged. The Granero orthogneiss was assigned to the Permian magmatic cycle (Borghi et al., Reference Borghi, Cadoppi, Porro, Sacchi and Sandrone1984; Bussy & Cadoppi, Reference Bussy and Cadoppi1996), but recently, Nosenzo et al. (Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024) dated three samples of the Granero orthogneiss, obtaining Concordia ages of 446.7 ± 1.4, 455.6 ± 2.4 and 440.0 ± 2.0 Ma, respectively.

The remaining orthogneiss groups (Brossasco, Sangone, Borgone, Luserna and Castlüs orthogneiss) and metadiorites were partly dated and largely assigned based on correlations and field/geochemical observations to the Permo-Carboniferous magmatic cycle (Vialon, Reference Vialon1966; Sandrone et al., Reference Sandrone, Cordola and Fontan1986, Reference Sandrone, Sacchi, Cordola, Fontan and Villa1988, Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993; Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Cadoppi et al., Reference Cadoppi, Castelletto, Sacchi, Baggio, Carraro and Giraud2002). We also recognize the Cialancia orthogneiss (Fig. 2) after the type locality, previously grouped in the ‘Freidour-type orthogneisses’ by Sandrone et al. (Reference Sandrone, Cordola and Fontan1986) and Bussy & Cadoppi (Reference Bussy and Cadoppi1996), hosted in the ‘Basement Complex’. Sandrone et al. (Reference Sandrone, Cordola and Fontan1986) documented cross-cutting relationships between this orthogneiss and the Ghiandone orthogneiss, leading to interpret all the ‘Freidour-type orthogneisses’ as late-Variscan granitoids (Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Sandrone et al., Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993), although geochronological data are still lacking.

Several metadioritic to tonalitic bodies (Metadiorites, Auct.) have been identified in the ‘Basement Complex’ Units. Occurrences were reported in the Upper Pellice Valley (Novarese, Reference Novarese1896; Giasset diorite of Sandrone et al., Reference Sandrone, Sacchi, Cordola, Fontan and Villa1988). Sandrone et al. (Reference Sandrone, Sacchi, Cordola, Fontan and Villa1988) proposed a Permian age for these rocks based on lithology and regional correlations, although Ar/Ar geochronology did not return geologically meaningful ages. During our fieldwork, we identified a new km-scale metadiorite body on the southeastern slope of Mount Cornour, hereafter referred to as Giasset metadiorite (Tab. 1, Fig. 2).

The Brossasco orthogneiss within the Brossasco-Isasca UHP Unit is the most studied meta-intrusive of the ‘Basement Complex’ (Chopin, Reference Chopin1984; Chopin et al., Reference Chopin, Henry and Michard1991; Paquette et al., Reference Paquette, Montel and Chopin1999; Compagnoni et al., Reference Compagnoni, Rolfo, Groppo, Hirajima and Turello2012; Chen et al., Reference Chen, Schertl, Zheng, Huang, Zhou and Gong2016, Reference Chen, Zhou, Zheng and Schertl2017; Fig. 2). Despite having experienced Alpine peak metamorphic conditions of approximately 660–730°C and 3.8–4.3 GPa (Kienast et al., Reference Kienast, Lombardo, Biino and Pinardon1991; Ferrando et al., Reference Ferrando, Groppo, Frezzotti, Castelli and Proyer2017; Groppo et al., Reference Groppo, Ferrando, Gilio, Botta, Nosenzo, Balestro, Festa and Rolfo2019), it locally preserves undeformed magmatic textures (Biino & Compagnoni, Reference Biino and Compagnoni1992; Bruno et al., Reference Bruno, Compagnoni and Rubbo2001). Low-strain domains of Brossasco orthogneiss show a gradual transition into high-strain and mylonitic zones (e.g. Gilba mylonites; Henry et al., Reference Henry, Michard and Chopin1993). The age of the protolith is loosely constrained by concordant zircon U–Pb ages at 304 ± 3 Ma (Paquette et al., Reference Paquette, Montel and Chopin1999) and in the range of 260–270 Ma (Chen et al., Reference Chen, Zhou, Zheng and Schertl2017). A similar orthogneiss, cropping out in the northern Dora-Maira (Sangone Valley, Fig. 2), is the Sangone orthogneiss, a metamorphosed porphyritic monzogranite (Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Cadoppi, Reference Cadoppi1990), with a magmatic age between 267 and 279 Ma (zircon U–Pb dating, Bussy & Cadoppi, Reference Bussy and Cadoppi1996). The Borgone orthogneiss is a coarse-grained, porphyritic to equigranular, slightly peraluminous, two-mica orthogneiss cropping out in the lower Susa Valley (Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Cadoppi et al., Reference Cadoppi, Castelletto, Sacchi, Baggio, Carraro and Giraud2002). To date, there are no published geochronological data for this orthogneiss.

The Castlüs orthogneiss (previously included in the Freidour-type orthogneisses of Sandrone et al., Reference Sandrone, Cordola and Fontan1986; Tab. 1) crop out along the left slope of the middle Pellice Valley (Fig. 2), where they display kilometre-scale lateral continuity and highly variable thicknesses, ranging from a few metres to several tens of metres. They consist of grey-yellowish orthogneiss with an augen texture and medium grain size. Centimetric feldspar porphyroclasts occur only in the thickest levels (exceeding 3–5 m), whereas thinner layers (<5 m) are typically whitish-yellowish, finer-grained and characterized by feldspar porphyroclasts never larger than 1 cm. Based on their intrusive relationships with the Ghiandone orthogneiss and with meta-mafic rocks, the Castlüs orthogneiss are proposed to be Permian or at least post-Ordovician in age (Sandrone et al., Reference Sandrone, Cordola and Fontan1986).

The Luserna orthogneiss encompasses a heterogeneous group of leucocratic, generally grey-greenish felsic orthogneisses, sometimes with an augen texture, frequently associated with ‘silvery’ phengitic schists or rare quartzite layers (e.g. Sandrone et al., Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993; Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Cadoppi et al., Reference Cadoppi, Castelletto, Sacchi, Baggio, Carraro and Giraud2002; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024). According to Vialon (Reference Vialon1966), it represents a metamorphosed volcano-sedimentary sequence (‘porphyroids arkosiques’) of Permian age. A volcanic origin is also supported by its association with remnants of the Mesozoic carbonate covers (Wheeler, Reference Wheeler1991; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024). Other authors proposed that the Luserna orthogneiss results from Alpine metamorphism of an original association of granites, leucogranites and aplitic and pegmatitic sills (Bortolami & Dal Piaz, Reference Bortolami and Dal Piaz1970; Barisone et al., Reference Barisone, Bottino, Coccolo, Compagnoni, Del Greco, Mastrangelo, Sandrone and Zucchetti1979; Cadoppi, Reference Cadoppi1990; Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Compagnoni & Sandrone, Reference Compagnoni and Sandrone1981), which were emplaced at very shallow crustal levels. The Luserna orthogneiss is widespread in the northern-central portion of the Dora-Maira Massif (Fig. 2). Three main bodies can be distinguished: (i) a long, stripe-shaped body along the western margin of the Dora-Maira, extending from the Chisone Valley to the Pellice Valley, where it progressively thins southward. In the Pellice Valley, its thickness locally decreases to only a few metres, and in the Subiasco Valley, it reaches just a few centimetres (Wheeler, Reference Wheeler1991); (ii) a large body cropping out in the central portion of the Dora-Maira, south of Luserna San Giovanni village (Vialon, Reference Vialon1966; Compagnoni et al., Reference Compagnoni, Crisci and Sandrone1983), which hosts all the main quarries of the Luserna orthogneiss, extensively exploited as building material (e.g. Borghi et al., Reference Borghi, Cadoppi and Dino2016); and (iii) a wide area in the Susa Valley (northern Dora-Maira, Fig. 2), where the Luserna orthogneiss is reported according to published geological maps (Cadoppi et al., Reference Cadoppi, Castelletto, Sacchi, Baggio, Carraro and Giraud2002; Sacchi et al., Reference Sacchi, Balestro, Cadoppi, Carraro, Delle Piane, Di Martino, Enrietti, Gallarà, Gattiglio, Martinotti and Perello2004; Balestro et al., Reference Balestro, Gattiglio, Roà and Festa2025). Despite the extensive distribution of the Luserna orthogneiss, only the first body (‘Clapier orthogneiss’ of Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024; Tab. 1) has been dated so far. A sample collected along the Chisone–Germanasca Valley watershed from this body yielded a U–Pb zircon crystallization age of 270.0 ± 1.8 Ma (Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024).

2.d.3. Orthogneisses and metadiorites of the Valmala Shear Zone

In the southern portion of the Dora-Maira Massif, tectonic lenses made by metabasite and metadiorite are hosted within the Valmala Shear Zone (Henry, Reference Henry1990; Compagnoni et al., Reference Compagnoni, Rolfo, Groppo, Hirajima and Turello2012; Balestro et al., Reference Balestro, Nosenzo, Cadoppi, Fioraso, Groppo and Festa2020; Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022). Despite their abundance and distribution, these lenses are still poorly studied. Balestro et al. (Reference Balestro, Festa, Cadoppi, Groppo and Roà2022) obtained a U–Pb zircon age of 514 ± 3 Ma for the Valmala metadiorite hosted within the Valmala Shear Zone (Fig. 2).

2.d.4. Orthogneiss of the Dronero Unit

The orthogneiss body in the Dronero Unit (Michard et al., Reference Michard, Henry and Chopin1993) (Fig. 2) was named Birrone orthogneiss (Tab. 1; Henry, Reference Henry1990). It consists of a heterogeneous set of K-feldspar-bearing, coarse- to medium-grained augengneiss derived from granite and tonalite (Balestro et al., Reference Balestro, Cadoppi, Di Martino and Sacchi1995; Reference Balestro, Festa, Cadoppi, Groppo and Roà2022), with minor medium-grained leucogneiss of granodioritic composition. This orthogneiss has not been dated directly, but a Permian age has been suggested (Henry, Reference Henry1990; Balestro et al., Reference Balestro, Festa, Cadoppi, Groppo and Roà2022; Michard et al., Reference Michard, Henry and Chopin1993, Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022) based on the association with rhyolite porphyroids attributed to the Permian s.l. (Michard & Vialon, Reference Michard and Vialon1966). The Dronero Unit also contains intercalations of micro-augengneiss levels within metasediments, which have been interpreted as reworked acidic volcaniclastic layers (Bonioli et al., Reference Bonioli, Cadoppi, Sacchi, Carmignani and Sassi1992) and dated at 253.8 ± 2.7 Ma based on zircon U–Pb (Balestro et al., Reference Balestro, Festa, Cadoppi, Groppo and Roà2022; Fig. 2).

3. Methodology

Twenty-one samples (Tab. 2) of meta-intrusive rocks from the main tectonic units building up the Dora-Maira Massif exposed in the Pellice-Chisone Valleys were studied. Thin sections have been cut parallel to the mineral lineation and perpendicular to the main foliation to better approximate the XZ section of the finite strain ellipsoid. Petrography and microstructure were determined by optical microscopy and Scanning Electron Microscope (SEM) equipped with energy dispersive X-ray spectroscopy (EDS), hosted at the Earth Sciences Department of the University of Turin.

All samples were analysed for major, minor and trace elements with ICP-MS following the 4Lithores protocol at Activation Labs (Ancaster, Canada). The rock material was crushed and milled for eight selected samples to separate zircon crystals from the light mineral fraction using a Wilfley Table. Zircon crystals were handpicked and mounted in epoxy resin and polished to expose the grain centres. Cathodoluminescence (CL) imaging was carried out with a ZEISS EVO50 hosted at the Institute of Geological Sciences, University of Bern, at low vacuum conditions (15.00 kV and 9.6 mm working distance). Post-analysis imaging of spot positions was carried out with a SEM using combined secondary and backscattered electrons and similar working conditions (12 kV, 750 pA and 9.5 mm working distance). The mineralogy of selected inclusions in zircons was identified by EDS analysis at the SEM.

U–Pb geochronology of zircon was performed by LA-ICP-MS with a Resonetics RESOlutionSE 193 nm excimer laser system coupled to an Agilent 7900 quadrupole ICP-MS at the University of Bern. The ICP-MS was tuned for low oxide production (ThO/Th < 0.2%) and a Th/U ratio close to one (Th/U > 97%). Analyses were performed with a 25 μm spot size with pre-ablation at 30 µm, at a repetition rate of 2.5 Hz and an energy density on the sample of 5 J/cm2. Temora zircon (Black et al., Reference Black, Kamo, Allen, Aleinikoff, Davis, Korsch and Foudoulis2003) was used as a primary standard and yielded a Concordia age of 416.1 ± 1.2 Ma (95% c.i.; MSWD (concordance + equivalence) = 0.57; n = 34). The Plešovice (Sláma et al., Reference Sláma, Košler, Condon, Crowley, Gerdes, Hanchar and Whitehouse2008) and 91500 zircon (Wiedenbeck et al., Reference Wiedenbeck, Alle, Corfu, Griffin, Meier, Oberli and Spiegel1995) were used as secondary reference materials and yielded a Concordia age of 344.6 ± 1.1 Ma (MSWD = 1.2; n = 20) and 1061.9 ± 3.3 Ma (MSWD = 0.56; n = 21). Raw data were processed with the software Iolite v4 (Paton et al., Reference Paton, Woodhead, Hellstrom, Hergt, Greig and Maas2010, Reference Paton, Hellstrom, Paul, Woodhead and Hergt2011). Concordia ages and diagrams were generated with IsoplotR (Vermeesch, Reference Vermeesch2018). Concordia ages are reported at the 95% confidence level, with an additional 1% external long-term uncertainty propagated to the average ages. Ages are defined by consistent groups of analyses (n > 3 and MSWD < 2.5). Mean square of weighted deviates (MSWD) values are reported for concordance plus equivalence.

4. Sample description

Samples are described according to the structural position of the related tectonic units within the nappe pile (from the bottom to the top). From the Pinerolo–Sanfront Unit, we collected three samples: two (C175, C168) from the Malanaggio metadiorite and one (DD112) from the Freidour orthogneiss, on the right side of the Chisone Valley north of Perosa Argentina (Fig. 2, Tab. 2).

Seventeen samples of meta-intrusive have been collected in the ‘Basement Complex’: (i) four samples from the Luserna orthogneiss: one sample from the body cropping out at the western border of the Dora-Maira Massif (DD109), one (SC41) from the Subiasco Valley (SE of Punta Cornour, Fig. 2, Tab. 2) and two from the large body cropping out south of the Luserna San Giovanni village (DC76, DC77); (ii) five samples from the Cialancia orthogneiss: four from the leucocratic facies cropping out in the Germanasca (Y6 and SC2) and Lower Pellice valley (DC75 and C211) and one (Y16) from the dark band layered facies; (iii) two samples of Giasset metadiorite: one from a dyke, cross-cutting the Ghiandone orthogneiss body cropping out in the Pellice Valley (SC59), and one from the main intrusive body in the Pellice Valley (SC58); (iv) two samples of Ghiandone orthogneiss: one collected north of Punta Cornour (Germanasca-Pellice watershed ridge, 36A, Fig. 2) and one collected north of mount Vandalino (DC51; Fig. 2); (v) three samples (SD2, DD244, DD245) from the Ferrera orthogneiss cropping out west of Bobbio Pellice (Upper Pellice Valley). Finally, a previously unknown orthogneiss sample (hereafter referred to as Giulian orthogneiss, AB21) from CLSZ was collected near the Germanasca-Pellice watershed ridge (Fig. 2; Tab. 2). Mineral phases documented in the collected samples are listed in Table 2.

4.a. Malanaggio metadiorite

The Malanaggio metadiorite of the Pinerolo–Sanfront Unit is composed of fine-grained dioritic gneisses with a typical grey colour and a banded appearance (Fig. 3a), locally containing centimetre to decimetre-sized mafic microgranular enclaves (MME). The collected samples (C168, C175) have a spaced foliation defined by the presence of microlithons alternating with cleavage domains, with quartz-albite domains alternating with actinolite, biotite and chlorite layers (Fig. 3b). Rare relicts of magmatic hornblende are preserved. Quartz exhibits undulose extinction, and it is recrystallized by the subgrain rotation (SGR) mechanism (Stipp et al., Reference Stipp, StuÈnitz, Heilbronner and Schmid2002a, Reference Stipp, Stünitz, Heilbronner and Schmid2002b).

Figure 3. Selected field images and photomicrographs of the studied lithologies. (a) Outcrop of Malanaggio metadiorite with the typical banded appearance and microgranular mafic enclaves. (b) Crossed polarized light (CPL hereafter) microphoto of the Malanaggio metadiorite. (c) Fine-grained facies of the Freidour orthogneiss. (d) Dyke of Freidour orthogneiss cross-cutting the Malanaggio metadiorite. (e) CPL photomicrograph of a large K-feldspar porphyroclast wrapped by the phengite foliation in the Freidour orthogneiss. (f) Field image of the Ferrera orthogneiss felsic facies, W of Bobbio Pellice village. (g) Field image of the Ferrera orthogneiss leucocratic facies (100 m from the outcrop shown in f). (h) CPL photomicrograph of the Ferrera orthogneiss (sample SD2) showing the spaced foliation defined by white mica.

4.b. Freidour orthogneiss

The Freidour orthogneiss sample (DD112, Pinerolo–Sanfront Unit) is a fine-grained metagranite (Fig. 3c), locally containing decimetre-sized MME in the coarser-grained portions. Aplitic, fine-grained sills and dykes, connected to the main orthogneiss body, are observed in the host micaschist at the contact with the Freidour orthogneiss, and they also crosscut the Malanaggio metadiorite (Fig. 3d). In the collected sample, K-feldspar porphyroclasts (microcline) are generally centimetre or sub-centimetre in length, whereas in the biotite-rich coarser-grained portions, they can reach 3–4 cm in length (Fig. 3e). The main foliation is a spaced one (Passchier & Trouw, Reference Passchier and Trouw2005) and it is defined by white mica (phengite), chlorite and rare biotite, alternating with fine-grained quartz-feldspathic microlithons. Quartz exhibits undulose extinction, and it is recrystallized by the SGR mechanism. Biotite generally grows at the edges of white mica and is partially replaced by chlorite.

4.c. Ferrera orthogneiss

The Ferrera orthogneiss samples SD2 and DD245 are greenish-grey felsic orthogneisses (Fig. 3f), whereas sample SD244 comes from a more leucocratic facies (Fig. 3g; 100 m from the latter samples). Intrusive relationships at the cartographic scale are observed between the Ghiandone orthogneiss and the host Ferrera orthogneiss, which is likely older. At the microscale, the samples are characterized by a continuous to spaced foliation defined by white mica (mainly phengite). Euhedral titanite and fine-grained chlorite are also observed along the foliation planes. Plagioclase is subhedral, showing polysynthetic twinning, whereas K-feldspar forms delta-type porphyroclasts with simple twinning, and quartz and white mica inclusions are often concentrated in the core of crystals (Fig. 3h). Feldspars show intense synkinematic recrystallization in the BLG regime. Rare mica fish, occurring in the fine-grained quartz-rich matrix, suggest a top-to-W sense of shear. Quartz shows microstructures pointing to a grain-boundary migration (GBM) recrystallization mechanism (Stipp et al., Reference Stipp, StuÈnitz, Heilbronner and Schmid2002a, Reference Stipp, Stünitz, Heilbronner and Schmid2002b), with local evidence for annealing and overprinting of SGR.

4.d. Ghiandone orthogneiss

The Ghiandone orthogneiss samples (36A and DC51) are characterized by the presence of abundant (i) large polycrystalline aggregates resembling K-feldspar megacrysts (up to 10 cm), hosted in a fine-grained matrix (Fig. 4a), and (ii) mafic microgranular enclaves. Because of Alpine metamorphisms, the K-feldspar is almost entirely replaced by fine-grained aggregates of albite and quartz. Fresh K-feldspar is preserved only in very low-strain domains, such as the Punta Muret outcrop (Novarese, Reference Novarese1895; Bouffette et al., Reference Bouffette, Lardeaux and Caron1993; the Muret orthogneiss of Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022). The main foliation is spaced, and a crenulation cleavage is rarely present. The foliation is defined by abundant white mica (mainly phengite), associated with minor biotite and chlorite along the cleavage domains. Small garnet (up to 2–3 % in volume), epidote and titanite are found along the foliation (Fig. 4b). Microlithons are mainly made of fine-grained aggregates of quartz and plagioclase.

Figure 4. Selected field images and photomicrographs of the studied lithologies. (a) Ghiandone orthogneiss showing the distinctive large polycrystalline aggregates porphyroclasts and microgranular mafic enclaves. (b) CPL photomicrograph of the Ghiandone orthogneiss. (c) Field image of the Granero orthogneiss with cm-sized K-feldspar porphyroclast. (d) CPL photomicrograph of Granero orthogneiss (sample DD43) showing K-feldspar porphyroclast enveloped by the main foliation defined by white mica and quartz-felspathic levels. (e) Typical banding of the Cialancia orthogneiss, with alternating dark and leucocratic levels. (f) Aplitic dyke associated to the Cialancia orthogneiss intruding the host micaschist. (g) CPL photomicrograph of Cialancia orthogneiss (sample Y6) showing a K-feldspar porphyroclast surrounded by a matrix of plagioclase, quartz and phengite. (h) Giasset metadiorite dyke cross-cutting the Ghiandone orthogneiss, suggesting a younger protolith age for the metadiorites.

4.e. Granero orthogneiss

The Granero orthogneiss sample (DD43) is a white to grey augengneiss (Fig. 4c), collected in the same outcrop dated by Nosenzo et al. (Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024). It displays a spaced foliation defined by fine-grained layers of quartz and the shape-preferred orientation of white mica (Fig. 4d). K-feldspar porphyroclasts (mainly microcline) up to cm-sized are wrapped by the main foliation and contain inclusions of epidote and white mica. Quartz is dynamically recrystallized in the GBM/SGR regime.

4.f. Cialancia orthogneiss

The Cialancia orthogneiss displays different facies grain sizes and deformation features, ranging from portions where the magmatic fabric is still preserved to a banding with alternations of leucocratic and darker levels with different mineralogy (Fig. 4e). Leucocratic aplitic sills, which are boudinaged and folded, are associated with the main Cialancia orthogneiss (Fig. 4f). Samples SC2, DC75, C211 and Y6 were collected within the dominant strongly deformed leucocratic facies with K-feldspar porphyroclasts, reaching up to 4–5 cm, wrapped by the main spaced foliation defined by white mica and biotite. Sample Y18 was collected in the dark layered facies of the Cialancia orthogneiss. Magmatic biotite relicts are rare in all samples. Quartz often occurs in ribbons, associated with plagioclase, and it is dynamically recrystallized in the GBM/SGR regime (Fig. 4g). K-feldspar is commonly microcline and more rarely orthoclase (with the typical Carlsbad twinning).

4.g. Giasset metadiorite

The Giasset metadiorite is fine-grained and variably deformed (Fig. 4h and 5a). Where intensively deformed, a continuous to spaced foliation, defined by white mica and amphibole, is observed. Fine aggregates of biotite and amphibole could represent pseudomorphs after original magmatic amphibole. Garnet is observed in very small grains, partially overgrown by biotite rims. Sample SC58 was collected from a hundred-metre-sized metadiorite body associated with several dykes. In the main body, microgranular mafic enclaves with dark colours and elongated shape are common (Fig. 5a). Sample SC59 was collected from one of the aforementioned dykes, which crosscuts the Ghiandone orthogneiss (Tab. 1), suggesting that Giasset metadiorite is younger (Fig. 4h).

Figure 5. Selected field images and photomicrographs of the studied lithologies. (a) Giasset metadiorite containing cm-sized mafic microgranular enclaves. (b) CPL photomicrograph of the Giasset metadiorite showing fine-grained Alpine mineral crystallization (sample SC58). (c) Field image of the Luserna orthogneiss, with associated quartz-phengitic micaschist. (d) CPL photomicrograph of the Luserna orthogneiss (sample DD109) showing a K-feldspar porphyroclast, wrapped by the main foliation. (e) Field view of the Giulian orthogneiss body sampled within the Cima Lubin Shear Zone, view from the Pellice side of the Germanasca-Pellice watershed ridge. (f) CPL photomicrograph of the Giulian orthogneiss sample (AB21) showing quartz-rich microlithons and a continuous to spaced foliation defined by white mica.

4.h. Luserna orthogneiss

Collected samples (DD109, DC76, DC77 and SC41) are felsic orthogneiss with mm-sized K-feldspar porphyroclasts, frequently associated with quartz-phengitic (Fig. 5c) micaschist and quartzite levels. In all the samples, at the microscale, porphyroclasts are surrounded by a spaced foliation defined by alternation of micaceous cleavage domains and quartz-feldspathic-rich microlithons (Fig. 5d). K-feldspar is locally recrystallized by bulging (BLG), whereas quartz shows recrystallization mechanisms by SGR, locally overprinted by a later BLG mechanism.

4.i. Giulian orthogneiss

Based on our field mapping, we report the presence of several lenses of leucocratic orthogneiss, no more than 10 m thick, dismembered within a thick band of mylonitic calcschist separating the Dora-Maira Massif from the overlying oceanic Monviso Unit in the Pellice Valley. These lenses (Fig. 5e), among other tectonic slices of dolostone, serpentinite, metabasite and micaschist, are particularly widespread in the Colle Giulian area (Pellice Valley, Fig. 2), where sample AB21 was collected. This high-strain zone, with both tectonic slices of oceanic and continental rocks, was previously considered part of the continental Giulian-Sea Bianca Unit of Balestro et al. (Reference Balestro, Fioraso and Lombardo2011) and is interpreted here as the northern extension of the CLSZ (Piana et al., Reference Piana, Fioraso, Irace, Mosca, d’Atri, Barale, Falletti, Monegato, Morelli, Tallone and Vigna2017), defined further south by Michard et al. (Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022).

This sample consists of a leucocratic fine-grained orthogneiss, with a continuous to spaced foliation defined by alternating white mica (mainly phengite), titanite and quartz-felspathic domains. It also contains very small feldspar porphyroclasts, mainly microcline, and minor albite. Quartz is recrystallized in the GBM regime (Fig. 5f).

5. Geochemistry results

The major element composition of the 21 analysed samples (Table 2; Additional file 1) plots in the granite field in a Total Alkali (Na2O + K2O) versus Silica (TAS) diagram (Fig. 6a; Middlemost, Reference Middlemost1994), with the exceptions of the Ghiandone orthogneiss (between the granodiorite and granite fields), Malanaggio metadiorite (granodiorite and diorite fields) and the Giasset metadiorite (among the monzo-diorite, gabbroic diorite and diorite fields). In the De la Roche et al. Reference De la Roche, Leterrier, Grandclaude and Marchal(1980) R1–R2 plot (R1 (millications) = 4Si–11(Na + K)–2(Fe + Ti) and R2 (millications) = 6Ca + 2 Mg + Al; Fig. 6b), the Freidour, Ferrera, Granero, Cialancia, Luserna and Giulian orthogneisses plot between the granite and alkaline-granite fields. In contrast, the Ghiandone orthogneiss samples plot closer to the granite-granodiorite transition, the two Giasset metadiorite samples plot in the gabbrodiorite and diorite fields and the Malanaggio metadiorite samples plot in the granodiorite field. According to the classification of Frost and Frost (Reference Frost and Frost2008), the studied meta-intrusives are peraluminous (Alumina Saturation Index (ASI) between 1.10 and 1.50; Fig. 6c), with only the Giasset metadiorite samples being metaluminous. The studied Malanaggio metadiorite samples are peraluminous, although some literature data plot within the metaluminous field. The samples vary in SiO2 content between 52 and 79 wt %, with the Giasset metadiorite being the lowest in content and the Ferrera orthogneiss being the most quartz-rich. In the Pearce (Reference Pearce and Wyman1996) Zr/TiO2–Nb/Y diagram (Fig. 6d), most of the studied samples plot within the granite–granodiorite field, whereas the Malanaggio metadiorite plots in the diorite–gabbrodiorite, gabbro and monzonite fields, and the Giasset metadiorite falls between the diorite–gabbrodiorite and gabbro fields.

Figure 6. Results of whole-rock bulk-rock geochemistry. (a) TAS diagram for intrusive rocks (Middlemost, Reference Middlemost1994). (b) R1–R2 classification diagram of the De la Roche et al. (Reference De la Roche, Leterrier, Grandclaude and Marchal1980). R1 (millications) = 4Si–11(Na + K)–2(Fe + Ti) and R2 (millications) = 6Ca + 2 Mg + Al. (c) Aluminium Saturation Index (Al/Ca-1.67P+N+K) versus silica plot of Frost and Frost (Reference Frost and Frost2008). (d) Zr/TiO2 -Nb/Y classification of granitic rocks after Pearce (Reference Pearce and Wyman1996), note that fewer data points are shown due to the absence of complete trace element analyses in several literature sources. Literature data are from: Vialon (Reference Vialon1966), Sandrone et al. (Reference Sandrone, Barla, Bianco, Compagnoni and Giani1982); Borghi (Reference Borghi1983); Sandrone et al. (Reference Sandrone, Cordola and Fontan1986); Sandrone et al. (Reference Sandrone, Sacchi, Cordola, Fontan and Villa1988), Cadoppi (Reference Cadoppi1990); Bussy & Cadoppi (Reference Bussy and Cadoppi1996); Chen et al. (Reference Chen, Schertl, Zheng, Huang, Zhou and Gong2016); Nosenzo et al. (Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022) and Nosenzo et al. (Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024).

In the Pearce et al. (Reference Pearce, Harris and Tindle1984) Nb–Y and Ta–Yb petrotectonic discrimination diagrams (Additional File 2), the orthogneiss groups plot mainly in the volcanic-arc and syn-collisional granite fields; the Luserna orthogneiss samples also display within-plate affinities, whereas the Cialancia orthogneisses plot exclusively in the syn-collisional field.

The analysed samples have Rare Earth Element (REE) patterns typical of felsic crustal rocks with enrichment of Light Rare Earth Elements (LREE) over Heavy Rare Earth Elements (HREE) and a variable Eu anomaly but also show significant variability across rock types (Fig. 6). The REE composition of the Freidour orthogneiss sample matches what has been reported by previous studies (Cadoppi, Reference Cadoppi1990; Bussy & Cadoppi, Reference Bussy and Cadoppi1996), with a moderate negative Eu anomaly (Eu/Eu* = 0.25; calculated with the formula Eun/SQRT(Smn*Gdn)). REE pattern and negative Eu anomaly of the Freidour orthogneiss is not significantly different from that of the Cavour orthogneiss, also part of the Pinerolo–Sanfront Unit, as studied by Bussy & Cadoppi (Reference Bussy and Cadoppi1996; Fig. 7a). The newly defined Ferrera orthogneiss is depleted in sample SD2 in all REE, whereas sample DD244 and DD245 compare well with the Giulian orthogneiss sample (Fig. 7b). Ferrera orthogneiss presents a moderate fractionation of LREE with respect to HREE and a moderate to strong negative Eu anomaly (Eu/Eu* = 0.37-0.17). The REE pattern of the Ghiandone orthogneiss sample shows a weaker Eu anomaly (Eu/Eu* = 0.40–0.55; Fig. 7c), in agreement with previous data (Sandrone et al., Reference Sandrone, Cordola and Fontan1986; Bussy & Cadoppi, Reference Bussy and Cadoppi1996), and well comparable with the Granero orthogneiss sample pattern (Fig. 7c). The REE pattern of the Cialancia orthogneiss, until now considered as a post-Variscan orthogneiss equivalent to the Brossasco, Sangone and Castlüs orthogneiss (Sandrone et al., Reference Sandrone, Cordola and Fontan1986), is plotted against the latter (Fig. 7d). The five Cialancia orthogneiss samples have a more restricted REE composition that is however significantly lower than what is reported in the literature for the considered equivalent orthogneiss (Castlüs and Brossasco orthogneiss, Fig. 7d), with the Eu anomaly ranging between 0.15 and 0.24. The two samples of Giasset metadiorite, hosted in ‘Basement Complex’ Units, have a particularly flat REE pattern comparable to previous data (Sandrone et al., Reference Sandrone, Sacchi, Cordola, Fontan and Villa1988) (Fig. 7e). These are the only samples lacking a significant negative Eu anomaly (Eu/Eu* = 0.97–0.83); this feature distinguishes the Giasset metadiorite from the Malanaggio metadiorite (Eu/Eu* = 0.67–0.60), hosted in the Pinerolo–Sanfront Unit. In accordance with the variable textures observed in the field, the four samples of Luserna orthogneiss show great variability in their REE content and a particularly strong negative Eu anomaly (Eu/Eu* = 0.04–0.27; Fig. 7f), as already reported by Cadoppi (Reference Cadoppi1990). The Giulian orthogneiss sample, part of the CLSZ, has a moderate fractionation of LREE compared to HREE and a negative Eu anomaly (Eu/Eu* = 0.20), comparing well with samples DD244 and DD245 of the Ferrera orthogneiss (Fig. 7b).

Figure 7. Summary of REE patterns of the studied samples. REE patterns are normalized to Chondrite values (Sun & McDonough, Reference Sun and McDonough1989) and compared to data from previous studies (grey areas). Where Pr, Tb and Tm were not analysed, dashed lines are used to join Ce and Nd, Gd and Dy and Er and Yb, respectively.

6. Geochronology results

Eight orthogneiss samples were selected for zircon U–Pb geochronology (Tab. 2). The isotopic results are available in Additional File 3.

6.a. Freidour orthogneiss

Zircon crystals from sample DD112 are prismatic and display oscillatory zoning in CL (Fig. 8a). Texturally older cores have not been observed. Th and U concentrations can be very high: Th is mainly between 52 and 587 µg/g with some analyses as high as 1000, 2530 and 3730 µg/g, and U concentrations range from 301 to 20910 µg/g. Th/U ranges from 0.08 to 0.54. Most of the analyses yield dates in the range of 210–286 Ma, with considerable scatter along Concordia. The oldest group of dates defines a Concordia age of 267.2 ± 2.7 Ma (MSWD = 2.4; n = 13; Fig. 8a), which is taken as the minimum protolith emplacement age. This interpretation, however, should be treated with caution due to the scatter in the data.

Figure 8. U–Pb zircon geochronology of orthogneiss samples: (a) Freidour orthogneiss; (b) Ferrera orthogneiss; (c) and d) Cialancia orthogneiss. Cathodoluminescence images of zircon crystals are shown on the left side, with the LA-ICP-MS spot position (red circles) and the relative date indicated. The Concordia diagrams on the right side report Concordia ages. Empty dashed ellipses represent dates excluded from the calculation.

6.b. Ferrera orthogneiss

Zircon grains from sample SD2 are either euhedral or subrounded with simple zoning (Fig. 8b). The concordant zircon analyses have Th and U concentrations in the range of 62–3180 and 141– 3648 µg/g, with Th/U between 0.09 and 0.98. One concordant zircon analysis yields a 1051.1 ± 16 Ma 206Pb/238U date, whereas the remaining concordant analyses yield Ediacaran–Lower Cambrian dates that define a Concordia age of 540.0 ± 5.4 Ma (MSWD = 0.94, n = 7, Fig. 8b). Two zircon rims yield concordant Ordovician dates, whereas two others provide Permian dates (of which only one is concordant at 267.9 ± 1.4 Ma). These zircons could reflect later metamorphic events; however, since they do not meet the minimum statistical requirements to define an age (n > 3), they are not considered geologically significant.

6.c. Cialancia orthogneiss

Zircon grains in the three studied Cialancia Orthogneiss samples (Y6, Y18 and SC2, Table 1) exhibit common features in all samples. Crystals are prismatic and show weak to marked oscillatory zoning. Zircon cores with a rounded or prismatic shape are less common in samples Y6 and SC2 but more common in sample Y18, which represents the darker portion of the Cialancia orthogneiss. Th and U concentrations are strongly variable in the range of 30–1110 and 373–13570 µg/g, respectively, with Th/U mainly between 0.01 and 0.17. Analyses obtained on inherited cores yield concordant dates around 1 Ga or between 500 and 600 Ma. Most of the remaining analyses provide concordant dates in the range of 416–481 Ma. The main cluster defines a Concordia age at 452.0 ± 4.5 Ma (MSWD = 2.3; n = 11; Fig. 8c) with a subset of data scattering along Concordia to lower values. In Sample Y18, zircon cores have Th/U comprised between 0.40 and 0.50, whereas the rims can reach very high U contents (596–6500 µg/g) and have lower Th/U between 0.03 and 0.17. A few inherited cores yield concordant dates of 1649 ± 21, 1366 ± 18 Ma and two between 600 and 650 Ma. A cluster of core analyses gives a Concordia age of 538.0 ± 5.4 Ma (MSWD = 1.9; n = 6; Fig. 8d). The rim analyses define a Concordia age of 452.0 ± 4.5 Ma (MSWD = 1.8; n = 5; Fig. 8d). In sample SC2, a few U-rich inherited zircon cores have dates older than 1 Ga (three concordant 207Pb/208Pb dates at 1914 ± 33, 2442 ± 30 and 2911 ± 29 Ma), and the majority is between 500 and 550 Ma. An isolated zircon dated at c. 393 Ma is not interpreted as geologically significant. The youngest analyses define a significant cluster with a Concordia age of 445.3 ± 4.5 Ma (MSWD = 1.7; n = 9; Fig. 9a).

Figure 9. U–Pb zircon geochronology of orthogneiss samples: (a) Cialancia orthogneiss; (b) Giasset metadiorite; (c) Luserna orthogneiss; (d) Giulian orthogneiss (tectonic slice in Cima Lubin shear zone). Cathodoluminescence images of zircon crystals are shown on the left side, with the LA-ICP-MS spot position (red circles) and the relative date indicated. The Concordia diagrams on the right side report Concordia ages. Empty dashed ellipses represent dates excluded from the calculation.

6.d. Giasset metadiorite

Zircon crystals extracted from sample SC59 are colourless subhedral crystals with mainly oscillatory zoning (Fig. 9b). Th and U concentrations are in the range of 87–3000 µg/g and 378–4770 µg/g, respectively, with Th/U between 0.13 and 0.86. U–Pb analyses scatter along Concordia without defining significant clusters (Fig. 9b), preventing the estimation of a geologically meaningful age. Concordant analyses are mainly around 500–600 Ma and 300 Ma.

6.e. Luserna orthogneiss

Zircon crystals from sample DD109 are prismatic and display oscillatory zoning in CL (Fig. 9c). An inherited core with a dark CL emission occurs in one grain only. Th and U concentrations of concordant analyses range between 30–80 and 407–6040 μg/g, respectively, with a variable Th/U ratio from 0.01 to 1.19. The inherited core yields a concordant 206Pb/238U date of 379.1 ± 6.3 Ma. Most of the remaining analyses yield dates in the range of 208–297 Ma, with a cluster defining an age of 265.3 ± 2.7 (MSWD = 1.7; n = 11; Fig. 9c).

6.f. Giulian orthogneiss

Zircon crystals extracted from sample AB21 are subhedral with oscillatory zoning (Fig. 9d). Th concentrations are in the range of 28–760 µg/g with three analyses between 2500 and 3310 µg/g, while U concentrations range from 97 to 5630 µg/g. The Th/U of the concordant analyses is between 0.24 and 0.59, whereas the remaining grains scatter widely between 0.02 and 2.34. Five discordant analyses yield dates older than 550 Ma; the oldest concordant dates define a Concordia age at 548.0 ± 5.5 Ma (MSWD = 1.6; n = 4; Fig. 9d), whereas younger dates between 300 and 200 Ma are generally discordant. Despite being defined by only four analyses, the age of the oldest zircon population concurs with that found in the Ferrera sample, and so it can be taken as geologically significant.

7. Discussion

7.a. Geochronology and geochemistry of the Dora-Massif meta-intrusive rocks

U–Pb geochronology data, combined with whole-rock geochemical analyses, allow distinguishing different suites of meta-intrusive rocks. The following discussion on the geochemistry and geochronology of the samples considers the fact that the Dora-Maira rocks have undergone one or more orogenic cycles and, at least during Alpine metamorphism, HP to UHP metamorphic conditions were reached (Manzotti et al., Reference Manzotti, Schiavi, Nosenzo, Pitra and Ballèvre2022; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022), likely involving fluid circulation and metasomatism. However, as argued by Grevel et al. (Reference Grevel, Schreyer, Grevel, Schertl and Willner2009), the REE in the rocks of the Brossasco-Isasca Unit, reaching far more extreme metamorphic conditions with respect to our samples, likely remained immobile during metamorphism. This suggests that, although caution is required, some chemical indicators can still be used to classify the different samples. Metamorphism and deformation can also affect the zircon U–Pb systematics, although complete recrystallization of magmatic zircon in orthogneiss is not expected even under extreme conditions. For instance, the most intensely metamorphosed samples of the Brossasco-Isasca Unit are whiteschists derived from metasomatism and UHP metamorphism of a metagranite (Chen et al., Reference Chen, Zhou, Zheng and Schertl2017). In this rock type, magmatic zircon cores still preserve the protolith Permian age, whereas new growth of zircon rims testifies Alpine metamorphism (Gebauer et al., Reference Gebauer, Schertl, Brix and Schreyer1997; Gauthiez-Putallaz et al., Reference Gauthiez-Putallaz, Rubatto and Hermann2016; Chen et al., Reference Chen, Zhou, Zheng and Schertl2017). The zircons analysed in this work predominantly show oscillatory zoning, either in corroded and rounded cores (likely inherited) or in the euhedral domain. Only rarely do they show possible recrystallization textures (Corfu et al., Reference Corfu, Hanchar, Hoskin and Kinny2003; Rubatto, Reference Rubatto2017), suggesting partial reopening of the U–Pb system, possibly triggered by metamorphism or deformation (e.g. sample AB21, Giulian orthogneiss). Alpine metamorphic rims, if present, are generally very thin, preventing accurate age determination, and might contribute to mixed ages if partly sampled when analysing zircon external domains. We always take the age of the predominant, euhedral oscillatory zoning domains as the age of zircon, and thus magmatic protolith, crystallisation. Partial disturbance of the U–Pb system during metamorphism, while preserving magmatic zoning, would produce scattered dates that could still appear concordant within the uncertainty of the analyses. Therefore, age interpretation is limited to Concordia ages (according to IsoplotR, Vermeesch, Reference Vermeesch2018) defined by a consistent group of analyses (n > 3 and MSWD < 2.5). We discuss our results from the oldest to the youngest magmatic suites dated.

The Ferrera and Giulian orthogneisses represent the oldest magmatic suite recognized in the study area. A leucocratic block of Giulian orthogneiss, dismembered within the CLSZ (sample AB21), yields an Ediacaran zircon core age of 548.0 ± 5.5 Ma, which is within uncertainty of the age for the Ferrera orthogneiss (540.0 ± 5.4 Ma; Fig. 8b). The REE patterns of these samples are generally similar (samples DD244, DD245 and AB21), except for sample SD2, which shows a marked depletion of all REEs (Fig. 7b). The compositional differences observed among the Ferrera orthogneiss samples, despite being collected only a few metres apart but from different facies, suggest that the body records internal chemical variability, possibly related to primary magmatic heterogeneities (e.g. Poli & Tommasini, Reference Poli and Tommasini1991; Dini et al., Reference Dini, Innocenti, Rocchi, Tonarini and Westerman2002). These samples provide clear evidence that the ‘Basement Complex’ Units of the Dora-Maira Massif contain Cambrian-to-Precambrian components. In addition, the zircon crystals from the Giulian orthogneiss sample AB21 also yield three scattered concordant dates between 260 and 280 Ma. These dates might be the consequence of a Permian high-temperature overprint (Fig. 9d), a thermal event widely documented in the continental crust of the Western and Southern Alps (e.g. Kunz et al., Reference Kunz, Manzotti, von Niederhäusern, Engi, Darling, Giuntoli and Lanari2018).

Analyses of the Ghiandone, Granero and Cialancia orthogneisses (Tab. 1) confirm a geochemical difference between these Ordovician to Silurian magmatic suites (Fig. 7c and 7d). Zircon in the Cialancia orthogneiss yields Ordovician to Silurian ages, older than previously inferred Permian ages attributed by previous authors (e.g. Sandrone et al., Reference Sandrone, Cadoppi, Sacchi, Vialon and von Raumer J.F.1993; Bussy & Cadoppi, Reference Bussy and Cadoppi1996). Sample Y18 of the Cialancia orthogneiss shows different ages between the zircon core and rims, which is also reflected by different Th/U ratios. The rim Concordia age of 453.0 ± 4.5 Ma matches the age of the other Cialancia orthogneiss samples (Y6 = 452.0 ± 4.5 and SC2 = 445.3 ± 4.5 Ma), which is therefore interpreted as its most likely age of emplacement (approximately 450.0 ± 5.0 Ma). The Concordia age of 538.0 ± 5.4 Ma (Fig. 8c) obtained from six zircon cores is interpreted as the age of an earlier magmatic suite (the Ferrera orthogneiss, see above), which was likely recycled to form part of the Cialancia orthogneiss. Geochronological evidence of this source was observed only in sample Y18, collected from a melanocratic, dark and restitic portion of the Cialancia orthogneiss. The age of the Cialancia orthogneiss (c. 440–458 Ma) indicates that its emplacement was possibly contemporaneous with that of the Ghiandone orthogneiss (457 ± 2 Ma, Bussy & Cadoppi, Reference Bussy and Cadoppi1996; 442.2 ± 2.0 Ma, Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022) and the Granero orthogneiss (c. 438–457 Ma; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024). We interpret these orthogneisses as three distinct magmatic suites, based on field observations and petrographic differences, despite the similarity in REE distribution (Fig. 7c) observed between the Granero and Ghiandone orthogneiss. These differences include the presence of relict pre-Alpine biotite, the complete replacement of K-feldspar and the occurrence of garnet in the Ghiandone orthogneiss (Section 4.4).

Zircons from the Giasset metadiorite do not yield clear age information (scatter dates between c. 600 and 300 Ma). However, a younger-than-Ordovician age can be tentatively proposed from field relationships, as the Ghiandone orthogneiss is crosscut by several metadiorite dykes (Fig. 4h). Notably, there is a minimal geochemical difference in REE composition between the Giasset and the Malanaggio metadiorite, hosted in the Pinerolo–Sanfront Unit (Fig. 7e). Further geochronological analyses are needed to constrain the absolute age of the Giasset metadiorite. The Luserna orthogneiss sample dated in this study (DD109) yielded a Permian photolith age of 265.3 ± 2.7 Ma, which is within uncertainty the same as previous estimates made further north on the Col Clapier outcrop (270.5 ± 2.7 Ma with full uncertainties; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024). Geochemically, the Luserna orthogneiss is quite heterogeneous, with REE distribution and abundance showing considerable variations (Fig. 7f), and thus possibly composed of non-contemporaneous magma batches. As a result, it could be challenging to distinguish it from other orthogneisses only on the basis of geochemical data. Finally, for the Freidour orthogneiss sample (Pinerolo–Sanfront Unit), we report a minimum crystallization age of 268.4 ± 2.7 Ma that is at the lower end of the range 268–283 Ma proposed by Bussy & Cadoppi (Reference Bussy and Cadoppi1996).

Our new geochronological data show that the Dora-Maira continental basement assembled granitoid rocks from three different magmatic cycles of Ediacaran–Early Cambrian, Ordovician–Silurian and Permian age. These data confirm the presence of different magmatic suites, as proposed by previous authors, with pre-Variscan meta-intrusive much more widespread than previously thought in the northern-central Dora-Maira (Sandrone et al., Reference Sandrone, Cordola and Fontan1986; Reference Sandrone, Sacchi, Cordola, Fontan and Villa1988; Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024), where Permian ages were considered prevalent. To date, it seems that the Brossasco and Sangone orthogneisses (e.g. Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Paquette et al., Reference Paquette, Montel and Chopin1999; Chen et al., Reference Chen, Zhou, Zheng and Schertl2017) are the only dated Permian felsic meta-intrusives hosted within the ‘Basement Complex’ Units. The Luserna orthogneiss is also part of the Permian magmatism, although its intrusive or volcanic origin remains debated (e.g. Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024).

7.b. Tectonic implications: paleogeographic and geodynamic evolution

Paleogeographic significance and tectonic evolution of orthogneisses in the pre-Mesozoic basements across southern Europe and the Alps have long been the subject of interest and controversies (Von Raumer et al., Reference Von Raumer, Stampfli, Borel and Bussy2002; Reference Von Raumer, Stampfli, Arenas and Sánchez Martínez2015; Bergomi et al., Reference Bergomi, Dal Piaz, Malusà, Monopoli and Tunesi2017; Siegesmund et al., Reference Siegesmund, Oriolo, Broge, Hueck, Lammerer, Basei and Schulz2023). The results obtained in this study are discussed below, divided by different time periods to outline the evolution of the Dora-Maira crystalline basement and its paleogeographic context (Fig. 10).

Figure 10. Summary of the pre-Alpine evolution of the Dora-Maira Massif comparing the ‘Basement Complex’ units with the Pinerolo–Sanfront unit, modified from Nosenzo et al. (Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022). Discontinuous lines indicate hypothetical time periods that are not yet constrained by geochronology. Data are from: 1 – Chen et al. (Reference Chen, Schertl, Chopin, Lin, Lin, Li, Lv and Nowlan2024); 2 – Bussy & Cadoppi (Reference Bussy and Cadoppi1996); 3 – Manzotti et al. (2025); 4 – Nosenzo et al. (Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022); 5 – Paquette et al. (Reference Paquette, Montel and Chopin1999); 6 – Chen et al. (Reference Chen, Zhou, Zheng and Schertl2017); 7 - Nosenzo et al. (Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024); 8 – Cadoppi & Tallone (Reference Cadoppi and Tallone1992), Sacchi et al. (Reference Sacchi, Balestro, Cadoppi, Carraro, Delle Piane, Di Martino, Enrietti, Gallarà, Gattiglio, Martinotti and Perello2004); 9 – Merlo & Malaroda (Reference Merlo and Malaroda1990); 10 – Manzotti et al. (Reference Manzotti, Ballèvre and Poujol2016); 11 – Balestro et al. (Reference Balestro, Festa, Cadoppi, Groppo and Roà2022); 12 – Bonnet et al. (Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022); 13 – This work.

7.b.1. Pre-Cambrian and Cambrian Cycle

The Ferrera and Giulian orthogneisses returned magmatic ages typical of the Cadomian orogeny. In addition, a Cadomian age is found in the inherited cores of one of the Cialancia orthogneiss samples (Y16). The occurrence of such ancient rocks in the ‘Basement Complex’ Units agrees with the detrital zircon populations identified by other studies in the northern Dora-Maira Massif basement (Manzotti et al., Reference Manzotti, Ballèvre and Poujol2016; Reference Manzotti, Millonig, Gerdes, Whitehouse, Jeon, Poujol and Ballèvre2025a; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022; Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024; Fig. 10), as well as with the U–Pb zircon dating of a Pre-Cambrian basement in the ‘Briançonnais Zone’ (Thiéblemont et al., Reference Thiéblemont, Jacob, Lach, Guerrot and Leguérinel2023). These data suggest that these rocks were assembled along the north-western margin of Gondwana during the Late Proterozoic to Cambrian period (Siegesmund et al., Reference Siegesmund, Oriolo, Schulz, Heinrichs, Basei and Lammerer2021; Reference Siegesmund, Oriolo, Broge, Hueck, Lammerer, Basei and Schulz2023).

Previous studies have demonstrated that the sedimentation in the Dora-Maira ‘Basement Complex’ Units during Ediacaran to Ordovician was predominantly siliciclastic (Fig. 10), except for possible carbonate deposition during the early Cambrian (‘Archaeocyathan limestones’; Debrenne et al., Reference Debrenne, Gandin and Pillola1989; Carmignani et al., Reference Carmignani, Carosi, Di Pisa, Gattiglio, Musumeci, Oggiano and Pertusati1994; Pillola et al., Reference Pillola, Leone and Loi1998; Debrenne, Reference Debrenne2007), and occurred in a marine environment (Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022). This feature is common in other Paleozoic basements in the Eastern-Central Alps (Liati et al., Reference Liati, Gebauer and Fanning2009; Heinrichs et al., Reference Heinrichs, Siegesmund, Frei, Drobe and Schulz2012; Mandl et al., Reference Mandl, Kurz, Hauzenberger, Fritz, Klötzli and Schuster2018) and Europe (Paoli et al., Reference Paoli, Stokke, Rocchi, Sirevaag, Ksienzyk, Jacobs and Košler2017; Cocco et al., Reference Cocco, Oggiano, Funedda, Loi and Casini2018; Pieruccioni et al., Reference Pieruccioni, Vezzoni and Petrelli2018). It is possible that the Cambrian metadiorite, dated by Balestro et al. (Reference Balestro, Festa, Cadoppi, Groppo and Roà2022, Fig. 1 and 2) south of our study area, was originally part of the ‘Basement Complex’ Units and was later detached and incorporated into the Valmala Shear Zone during Alpine deformation.

7.b.2. Ordovician–Silurian Magmatic Cycle

The protoliths of the Ghiandone, Cialancia and Granero (Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022; Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024) orthogneisses are associated with the Ordovician–Lower Silurian magmatic cycle (Fig. 10). Ordovician to Lower Silurian granitoids and coeval basic rocks with ages ranging from 480 to 420 Ma are well-documented throughout the Alpine belt (Bertrand & Leterrier, Reference Bertrand and Leterrier1997; Schaltegger & Gebauer, Reference Schaltegger and Gebauer1999; Neubauer, Reference Neubauer2002; Von Raumer et al., Reference Von Raumer, Stampfli, Borel and Bussy2002). Our data show that different generations of peraluminous calc-alkaline granites were emplaced in the Dora-Maira basement during the Middle-Late Ordovician to Early Silurian. The cross-cutting relationships (i.e. dykes), previously observed by Sandrone et al. (Reference Sandrone, Cordola and Fontan1986) between the Cialancia and Ghiandone orthogneisses, remain valid despite the similar ages of these two rock types. Because of the presence of MME in the Ghiandone orthogneiss, we suggest that the Ghiandone orthogneiss is a slightly older granite generated by a more primitive melt.

Granitic protoliths of Ordovician to Lower Silurian age have been documented in other Briançonnais-derived basement units in the Alps, including the Ruitor (459.0 ± 2.3 and 456.4 ± 2.4 Ma; Bergomi et al., Reference Bergomi, Dal Piaz, Malusà, Monopoli and Tunesi2017; 468 ± 22, 465 ± 11 and 460 ± 7 Ma; Guillot et al., Reference Guillot, Schaltegger, Bertrand, Deloule and Baudin2002), Sapey (452 ± 5 Ma; Bertrand et al., Reference Bertrand, Pidgeon, Leterrier, Guillot, Gasquet and Gattiglio2000a), Siviez-Mischabel (c. 460 Ma; Genier et al., Reference Genier, Epard, Bussy and Magna2008) and Ligurian Briançonnais (470–455 Ma; Gaggero et al., Reference Gaggero, Cortesogno and Bertrand2004) units. Metagranitoids of similar age and geochemistry are also documented in the Central-Eastern Alps (Flisch, Reference Flisch1987; Liati et al., Reference Liati, Gebauer and Fanning2009; Bussien et al., Reference Bussien, Bussy, Magna and Masson2011; Boriani et al., Reference Boriani, Bergomi, Ferrario, Migliacci Bellante and Vari2012; Galli et al., Reference Galli, Le Bayon, Schmidt, Burg, Reusser, Sergeev and Larionov2012; Cavargna-Sani et al., Reference Cavargna-Sani, Epard, Bussy and Ulianov2014), the Southern Alps (Pezzotta & Pinarelli, Reference Pezzotta and Pinarelli1994; Bergomi et al., Reference Bergomi, Colombo, Tunesi, Caironi and Boriani2004) and Sardinia (Carmignani et al., Reference Carmignani, Carosi, Di Pisa, Gattiglio, Musumeci, Oggiano and Pertusati1994; Cocco et al., Reference Cocco, Oggiano, Funedda, Loi and Casini2018).

The protoliths of the Ordovician–Early Silurian orthogneisses were intruded into pelitic and arenaceous sequences of Late Neoproterozoic to Cambrian age, dated through detrital zircon studies (Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022; Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024 for the Dora-Maira basement; Scheiber et al., Reference Scheiber, Berndt, Mezger and Pfiffner2014). Zurbriggen (Reference Zurbriggen2017) suggested that, during the Upper Ordovician to Silurian, the Cenerian orogenic system developed along the northern Gondwana margin (Zurbriggen, Reference Zurbriggen2017). This event is extensively preserved in the Austroalpine basement (e.g. Thöny et al., Reference Thöny, Tropper, Schennach, Krenn, Finger, Kaindl and Hoinkes2008; Siegesmund et al., Reference Siegesmund, Oriolo, Broge, Hueck, Lammerer, Basei and Schulz2023), where it also caused Silurian to Ordovician partial melting (around 444 Ma) (Klötzli-Chowanetz et al., Reference Klötzli-Chowanetz, Klötzli and Koller1997). Ordovician deformation between 494 and 467 Ma is also documented in the southernmost Briançonnais basements (Maino et al., Reference Maino, Gaggero, Langone, Seno and Fanning2019, and references). Therefore, it is likely that the ‘Basement Complex’ rocks were also affected by deformation and magmatic processes during this time. Part of the Ordovician–Silurian magmatic material could have resulted from partial melting of older pre-existing crustal material. Evidence for this possibility is provided by early Cambrian zircon cores in sample Y18, collected from a dark, likely more restitic portion of the Cialancia orthogneiss.

Ordovician orthogneisses in the Paleozoic basement of the Western Alps are closely associated with boudins of mafic rocks (eclogites) and massive to banded amphibolite sills (Desmons & Ploquin, Reference Desmons and Ploquin1989; Desmons, Reference Desmons1992; Desmons & Fabre, Reference Desmons and Fabre1988; Thélin et al., Reference Thélin, Sartori, Lengeler and Schaerer1990, Reference Thélin, Sartori, Burri, Gouffon and Chessex1993; Desmons & Mercier, Reference Desmons and Mercier1993; Guillot et al., Reference Guillot, Schaltegger, Bertrand, Deloule and Baudin2002; Sartori et al., Reference Sartori, Gouffon and Marthaler2006). While most of these mafic rocks remain undated, protolith ages of 468 ± 22 and 471 ± 5 Ma have been determined for amphibolites from the Ruitor Unit (Guillot et al., Reference Guillot, Schaltegger, Bertrand, Deloule and Baudin2002; Reference Bertrand, Paquette and Guillot2005), gabbroic sills from the Mont Fort basement (Gauthiez et al., Reference Gauthiez, Bussy, Ulianov, Gouffon and Sartori2011) and mafic rocks from the Ligurian Briançonnais basement (469 ± 6 Ma; Gaggero et al., Reference Gaggero, Cortesogno and Bertrand2004; Giacomini et al., Reference Giacomini, Braga, Tiepolo and Tribuzio2007; Cortesogno et al., Reference Cortesogno, Gaggero and Capelli1997). Mafic magmatism of Ordovician age could also occur in the Dora-Maira basement since polycyclic metabasites are commonly observed associated or intruded by the Ordovician Ghiandone orthogneiss (Compagnoni & Sandrone, Reference Compagnoni and Sandrone1981; Wheeler, Reference Wheeler1991), suggesting that these mafic rocks may be pre- or syn-Middle-Late Ordovician, as observed in other Paleozoic basements of the Alps (e.g. Gauthiez et al., Reference Gauthiez, Bussy, Ulianov, Gouffon and Sartori2011; Zurbriggen, Reference Zurbriggen2020).

7.b.3. The Variscan orogenic cycle

The end of the Cenerian orogeny was followed by the deposition of a sedimentary succession, likely ranging from the Silurian to the Late Devonian (Siegesmud et al., Reference Siegesmund, Oriolo, Broge, Hueck, Lammerer, Basei and Schulz2023). In the Dora-Maira basement, layers of marble (i.e. Rocca Bianca marbles, Vola et al., Reference Vola, Ardit, Frijia, Cavallo, Natali, Mion, Lugli and Primavori2022) intercalated within polycyclic micaschist, sometimes mined for talc (Sandrone et al., Reference Sandrone, Borghi, Carosso, Morsetti, Tagliano and Zucchetti1990; Cadoppi et al., Reference Cadoppi, Camanni, Balestro and Perrone2016), are typically interpreted as Late Devonian to Early Carboniferous in age (Fig. 10), as they contain relicts of high-temperature, pre-Alpine minerals (Castelli et al., Reference Castelli, Rolfo, Groppo and Compagnoni2007; Groppo et al., Reference Groppo, Castelli and Rolfo2007; Ferrando et al., Reference Ferrando, Groppo, Frezzotti, Castelli and Proyer2017). However, direct age determination of the protolith of these marbles is not available. Their formation during this period could be linked to the sedimentary response to the migration of Gondwana towards more equatorial positions (Franke et al., Reference Franke, Cocks and Torsvik2020; Scotese, Reference Scotese2021).

Syn-to-late-collisional Variscan magmatism appears to be absent in the Dora-Maira basement or at least has not yet been documented. However, the Dora-Maira Ediacaran to Devonian basement was involved in the Variscan orogeny, as evidenced by numerous relict Variscan mineral assemblages (e.g. Cadoppi, Reference Cadoppi1990; Bouffette et al., Reference Bouffette, Lardeaux and Caron1993; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022). P-T-t estimates by Nosenzo et al. (Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022) indicate Variscan prograde metamorphic conditions reaching up to 0.6 GPa and 650°C, with age determined by monazite petrochronology at 324 ± 6 Ma in the northern Dora-Maira (Fig. 10). Zircon overgrowths dated at 304 ± 2 Ma in the Muret Unit (Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022) record crystallization during cooling and decompression at the final stages of the Variscan orogenic cycle, likely involving fluid infiltration. Comparable results have been reported by Bonnet et al. (Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022), who identified pre-Alpine rutile and titanite in the southern Dora-Maira Massif, dated at 356 ± 35 Ma and c. 326 Ma, respectively. P–T conditions of 500–560°C and c. 0.6 GPa were inferred based on Zr thermometry (Bonnet et al. Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022).

7.b.4. Carboniferous and Permian

During the Late Carboniferous, endorheic basins developed on the future Briançonnais microcontinent basement (Fabre, Reference Fabre1958; Mercier & Beaudoin, Reference Mercier and Beaudoin1987; Manzotti et al., Reference Manzotti, Poujol and Ballèvre2015). Fluvio-lacustrine sediments deposited in these basins later formed the ‘Zone Houillère’ of the ‘Brianconnais Zone’, as well as the Pinerolo–Sanfront (Fig. 10) and Money Unit of the Internal Crystalline Massifs (Bertrand et al., Reference Bertrand, Aillères, Gasquet and Macaudiere1996; Manzotti et al., Reference Manzotti, Poujol and Ballèvre2015, Reference Manzotti, Ballèvre and Poujol2016). Notably, at this time, the same crust forming the ‘Basement Complex’ Units also served as the substrate for the Carboniferous sediments of the future Pinerolo–Sanfront Unit. It remains unclear whether this unit was detached at the level of these metasediments or if it also includes portions of older, pre-Carboniferous basement (Carosi et al., Reference Carosi, Montomoli, Iaccarino, Dana, Corno, De Cesari and Spina2025). Since Carboniferous sediments appear to be absent in the Dora-Maira Massif ‘Basement Complex’ Units, as well as in the pre-Triassic basement of several Briançonnais units (e.g. Ruitor, Vanoise), Gran Paradiso s.s. and Monte Rosa Units, it is suggested that these sectors were structural highs during the Late Carboniferous. Permian sedimentation within the ‘Basement Complex’ Units has been inferred by several authors (e.g. Cadoppi & Tallone, Reference Cadoppi and Tallone1992; Sacchi et al., Reference Sacchi, Balestro, Cadoppi, Carraro, Delle Piane, Di Martino, Enrietti, Gallarà, Gattiglio, Martinotti and Perello2004; Fig. 10) based on the occurrence of monocyclic garnet–chloritoid micaschist, locally associated with ankerite layers. These lithologies commonly occur in spatial association with the Permian Luserna orthogneiss or with remnants of the Mesozoic cover, supporting their attribution to late Paleozoic depositional settings. In the Pinerolo–Sanfront Unit, conglomeratic quartzite exposed in the Sanfront and Monte Bracco areas has been correlated with a Permian–Triassic siliciclastic sequence (Michard, Reference Michard1967; Michard et al., Reference Michard, Schmid, Lahfid, Ballèvre, Manzotti, Chopin, Iaccarino and Dana2022). In the Dora-Maira Massif, several generations of post-Variscan magmatic rocks have intruded the ‘Basement Complex’ Units and the Carboniferous metasediments of the Pinerolo–Sanfront Unit. Among these intrusions, the Brossasco metagranite (Gebauer et al., Reference Gebauer, Schertl, Brix and Schreyer1997; Paquette et al., Reference Paquette, Montel and Chopin1999; Chen et al., Reference Chen, Zhou, Zheng and Schertl2017) is potentially one of the oldest of these intrusions (with ages ranging from c. 304 to 260 Ma), followed by the Sangone orthogneiss (Fig. 10; Bussy & Cadoppi, Reference Cadoppi1990). These intrusions are associated with contact metamorphism, although such relicts are sparsely documented (pseudomorphs after cordierite, andalusite, sillimanite and spinel; Compagnoni & Rolfo, Reference Compagnoni and Rolfo2003; Groppo et al., Reference Groppo, Castelli and Rolfo2007; Castelli et al., Reference Castelli, Rolfo, Groppo and Compagnoni2007). The Luserna orthogneiss likely formed during this period (Fig. 10), but it remains unclear whether it represents volcanic or sub-volcanic rocks or shallow intrusions. Geochronological data indicate a transition from calc-alkaline magmatism to more mafic magmatism, exemplified by the monocyclic metabasites dated as Permian (250–255 Ma; Chen et al., Reference Chen, Schertl, Chopin, Lin, Lin, Li, Lv and Nowlan2024) in the Brossasco-Isasca Unit. These metabasites exhibit a pattern typical of Enriched Mid-Ocean Ridge Basalts or Ocean Island Basalts, in contrast to the Ordovician metabasites discussed above.

It remains unclear whether the intrusion of Giasset metadiorite bodies and the Castlüs orthogneiss into the ‘Basement Complex’ Units also occurred during the Permian or earlier. Vialon (Reference Vialon1966) reported the occurrence of migmatites in the Pellice Valley, associated with the Castlüs orthogneiss, suggesting that at least part of these rocks may have originated through partial melting of pre-existing crustal material. The Malanaggio metadiorite intruded the future Pinerolo–Sanfront Unit metasediments around 290 ± 2 and 288 ± 2 Ma (quartz-diorite and granodiorite facies respectively, Bussy & Cadoppi, Reference Bussy and Cadoppi1996), shortly followed by the emplacement of the large Freidour orthogneiss body (Fig. 10). Evidence for high-temperature, low-pressure contact metamorphism in the metasediments of the Pinerolo–Sanfront Unit, associated with Permian intrusions, is provided by pseudomorphs after andalusite (Franchi & Novarese, Reference Franchi and Novarese1895; Bussy & Cadoppi, Reference Bussy and Cadoppi1996) and relict titanite dated at c. 291 Ma, which crystallized at c. 600 °C assuming a pressure of c. 0.3 GPa (Bonnet et al., Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022).

The Permian magmatic event is widely acknowledged in the pre-Triassic basement of the South Alpine, Austroalpine and Salassic domains (Dal Piaz et al., Reference Dal Piaz, De Vecchi and Hunziker1977; Voshage et al., Reference Voshage, Hunziker, Hofmann and Zingg1987; Vavra et al., Reference Vavra, Gebauer, Schmidt and Compston1996, Reference Vavra, Schmid and Gebauer1999; Mayer et al., Reference Mayer, Mezger and Sinigoi2000; Schuster & Stüwe, Reference Schuster and Stüwe2008; Mohn et al., Reference Mohn, Manatschal, Müntener, Beltrando and Masini2010; Cenki-Tok et al., Reference Cenki-Tok, Oliot, Rubatto, Berger, Engi, Janots, Thomsen, Manzotti, Regis, Spandler, Robyr and Goncalves2011; Manzotti et al., Reference Manzotti, Rubatto, Darling, Zucali, Cenki-Tok and Engi2012, Reference Manzotti, Ballèvre and Dal Piaz2017; Petri et al., Reference Petri, Mohn, Skrzypek, Mateeva, Galster and Manatschal2017; Kunz et al., Reference Kunz, Manzotti, von Niederhäusern, Engi, Darling, Giuntoli and Lanari2018). Permian magmatism is also widespread in other Briançonnais-derived units: (i) Lower Permian porphyritic granite intrusions in Variscan high-grade paragneisses and migmatites in the Internal Crystalline Massifs (Bertrand et al., Reference Bertrand, Pidgeon, Leterrier, Guillot, Gasquet and Gattiglio2000b; Reference Bertrand, Paquette and Guillot2005; Liati et al., Reference Liati, Gervers, Froitzheim and Fanning2001; Pawlig & Baumgartner, Reference Pawlig and Baumgartner2001; Ring et al., Reference Ring, Collins and Kassem2005), (ii) Lower Permian orthogneisses in the ‘Zone Houillère’ and other Briançonnais basements (Bussy et al., Reference Bussy, Derron, Jacquod, Sartori and Thélin1996; Marquer et al., Reference Marquer, Challandes and Schaltegger1998; Scheiber et al., Reference Scheiber, Adrian Pfiffner and Schreurs2013, Reference Scheiber, Berndt, Mezger and Pfiffner2014; Bergomi et al., Reference Bergomi, Dal Piaz, Malusà, Monopoli and Tunesi2017) and (iii) widespread rhyolitic volcanism in the Classic Briançonnais units of the French and Ligurian Alps (286 ± 3 to 273 ± 2 Ma; Cortesogno et al., Reference Cortesogno, Oddone, Oxilia, Vanossi and Vannucci1982; Dallagiovanna et al., Reference Dallagiovanna, Gaggero, Maino, Seno and Tiepolo2009; Ballèvre et al., Reference Ballèvre, Camonin, Manzotti and Poujol2020). Metamorphism associated with the Permian magmatic event has been well-documented in the Alps (e.g. Hermann & Rubatto, Reference Hermann and Rubatto2003; Schuster et al., Reference Schuster, Scharbert, Abart and Frank2001; Schuster & Stüwe, Reference Schuster and Stüwe2008; Galli et al., Reference Galli, Le Bayon, Schmidt, Burg, Caddick and Reusser2011; Kunz et al., Reference Kunz, Johnson, White and Redler2014, Reference Kunz, Manzotti, von Niederhäusern, Engi, Darling, Giuntoli and Lanari2018). This metamorphic episode may also be partially recorded in the polycyclic rocks of the Dora-Maira Massif, where Bonnet et al. (Reference Bonnet, Chopin, Locatelli, Kylander-Clark and Hacker2022) documented relict rutile and titanite dated to c. 292 Ma and 238 ± 32 Ma, respectively, in the southern Dora-Maira. In this study, three of the analysed samples yield scattered Permian dates (Figs. 8 and 9), tentatively interpreted as evidence of a partial reopening of the U–Pb systematics during a Permian thermal event.

8. Conclusions

Field mapping, combined with zircon U–Pb dating and whole-rock geochemistry of orthogneisses from the northern-central Dora-Maira Massif, provides new insights into the Paleozoic evolution of the Variscan and pre-Variscan continental crust involved in the Alpine orogeny:

  1. 1. The pre-Triassic basement of the Dora-Maira Massif records a complex tectono-magmatic history, including Cambrian–Ordovician magmatism, the Early Carboniferous Variscan orogeny and Permian lithospheric thinning, in agreement with previous studies (Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022, Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024).

  2. 2. While an Ordovician to Silurian age for the Ghiandone orthogneiss has been previously established for the Punta Muret body (Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Nosenzo et al., Reference Nosenzo, Manzotti, Poujol, Ballèvre and Langlade2022) and for the Granero orthogneiss (Nosenzo et al., Reference Nosenzo, Manzotti, Krona, Ballèvre and Poujol2024), our study identifies additional Lower Paleozoic magmatic suites within the Dora-Maira basement. U–Pb zircon dating of the Cialancia orthogneiss (c. 440–455 Ma), previously interpreted as Permian (Bussy & Cadoppi, Reference Bussy and Cadoppi1996; Sandrone et al., Reference Sandrone, Cordola and Fontan1986), as well as the Ferrera orthogneiss (Ediacaran–Cambrian), reveals a more intricate magmatic evolution than previously recognized. Although the age of the Giasset metadiorite cannot be precisely constrained, field relationships suggest a post–Middle-Late Ordovician emplacement. A tectonic lens of Ediacaran–Cambrian orthogneiss has also been identified in the northern prolongation of the CLSZ, most likely derived from the ‘Basement Complex’ and incorporated in the shear zone during the Alpine deformation.

  3. 3. Field, geochemical and geochronological data from the Pinerolo–Sanfront Unit confirm a Permian magmatic framework evolving from calc-alkaline (Malanaggio metadiorite and Cavour orthogneiss) to alkaline affinity (Freidour orthogneiss), consistent with previous studies (Bussy & Cadoppi, Reference Bussy and Cadoppi1996).

  4. 4. The occurrence of an Ediacaran to Cambrian magmatic suite within the polycyclic basement rocks of the Dora-Maira Massif further supports a Gondwanan affinity for this basement, as proposed by previous authors (Bussy & Cadoppi, Reference Bussy and Cadoppi1996; von Raumer et al., Reference Von Raumer, Stampfli, Borel and Bussy2002, Reference Von Raumer, Bussy, Schaltegger, Schulz and Stampfli2013).

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S0016756825100332.

Acknowledgements

Thorsten Markmann is thanked for his assistance with LA-ICP-MS analyses; Simona Cavagna and Luca Pacchiega are thanked for their support with CL imaging; Gabriele Vola is acknowledged for providing the database of ‘historical’ geochemical analyses. We would like to thank Stefano Dolce, Chiara Scabin, Simone Lenci and Andrea Bigongiari for their valuable help in the field. We gratefully acknowledge the financial support provided by ISPRA and Regione Piemonte. Michel Ballèvre and Guillaume Bonnet are thanked for their engaging reviews and numerous constructive suggestions, which significantly improved this work. We also thank the editor, Olivier Lacombe, for his careful handling of our manuscript.

Author contributions

Davide Dana: Writing – original draft, Investigation, Methodology, Formal analysis, Data curation, Conceptualization. Francesco De Cesari: Writing – review & editing, Investigation, Methodology, Formal analysis, Data curation, Conceptualization. Chiara Montomoli: Writing – review & editing, Methodology, Investigation, Formal analysis, Data curation, Validation, Supervision, Conceptualization. Salvatore Iaccarino: Writing – review & editing, Methodology, Investigation, Formal analysis, Data curation, Validation, Supervision, Conceptualization. Alberto Corno: Writing – review & editing, Investigation, Formal analysis, Data curation. Daniela Rubatto: Writing – original draft, Investigation, Methodology, Formal analysis, Data curation, Conceptualization. Rodolfo Carosi: Writing – review & editing, Supervision, Investigation, Formal analysis, Conceptualization, Funding acquisition, Validation, Supervision, Project administration.

Financial support

This work was funded by the Geological Map of Italy CARG project – Sheet 172 – ‘Pinerolo’ (Resp. Prof. R. Carosi).

Competing interests

The authors declare that they have no competing interests.

References

Agard, P and Handy, M (2021) Ocean subduction dynamics in the Alps. Elements 17(1), 916. https://doi.org/10.2138/gselements.17.1.9 CrossRefGoogle Scholar
Balestro, G, Cadoppi, P, Di Martino, L and Sacchi, R (1995) Il settore meridionale del Massiccio Dora-Maira (Valli Maira e Varaita): Inquadramento, carta geologica e guida a un’escursione. Rapporti Alpi-Appennino e guida alle escursioni. Accademia Nazionale Delle Scienze 14, 501–29.Google Scholar
Balestro, G, Festa, A, Cadoppi, P, Groppo, C and Roà, M (2022) Pre-orogenic tectonostratigraphic evolution of the European distal margin-Alpine Tethys transition zone in high-pressure units of the southwestern Alps. Geosciences 12(10), 358.CrossRefGoogle Scholar
Balestro, G, Fioraso, G and Lombardo, B (2011) Geological map of the upper Pellice valley (Italian Western Alps). Journal of Maps 7(1), 634–54.CrossRefGoogle Scholar
Balestro, G, Gattiglio, M, Roà, M and Festa, A (2025) Tectonic relationships between the Paleozoic basement and the Mesozoic cover succession in the northern Dora-Maira Massif (Susa Valley, Western Alps). Journal of Maps 21(1), 2451785.CrossRefGoogle Scholar
Balestro, G, Nosenzo, F, Cadoppi, P, Fioraso, G, Groppo, C and Festa, A (2020) Geology of the southern Dora-Maira Massif: Insights from a sector with mixed ophiolitic and continental rocks (Valmala tectonic unit, Western Alps). Journal of Maps 16(2), 736–44.CrossRefGoogle Scholar
Ballèvre, M, Camonin, A, Manzotti, P and Poujol, M (2020) A step towards unraveling the paleogeographic attribution of pre-Mesozoic basement complexes in the Western Alps based on U–Pb geochronology of Permian magmatism. Swiss Journal of Geoscience 113, 12. https://doi.org/10.1186/s00015-020-00367-1 CrossRefGoogle Scholar
Ballèvre, M, Fourcade, S, Capdevila, R, Peucat, JJ, Cocherie, A and Fanning, C M (2012) Geochronology and geochemistry of Ordovician felsic volcanism in the Southern Armorican Massif (Variscan belt, France): Implications for the breakup of Gondwana. Gondwana Research 21(4), 1019–36.CrossRefGoogle Scholar
Ballèvre, M, Manzotti, P and Dal Piaz, GV (2018) Pre-Alpine (Variscan) inheritance: a key for the location of the future Valaisan Basin (Western Alps). Tectonics 37(3), 786817. https://doi.org/10.1002/2017TC004633 CrossRefGoogle Scholar
Barisone, G, Bottino, G, Coccolo, V, Compagnoni, R, Del Greco, O, Mastrangelo, F, Sandrone, R & Zucchetti, S (1979 ) Il bacino estrattivo della “Pietra di Luserna” (Alpi Cozie). Notiziario dell’Associazione Mineraria Subalpina 5, 3550.Google Scholar
Bellanger, M, Bellahsen, N, Jolivet, L, Baudin, T, Augier, R and Boutoux, A (2014) Basement shear zones development and shortening kinematics in the Ecrins Massif, Western Alps. Tectonics 33(2), 84111.CrossRefGoogle Scholar
Beltrando, M., Compagnoni, R. and Lombardo, B. (2010) (Ultra-) High-pressure metamorphism and orogenesis: An Alpine perspective. Gondwana Research 18, 147–66. https://doi.org/10.1016/j.gr.2010.01.009 CrossRefGoogle Scholar
Bergomi, MA, Colombo, A, Tunesi, A, Caironi, V and Boriani, A (2004) Ordovician granitoids in the Southern Alps basement (Monte Fioraro Complex-Central Alps): geochemistry and SHRIMP U-Pb zircon age. Paper presented at 32nd IGC conference, Firenze, IT. Abstract volume, 2, 1062.Google Scholar
Bergomi, MA, Dal Piaz, GV, Malusà, MG, Monopoli, B and Tunesi, A (2017) The Grand St Bernard-Briançonnais nappe system and the Paleozoic inheritance of the Western Alps unraveled by zircon U-Pb dating. Tectonics 36(12), 2950–72.CrossRefGoogle Scholar
Bertrand, JM, Aillères, L, Gasquet, D and Macaudiere, J (1996) The Pennine Front zone in Savoie (Western Alps), a review and new interpretations from the Zone Houillère Briançonnaise. Eclogae Geologicae Helvetiae 89(1), 297320.Google Scholar
Bertrand, JM, Guillot, F and Leterrier, J (2000a) Âge Paléozoïque inférieur (U—Pb sur zircon) de métagranophyres de la nappe du Grand-Saint-Bernard (Zona Interna, France, Vallée d’Aoste, Italie). Comptes Rendus de l’Académie des Sciences-Series IIA-Earth and Planetary Science 330(7), 473–78.Google Scholar
Bertrand, JM and Leterrier, J (1997) Granitoïdes d’âge Paléozoïque inférieur dans le socle de Vanoise méridionale: géochronologie U-Pb du métagranite de l’Arpont France de Savoie, France). Comptes Rendus de l’Académie des Sciences-Series IIA-Earth and Planetary Science 325(11), 839–44.Google Scholar
Bertrand, JM, Paquette, JL and Guillot, F (2005) Permian zircon U-Pb ages in the Gran Paradiso massif: revisiting post-Variscan events in the Western Alps. Schweizerische Mineralogische und Petrographische Mitteilungen 85(1), 1529.Google Scholar
Bertrand, JM, Pidgeon, RT, Leterrier, J, Guillot, F, Gasquet, G and Gattiglio, M (2000b) SHRIMP and IDTIMS U-Pb zircon ages of the pre-Alpine basement in the Internal Western Alps (Savoy and Piemont). Schweizerische Mineralogische und Petrographische Mitteilungen 80, 225248.Google Scholar
Biino, GG and Compagnoni, R (1992) Very-high pressure metamorphism of the Brossasco coronite metagranite, southern Dora Maira Massif, Western Alps. Schweizerische mineralogische und petrographische Mitteilungen 72(3), 347–63.Google Scholar
Black, LP, Kamo, SL, Allen, CM, Aleinikoff, JN, Davis, DW, Korsch, RJ and Foudoulis, C (2003) TEMORA 1: a new zircon standard for Phanerozoic U–Pb geochronology. Chemical geology 200(1-2), 155–70.CrossRefGoogle Scholar
Bonin, B (1988) From orogenic to anorogenic environments: evidence from associated magmatic episodes. Schweizerische Mineralogische und Petrographische Mitteilungen 68(3), 301–11.Google Scholar
Bonin, B, Brändlein, P, Bussy, F, Desmons, J, Eggenberger, U, Finger, F, Graf, K, Marro, C, Mercolli, F, Oberhänsli, R, Ploquin, A, von Quadt, A, von Raumer, J, Schaltegger, U, Steyrer, HP, Visonà, D & Vivier, G (1993) Late Variscan magmatic evolution of the Alpine basement. In Pre-Mesozoic geology in the Alps (pp. 171201). Berlin, Heidelberg: Springer Berlin Heidelberg.CrossRefGoogle Scholar
Bonioli, L, Cadoppi, P and Sacchi, R (1992) Occurrence of metavolcanics in the southernmost Dora-Maira Massif (Italian Western Alps). In Contributions to the geology of Italy with special regard to the Paleozoic basements. A volume dedicated to Tommaso Cocozza (eds Carmignani, L., & Sassi, F. P. ), vol. 5, pp. 237–40. Council of the International Geoscience Programme, 276, Newsletter.Google Scholar
Bonnet, G, Chopin, C, Locatelli, M, Kylander-Clark, ARC and Hacker, BR (2022) Protracted subduction of the European hyperextended margin revealed by rutile U-Pb geochronology across the Dora-Maira massif (western Alps). Tectonics 41, e2021TC007170. https://doi.org/10.1029/2021TC007170 CrossRefGoogle Scholar
Borghi, A (1983) Studio geologico della media Val Chisone (versante sinistro). Published Thesis. University of Turin.Google Scholar
Borghi, A, Cadoppi, P and Dino, GA (2016) Heritage stone 2. The Dora-Maira Unit (Italian Cottian Alps): A reservoir of ornamental stones since Roman times. Geoscience Canada 43, 1330.CrossRefGoogle Scholar
Borghi, A, Cadoppi, P, Poli, G, Sacchi, R, Sandrone, R (1989) Metaintrusives in the north-central part of the Dora-Maira Massif (Western Alps). Periodico di Mineralogia 58, 171–72.Google Scholar
Borghi, A, Cadoppi, P, Porro, A and Sacchi, R (1985) Metamorphism in the northern part of the Dora Maira massif (Cottian Alps). Bollettino / Museo Regionale di Scienze Naturali 3, 369–80.Google Scholar
Borghi, A, Cadoppi, P, Porro, A, Sacchi, R and Sandrone, R (1984) Osservazioni geologiche nella Valle Germanasca e nella media Val Chisone (Alpi Cozie). Bollettino del Museo Regionale di Scienza Naturali, Torino 2, 503–30.Google Scholar
Boriani, A, Bergomi, MA, Ferrario, A and Migliacci Bellante, R (2012) Basamento metamorfico austroalpino. In Note Illustrative della Carta Geologica d’Italia alla scala 1:50.000, Foglio 056 Sondrio (ed Vari, Autori), pp. 4969. ISPRA.Google Scholar
Bortolami, G and Dal Piaz, GV (1970) Il substrato cristallino dell’anfiteatro morenico di Rivoli-Avigliana (Provincia di Torino). Memorie della Società italiana di Scienze Naturali 18, 125–69.Google Scholar
Boschetti, L, Mouthereau, F, Schwartz, S, Rolland, Y, Bernet, M, Balvay, M, … & Lahfid, A (2025) Thermochronology of the western Alps (Pelvoux massif) reveals the longterm multiphase tectonic history of the European paleomargin. Tectonics 44(2), e2024TC008498.CrossRefGoogle Scholar
Bouffette, J, Lardeaux, JM and Caron, JM (1993) Le passage des granulites aux éclogites dans les métapélites de l’unité de la Punta Muret (Massif Dora-Maira, Alpes occidentales). Comptes rendus de l’Académie des sciences. Série 2, Mécanique, Physique, Chimie, Sciences de l’univers, Sciences de la Terre 317(12), 1617–24.Google Scholar
Bousquet, R, Oberhänsli, R, Goffé, B, Wiederker, M, Koller, F, Schmid, SM, Schuster, R, Engi, M, Berger, A and Martinotti, G (2008) Metamorphism of metasediments at the scale of an orogen: a key to the Tertiary geodynamic evolution of the Alps. Geological Society of London Special Publications 298, 393411. https://doi.org/10.1144/SP298.18 CrossRefGoogle Scholar
Bruno, M, Compagnoni, R and Rubbo, M (2001) The ultra-high pressure coronitic and pseudomorphous reactions in a metagranodiorite from the Brossasco-Isasca Unit, Dora-Maira massif, Western Italian Alps: A petrographic study and equilibrium thermodynamic modelling. Journal of Metamorphic Geology 19, 3343.CrossRefGoogle Scholar
Brunsmann, Q, Rosenberg, CL and Bellahsen, N (2024) The Western Alpine arc: a review and new kinematic model. Comptes Rendus Géoscience 356(S2), 231–63.CrossRefGoogle Scholar
Bussien, D., Bussy, F., Magna, T. and Masson, H. (2011) Timing of Palaeozoic magmatism in the Maggia and Sambuco nappes and paleogeographic implications (Central Lepontine Alps). Swiss Journal of Geosciences 104, 129.CrossRefGoogle Scholar
Bussy, F and Cadoppi, P (1996) U-Pb zircon dating of granitoids from the Dora-Maira massif (western Italian Alps). Schweizerische Mineralogische und Petrographische Mitteilungen 76, 217–33.Google Scholar
Bussy, F, Derron, MH, Jacquod, J, Sartori, M and Thélin, P (1996) The 500 Ma-old Thyon metagranite: A new A-type granite occurrence in the western Penninic Alps (Wallis, Switzerland). European Journal of Mineralogy 8, 565–75.CrossRefGoogle Scholar
Cadoppi, P (1990) Geologia del basamento cristallino nel settore settentrionale del Massiccio Dora-Maira (Alpi Occidentali). PhD thesis, University of Turin, Turin, Italy. Published thesis.Google Scholar
Cadoppi, P, Camanni, G, Balestro, G and Perrone, G (2016) Geology of the Fontane talc mineralization (Germanasca valley, Italian Western Alps). Journal of Maps 12(5), 1170–77.CrossRefGoogle Scholar
Cadoppi, P., Castelletto, M., Sacchi, R., Baggio, P., Carraro, F. and Giraud, V. (2002) Note illustrative della Carta Geologica d’Italia alla scala 1: 50.000–Foglio 154, Susa, pp. 1123. Servizio Geologico d’Italia.Google Scholar
Cadoppi, P and Tallone, S (1992) Structural and lithostratigraphic heterogeneity of the northern part of the Dora-Maira Massif (Western Alps). Atti Ticinensi di Scienze della Terra 35, 918.Google Scholar
Carmignani, L, Carosi, R, Di Pisa, A, Gattiglio, M, Musumeci, G, Oggiano, G and Pertusati, CP (1994) The Hercynian chain in Sardinia (Italy). Geodinamica Acta 7(1), 3147.CrossRefGoogle Scholar
Caron, JM (1977) Lithostratigraphie et tectonique des Schistes lust& dans les Alpes cottiennes septentrionales et en Corse orientale. Thesis, Univ. L. Pasteur de Strasbourg, Strasbourg, 326 pp.Google Scholar
Carosi, R, Montomoli, C, Iaccarino, S, Dana, D, Corno, A, De Cesari, F and Spina, A (2025) Buried, not erased: palynofloras in ultra-high-pressure metamorphic rocks. Scientific Reports 15(1), 35865.CrossRefGoogle Scholar
Carosi, R, Montomoli, C, Tiepolo, M and Frassi, C (2012) Geochronological constraints on post-collisional shear zones in the Variscides of Sardinia (Italy). Terra Nova 24(1), 4251.CrossRefGoogle Scholar
Castelli, D, Rolfo, F, Groppo, C and Compagnoni, R (2007) Impure marbles from the UHP Brossasco-Isasca Unit (Dora-Maira Massif, western Alps): evidence for Alpine equilibration in the diamond stability field and evaluation of the X (CO2) fluid evolution. Journal of Metamorphic Geology 25(6), 587603.CrossRefGoogle Scholar
Cavargna-Sani, M, Epard, JL, Bussy, F and Ulianov, A (2014) Basement lithostratigraphy of the Adula nappe: implications for Palaeozoic evolution and Alpine kinematics. International Journal of Earth Sciences 103, 6182.CrossRefGoogle Scholar
Cawood, PA, Hawkesworth, CJ and Dhuime, B (2013) The continental record and the generation of continental crust. Geological Society of America Bulletin 125(1-2), 1432.CrossRefGoogle Scholar
Cenki-Tok, B, Oliot, E, Rubatto, D, Berger, A, Engi, M, Janots, E, Thomsen, TB, Manzotti, P, Regis, D, Spandler, C, Robyr, M and Goncalves, P (2011) Preservation of Permian allanite within an Alpine eclogite facies shear zone at Mt Mucrone, Italy: Mechanical and chemical behavior of allanite during mylonitization. Lithos 125, 4050. https://doi.org/10.1016/j.lithos.2011.01.005 CrossRefGoogle Scholar
Ceriani, S, Fügenschuh, B and Schmid, SM (2001) “Multi-stage thrusting at the” Penninic Front” in the Western Alps between Mont Blanc and Pelvoux massifs. International Journal of Earth Sciences 90(3), 685702.CrossRefGoogle Scholar
Chen, X, Schertl, HP, Chopin, C, Lin, C, Lin, H, Li, H, Lv, J and Nowlan, EU (2024) From divergent to convergent plate boundary: A ca. 200 Ma Wilson cycle recorded by ultrahigh-pressure eclogites in the Dora-Maira Massif, Western Alps. GSA Bulletin 136(5-6), 2178–94.Google Scholar
Chen, YX, Schertl, HP, Zheng, YF, Huang, F, Zhou, K and Gong, YZ (2016) Mg–O isotopes trace the origin of Mg-rich fluids in the deeply subducted continental crust of Western Alps. Earth and Planetary Science Letters 456, 157–67.CrossRefGoogle Scholar
Chen, YX, Zhou, K, Zheng, YF and Schertl, HP (2017) Zircon geochemical constraints on the protolith nature and metasomatic process of the Mg-rich whiteschist from the Western Alps. Chemical Geology 467, 177–95.CrossRefGoogle Scholar
Chopin, C (1984) Coesite and pure pyrope in high-grade blueschists of the Western Alps: a first record and some consequences. Contribution to Mineralogy and Petrology 86, 107–18.CrossRefGoogle Scholar
Chopin, C, Henry, C and Michard, A (1991) Geology and petrology of the coesite-bearing terrain, Dora Maira massif, Western Alps. European Journal of Mineralogy 3(2), 263–91.CrossRefGoogle Scholar
Cocco, F, Oggiano, G, Funedda, A, Loi, A and Casini, L (2018) Stratigraphic, magmatic and structural features of Ordovician tectonics in Sardinia (Italy): a review. Journal of Iberian Geology 44, 619–39.CrossRefGoogle Scholar
Compagnoni, R (1977) The Sesia-Lanzo Zone, a slice of continental crust with Alpine high-pressure-low temperature assemblages in the western Italian Alps. Rendiconti Societa Italiana di Mineralogia e Petrologia 33, 281334.Google Scholar
Compagnoni, R, Crisci, GM and Sandrone, R (1983) Caratterizzazione chimica e petrografica degli “gneiss di Luserna” (Massiccio cristallino Dora-Maira, Alpi Occidentali). Rendiconti Società Italiana Mineralogia e Petrologia 38, 498.Google Scholar
Compagnoni, R and Rolfo, F (2003) UHPM units in the Western Alps. Ultrahigh Pressure Metamorphism 5, 1349.CrossRefGoogle Scholar
Compagnoni, R, Rolfo, F, Groppo, C, Hirajima, T and Turello, R (2012) Geological map of the ultra-high pressure Brossasco-Isasca unit (Western Alps, Italy). Journal of Maps 8(4), 465–72.CrossRefGoogle Scholar
Compagnoni, R and Sandrone, R (1981) Lineamenti geo-petrografici delle Alpi Cozie Italiane tra la Val di Susa e la Valle Po. Associazione piemontese di mineralogia e paleontologia F. Meda.Google Scholar
Condie, KC (2014) Growth of continental crust: a balance between preservation and recycling. Mineralogical Magazine 78(3), 623–37.CrossRefGoogle Scholar
Corfu, F, Hanchar, JM, Hoskin, PW and Kinny, P (2003) Atlas of zircon textures. Reviews in Mineralogy and Geochemistry 53(1), 469500.CrossRefGoogle Scholar
Corsini, M and Rolland, Y (2009) Late evolution of the southern European Variscan belt: Exhumation of the lower crust in a context of oblique convergence. Comptes Rendus. Géoscience 341(2-3), 214–23.CrossRefGoogle Scholar
Cortesogno, L, Gaggero, L and Capelli, C (1997) Petrology of pre-Alpine eclogites and amphibolites from the Ligurian Briançonnais basement. Atti Ticinensi Scienze della Terra 39, 329.Google Scholar
Cortesogno, L, Oddone, M, Oxilia, M, Vanossi, M and Vannucci, R (1982) Le metavulcaniti a chimismo andesitico del perno-carbonifero brianzonese (Alpi Marittime): caratterizzazione petrografica e tentativo di interpretazione geodinamica. Rendiconti della Societa Italiana di Mineralogia e Petrologia 38(2), 581606.Google Scholar
Couzinié, S, Laurent, O, Poujol, M, Mintrone, M, Chelle-Michou, C, Moyen, JF, … and Marko, L (2017) Cadomian S-type granites as basement rocks of the Variscan belt (Massif Central, France): Implications for the crustal evolution of the north Gondwana margin. Lithos 286, 1634.CrossRefGoogle Scholar
Dal Piaz, GV (1993) Evolution of Austro-Alpine and Upper Penninic basement in the northwestern Alps from Variscan convergence to post-Variscan extension. In Pre-Mesozoic geology in the Alps (pp. 327–44). Berlin, Heidelberg: Springer Berlin Heidelberg.CrossRefGoogle Scholar
Dal Piaz, GV (2010). The Italian Alps: a journey across two centuries of Alpine geology. Journal of the Virtual Explorer 36. https://doi.org/10.3809/jvirtex.2010.00234 CrossRefGoogle Scholar
Dal Piaz, GV, Bistacchi, A and Massironi, M (2003) Geological outline of the Alps. Episodes XXVI(3), 175–80.CrossRefGoogle Scholar
Dal Piaz, GV., De Vecchi, G and Hunziker, JC (1977) The Austroalpine layered gabbros of the Matterhorn and Mt Collon-Dents de Bertol. Schweizerische Mineralogische und Petrographische Mitteilungen 57, 5988.Google Scholar
Dal Piaz, GV and Martin, S (1998). Evoluzione litosferica e magmatismo nel dominio austro-sudalpino dall’orogenesi varisica al rifting permo-mesozoico. Memorie della Società Geologica Italiana 53, 4362.Google Scholar
Dallagiovanna, G, Gaggero, L, Maino, M, Seno, S and Tiepolo, M (2009) U–Pb zircon ages for post-Variscan volcanism in the Ligurian Alps (Northern Italy). Journal of the Geological Society 166(1), 101–14.CrossRefGoogle Scholar
Dana, D (2024) The Briançonnais margin units in the South-Western Alps (Italy-France). Geological Field Trips & Maps 16(2.2), 161.CrossRefGoogle Scholar
Dana, D, Iaccarino, S, Schmid, SM, Petroccia, A and Michard, A (2023) Structural and metamorphic evolution of a subducted passive margin: insights from the Briançonnais nappes of the Western Alps (Ubaye–Maira valleys, France–Italy). Swiss Journal of Geosciences 116(1), 18.CrossRefGoogle Scholar
De Broucker, G, Siméon, Y, Stampfli, G, Thiéblemont, D, Lach, P and Marthaler, M (2021) Early Cretaceous accretionary complex of the Valaisan Ocean, Western Alps?. Ofioliti 46(2), 203227.Google Scholar
De la Roche, HD, Leterrier, JT, Grandclaude, P and Marchal, M (1980) A classification of volcanic and plutonic rocks using R1R2-diagram and major-element analyses—its relationships with current nomenclature. Chemical Geology 29(1-4), 183210.CrossRefGoogle Scholar
Debrenne, F (2007) Lower Cambrian archaeocyathan bioconstructions. Comptes Rendus Palevol 6(1-2), 519.CrossRefGoogle Scholar
Debrenne, F, Gandin, A and Pillola, GL (1989) Biostratigraphy and depositional setting of Punta Manna member type-section (Nebida formation, lower Cambrian, SW Sardina, Italy). Rivista Italiana di Paleontologia e Stratigrafia 94(4), 483514.Google Scholar
Desmons, J. (1992) The Briançon basement (Pennine Western Alps): mineral composition and polymetamorphic evolution. Schweizerische Mineralogische und Petrographische Mitteilungen 72(1), 3755.Google Scholar
Desmons, J and Fabre, J (1988) Contribution à la connaissance pétrographique du Mont Pourri (Savoie, France); conséquences structurales. Géologie Alpine 64, 1326.Google Scholar
Desmons, J and Mercier, D (1993) Passing through the Briançon zone. In Pre-Mesozoic geology in the Alps, pp. 279–95. Berlin, Heidelberg: Springer Berlin Heidelberg.CrossRefGoogle Scholar
Desmons, J and Ploquin, A (1989) Chimisme du Brianconnais–Grand Saint-Bernard Ante-Mesozoique (Alpes occidentales): une vue synoptique. Géologie Alpine 65, 131.Google Scholar
Dini, A, Innocenti, F, Rocchi, S, Tonarini, S and Westerman, DS (2002) The magmatic evolution of the late Miocene laccolith–pluton–dyke granitic complex of Elba Island, Italy. Geological Magazine 139(3), 257–79.CrossRefGoogle Scholar
Dumont, T, Schwartz, S, Guillot, S, Malusà, M, Jouvent, M, Monié, P and Verly, A (2022) Cross-propagation of the western Alpine orogen from early to late deformation stages: Evidence from the Internal Zones and implications for restoration. Earth-Science Reviews 104106. https://doi.org/10.1016/j.earscirev.2022.104106 CrossRefGoogle Scholar
Fabre, J (1958) Contribution à l’étude de la zone Houillère en Maurienne et en Tarentaise (Alpes de Savoie). PhD thesis, Faculté des Sciences de l’Université de Paris, Université de Paris, Paris, France. Published thesis.Google Scholar
Ferrando, S, Groppo, C, Frezzotti, ML, Castelli, D and Proyer, A (2017) Dissolving dolomite in a stable UHP mineral assemblage: Evidence from Cal-Dol marbles of the Dora-Maira Massif (Italian Western Alps). American Mineralogist 102(1), 4260.CrossRefGoogle Scholar
Filippi, M., Jouffray, F., Lardeaux, J. M., Tiepolo, M. and Spalla, M. I. (2024) On the occurrence of a Variscan eclogite in the Argentera-Mercantour Massif, Western Alps: Implications for the evolution of the southern Variscan belt. Geological Society of America Bulletin 136(9-10), 4018–34.CrossRefGoogle Scholar
Finger, F and Riegler, G (2023) The role of the proto-Alpine Cenerian Orogen in the Avalonian-Cadomian belt. Austrian Journal of Earth Sciences 116(1), 109–15.CrossRefGoogle Scholar
Flisch, M (1987) Teil 1: Geologische, petrographische und isotopengeologische Untersuchungen an Gesteinen des Silvrettakristallins, Teil 2: Die Hebungsgeschichte der oberostalpinen Silvretta-Decke seit der mittleren Kreide. PhD thesis, University of Bern, Bern, Switzerland. Published thesis.Google Scholar
Franchi, S (1898) Sull’età mesozoica della zona delle pietre verdi nelle Alpi Occidentali. Bollettino del Regio Comitato Geologico d’Italia 29, 173247 and 325-482.Google Scholar
Franchi, S. and Novarese, V. (1895) Appunti geologici e petrografici sui dintorni di Pinerolo. Bollettino del Reale Comitato Geologico d’Italia, XXVI 385429.Google Scholar
Franke, W, Cocks, LRM and Torsvik, TH (2020) Detrital zircons and the interpretation of palaeogeography, with the Variscan Orogeny as an example. Geological Magazine 157(4), 690–94.CrossRefGoogle Scholar
Frisch, W (1979) Tectonic progradation and plate tectonic evolution of the Alps. Tectonophysics 60(3-4), 121–39.CrossRefGoogle Scholar
Froitzheim, N (2001) Origin of the Monte Rosa nappe in the Pennine Alps—A new working hypothesis. Geological Society of America Bulletin 113(5), 604–14.2.0.CO;2>CrossRefGoogle Scholar
Frost, BR and Frost, CD (2008) A geochemical classification for feldspathic igneous rocks. Journal of Petrology 49(11), 1955–69.CrossRefGoogle Scholar
Gaggero, L, Cortesogno, L and Bertrand, JM (2004) The Pre-Namurian basement of the Ligurian Alps: A review of the lithostratigraphy, pre-Alpine metamorphic evolution and regional comparisons. Periodico di Mineralogia 73, 8596.Google Scholar
Galli, A, Le Bayon, B, Schmidt, MW, Burg, JP, Caddick, MJ and Reusser, E (2011) Granulites and charnockites of the Gruf Complex: evidence for Permian ultra-high temperature metamorphism in the Central Alps. Lithos 124(1-2), 1745.CrossRefGoogle Scholar
Galli, A, Le Bayon, B, Schmidt, MW, Burg, JP, Reusser, E, Sergeev, SA and Larionov, A (2012) U–Pb zircon dating of the Gruf Complex: disclosing the late Variscan granulitic lower crust of Europe stranded in the Central Alps. Contributions to Mineralogy and Petrology 163, 353–78.CrossRefGoogle Scholar
Gasco, I, Gattiglio, M and Borghi, A (2011) Lithostratigraphic setting and PT metamorphic evolution for the Dora Maira Massif along the Piedmont Zone boundary (middle Susa Valley, NW Alps). International Journal of Earth Sciences 100, 1065–85.CrossRefGoogle Scholar
Gasco, I, Gattiglio, M and Borghi, A (2013) Review of metamorphic and kinematic data from Internal Crystalline Massifs (Western Alps): PTt paths and exhumation history. Journal of Geodynamics 63, 119.CrossRefGoogle Scholar
Gauthiez, L, Bussy, F, Ulianov, A, Gouffon, Y and Sartori, M (2011) Ordovician mafic magmatism in the Métailler Formation of the Mont-Fort nappe (Middle Penninic domain, western Alps)—Geodynamic implications. 9th Swiss Geoscience Meeting, Zurich, CH.Google Scholar
Gauthiez-Putallaz, L, Rubatto, D, Hermann, J (2016) Dating prograde fluid pulses during subduction by in situ U-Pb and oxygen isotope analysis. Contributions to Mineralogy and Petrology 171, 15.CrossRefGoogle Scholar
Gebauer, D, Schertl, HPS, Brix, HPM and Schreyer, W (1997) 35 Ma old ultrahigh-pressure metamorphism and evidence for very rapid exhumation in the Dora Maira Massif, Western Alps. Lithos 41(1-3), 524.CrossRefGoogle Scholar
Gebauer, D (1993) The pre-Alpine evolution of the continental crust of the Central Alps - An overview. In Pre-Mesozoic Geology in the Alps (ed von Raumer J.F., NF), pp. 317–25. Berlin: Springer. https://doi.org/10.1007/978-3-642-84640-3_18 Google Scholar
Genier, F, Epard, JL, Bussy, F. and Magna, T (2008) Lithostratigraphy and U-Pb zircon dating in the overturned limb of the Siviez-Mischabel nappe: a new key for Middle Penninic nappe geometry. Swiss Journal of Geosciences 101, 431–52.CrossRefGoogle Scholar
Giacomini, F, Braga, R, Tiepolo, M and Tribuzio, R (2007) New constraints on the origin and age of Variscan eclogitic rocks (Ligurian Alps, Italy). Contributions to Mineralogy and Petrology 153, 2953.CrossRefGoogle Scholar
Gouffon, Y, Bernoulli, D, Dall’agnolo, S, Fantoni, R, Jordan, P, Madritsch, H, Mosar, J, Pfiffner, OA, Picotti, V and Schenker, FL (2024a) Tectonic Map of Switzerland. Swisstopo, Switzerland.Google Scholar
Gouffon, Y, Bernoulli, D, Dall’agnolo, S,Fantoni, R, Jordan, P, Madritsch, H, Mosar, J, Picotti, V, Pfiffner, OA, Schenker, FL, Schlunegger, F and Schmid, SM (2024b). Tectonic Map of Switzerland 1: 500 000, Explanatory notes. – FederalOffice of Topography swisstopo, Wabern.Google Scholar
Grevel, C, Schreyer, W, Grevel, KD, Schertl, HP and Willner, AP (2009) REE distribution, mobilization and fractionation in the coesite-bearing ‘pyrope quartzite’and related rocks of the Dora-Maira Massif, Western Alps. European Journal of Mineralogy 21(6), 1213–24.CrossRefGoogle Scholar
Groppo, C, Ferrando, S, Gilio, M, Botta, S, Nosenzo, F, Balestro, G, Festa, A & Rolfo, F (2019) What’s in the sandwich? New P–T constraints for the (U) HP nappe stack of southern Dora-Maira Massif (Western Alps). European Journal of Mineralogy 31(4), 665–83.CrossRefGoogle Scholar
Groppo, C, Ferrando, S, Tursi, F and Rolfo, F (2025) A New UHP-HP Tectono-Metamorphic Architecture for the Southern Dora-Maira Massif Nappe Stack (Western Alps) Based on Petrological and Microstructural Evidence. Journal of Metamorphic Geology 43(4), 359–83.CrossRefGoogle Scholar
Groppo, C. T., Castelli, D. C. C. and Rolfo, F. (2007) HT, Pre-Alpine relics in a spinel-bearing dolomite marble from the UHP Brossasco-Isasca Unit (Dora-Maira Massif, western Alps). Periodico di Mineralogia 76, 155–68.Google Scholar
Guillot, F, Liégeois, JP, Fabre, J (1991) Des granophyres du Cambrien terminal dans le Mont Pourri (Vanoise, zone briançonnaise): première datation U–Pb sur zircon d’un socle des zones internes des Alpes françaises. Comptes Rendus de l’Académie des Sciences II 313, 239–44Google Scholar
Guillot, F, Schaltegger, U, Bertrand, JM, Deloule, É and Baudin, T (2002) Zircon U–Pb geochronology of Ordovician magmatism in the polycyclic Ruitor Massif (internal W Alps). International Journal of Earth Sciences 91, 964–78.CrossRefGoogle Scholar
Guillot, S, di Paola, S, Ménot, RP, Ledru, P, Spalla, MI, Gosso, G and Schwartz, S (2009) Suture zones and importance of strike-slip faulting for Variscan geodynamic reconstructions of the External Crystalline Massifs of the western Alps. Bulletin de la Société géologique de France 180(6), 483500.CrossRefGoogle Scholar
Handy, M, Franz, L, Heller, F, Janott, B and Zurbriggen, R (1999) Multistage accretion and exhumation of the continental crust (Ivrea crustal section, Italy and Switzerland). Tectonics 18, 1154–77.CrossRefGoogle Scholar
Handy, M, Schmid, S, Paffrath, M and Friederich, W (2021) European tectosphere and slabs beneath the greater Alpine area–Interpretation of mantle structure in the Alps-Apennines-Pannonian region from teleseismic Vp studies. Solid Earth Discussions 2021, 161.Google Scholar
Handy, ML, Schmid, SM, Bousquet, R, Kissling, E and Bernoulli, D (2010) Reconciling plate-tectonic reconstructions of Alpine Tethys with the geological–geophysical record of spreading and subduction in the Alps. Earth-Science Reviews 102(3-4), 121–58. https://doi.org/10.1016/j.earscirev.2010.06.002 CrossRefGoogle Scholar
Hauschke, N. and Wilde, V. (1999) Trias: eine ganz andere Welt: Mitteleuropa im frühen Erdmittelalter. Verlag Dr. Friedrich Pfeil, 636.Google Scholar
Heinrichs, T, Siegesmund, S, Frei, D, Drobe, M and Schulz, B (2012) Provenance signatures from whole-rock geochemistry and detrital zircon ages of metasediments from the Austroalpine basement south of the Tauern Window (Eastern Tyrol, Austria). Géologie Alpine 9(2012), 156–85.Google Scholar
Henk, A, Franz, L, Teufel, S and Oncken, O (1997) Magmatic underplating, extension, and crustal reequilibration: Insights from a cross-section trough the Ivrea Zone and Strona-Ceneri Zone, Northern Italy. Journal of Geology 105, 367–77.CrossRefGoogle Scholar
Henry, C (1990) L’unité à coesite du massif Dora-Maira dans son cadre pétrologique et structural (Alpes occidentales, Italie). PhD thesis, Université Pierre et Marie Curie, Paris, France. Published thesis. https://hal.archives-ouvertes.fr/tel-03511047 Google Scholar
Henry, C, Michard, A and Chopin, C (1993) Geometry and structural evolution of ultra-high-pressure and high-pressure rocks from the Dora-Maira massif, Western Alps, Italy. Journal of Structural Geology XV(8), 965–81. https://doi.org/10.1016/0191-8141(93)90170-F CrossRefGoogle Scholar
Hermann, J and Rubatto, D (2003). Relating zircon and monazite domains to garnet growth zones: age and duration of granulite facies metamorphism in the Val Malenco lower crust. Journal of Metamorphic Geology 21(9), 833–52.CrossRefGoogle Scholar
Hunziker, JC, Desmons, J and Hurford, AJ (1992) Thirty-two years of geochronological work in the Central and Western Alps: a review on seven maps (p. 13). Université de Lausanne/Faculté des sciences/Institut de minéralogie et de pétrographie.Google Scholar
Kienast, JR, Lombardo, B, Biino, C and Pinardon, JL (1991) Petrology of very-high-pressure eclogitic rocks from the Brossasco-Isasca complex, Dora-Maira massif, Italian Western Alps. Journal of Metamorphic Geology 9(1), 1934.CrossRefGoogle Scholar
Klötzli-Chowanetz, E, Klötzli, U and Koller, F (1997) Lower Ordovician migmatisation in the Ötztal crystalline basement (Eastern Alps, Austria): linking U-Pb and Pb-Pb dating with zircon morphology. Schweizerische Mineralogische und Petrographische Mitteilungen 77(3), 315–24.Google Scholar
Kroner, U and Romer, RL (2013) Two plates—many subduction zones: the Variscan orogeny reconsidered. Gondwana Research 24(1), 298329.CrossRefGoogle Scholar
Kunz, BE, Johnson, TE, White, RW and Redler, C (2014) Partial melting of metabasic rocks in Val Strona di Omegna, Ivrea Zone, northern Italy. Lithos 190, 112.CrossRefGoogle Scholar
Kunz, BE, Manzotti, P, von Niederhäusern, B, Engi, M, Darling, JR, Giuntoli, F and Lanari, P (2018) Permian high-temperature metamorphism in the Western Alps (NW Italy). International Journal of Earth Sciences 107, 203–29.CrossRefGoogle Scholar
Liati, A, Gebauer, D and Fanning, CM (2009) Geochronological evolution of HP metamorphic rocks of the Adula nappe, Central Alps, in pre-Alpine and Alpine subduction cycles. Journal of the Geological Society 166(4), 797810.CrossRefGoogle Scholar
Liati, A, Gervers, M, Froitzheim, N and Fanning, C (2001) U-Pb SHRIMP geochronology of an amphibolitised eclogite and an orthogneiss from the Furgg zone (Western Alps) and implications for its geodynamic evolution. Schweizerische Mineralogische und Petrographische Mitteilungen 81, 379–93.Google Scholar
Linnemann, U, Gerdes, A, Drost, K and Buschmann, B (2007) The continuum between Cadomian orogenesis and opening of the Rheic Ocean: Constraints from LA-ICP-MS U-Pb zircon dating and analysis of plate-tectonic setting (Saxo-Thuringian zone, northeastern Bohemian Massif, Germany). Geological Society of America, Special Papers 423, 6196.Google Scholar
Maino, M Gaggero, L, Langone, A, Seno, S and Fanning, M (2019) Cambro-Silurian magmatisms at the northern Gondwana margin (Penninic basement of the Ligurian Alps). Geoscience Frontiers 10(1), 315–30.CrossRefGoogle Scholar
Mandl, M, Kurz, W, Hauzenberger, C, Fritz, H, Klötzli, U, Schuster, R (2018) Pre-Alpine evolution of the Seckau Complex (Austroalpine basement/Eastern Alps): Constraints from in-situ LA-ICP-MS U-Pb zircon geochronology. Lithos 296–299, 412–30CrossRefGoogle Scholar
Manzotti, P, Ballèvre, M and Dal Piaz, GV (2017) Continental gabbros in the Dent Blanche Tectonic System (Western Alps): From the pre-Alpine crustal structure of the Adriatic palaeo-margin to the geometry of an alleged subduction interface. Journal of the Geological Society 174(3), 541–56.CrossRefGoogle Scholar
Manzotti, P., Ballèvre, M. and Poujol, M. (2016) Detrital zircon geochronology in the Dora-Maira and Zone Houillère: a record of sediment travel paths in the Carboniferous. Terra Nova 28(4), 279–88.CrossRefGoogle Scholar
Manzotti, P, Ballèvre, M, Zucali, M, Robyr, M and Engi, M (2014) The tectonometamorphic evolution of the Sesia–Dent Blanche nappes (internal Western Alps): review and synthesis. Swiss Journal of Geosciences 107(2), 309–36.CrossRefGoogle Scholar
Manzotti, P, Millonig, LJ, Gerdes, A, Whitehouse, MJ, Jeon, H, Poujol, M and Ballèvre, M (2025a) Protolith age, and timing of burial and exhumation of the UHP Chasteiran Unit (Dora-Maira Massif, Western Alps), constrained by zircon, garnet and rutile petrochronology. Lithos 107951.CrossRefGoogle Scholar
Manzotti, P, Poujol, M and Ballèvre, M (2015) Detrital zircon geochronology in blueschist-facies meta-conglomerates from the Western Alps: implications for the late Carboniferous to early Permian palaeogeography. International Journal of Earth Sciences 104, 703–31.CrossRefGoogle Scholar
Manzotti, P, Rubatto, D, Darling, J, Zucali, M, Cenki-Tok, B and Engi, M (2012) From Permo-Triassic lithospheric thinning to Jurassic rifting at the Adriatic margin: Petrological and geochronological record in Valtournenche (Western Italian Alps). Lithos 146, 276–92.CrossRefGoogle Scholar
Manzotti, P, Rubatto, D, Zucali, M, Korh, AE, Cenki-Tok, B, Ballevre, M and Engi, M (2018) Permian magmatism and metamorphism in the Dent Blanche nappe: constraints from field observations and geochronology. Swiss Journal of Geosciences 111, 7997.CrossRefGoogle Scholar
Manzotti, P, Schiavi, F, Ballèvre, M and Nosenzo, F (2025b) Garnet growth across the quartz–coesite transition in metapelites: equilibrium vs. kinetics. European Journal of Mineralogy 37(4), 455–82.CrossRefGoogle Scholar
Manzotti, P, Schiavi, F, Nosenzo, F, Pitra, P and Ballèvre, M (2022) A journey towards the forbidden zone: a new, cold, UHP unit in the Dora-Maira Massif (Western Alps). Contributions to Mineralogy and Petrology 177(6), 122. https://doi.org/10.1007/s00410-022-01923-8 CrossRefGoogle Scholar
Marotta, AM and Spalla, MI (2007) Permian-Triassic high thermal regime in the Alps: Result of late Variscan collapse or continental rifting? Validation by numerical modeling. Tectonics 26(4), TC4016.CrossRefGoogle Scholar
Marquer, D, Challandes, N and Schaltegger, U (1998) Early Permian magmatism in Brianconnais terranes: Truzzo granite and Roffna rhyolite (eastern Penninic nappes, Swiss and Italian Alps). Schweizerische Mineralogische und Petrographische Mitteilungen 78, 397414.Google Scholar
Marthaler, M, Fudral, S, Deville, E and Rampnoux, JP (1986) Mise en évidence du Crétacé supérieur dans la couverture septentrionale de Dora Maira, région de Suse, Italie (Alpes occidentales). Conséquences paléogéographiques et structurales. Comptes rendus de l’Académie des sciences. Série 2, Mécanique, Physique, Chimie, Sciences de l’univers, Sciences de la Terre 302(2), 9196.Google Scholar
Matte, P, Maluski, H, Rajlich, P and Franke, W (1990) Terrane boundaries in the Bohemian Massif: result of large-scale Variscan shearing. Tectonophysics 177(1-3), 151–70.CrossRefGoogle Scholar
Mayer, A, Mezger, K and Sinigoi, S (2000) New Sm–Nd ages for the Ivrea–Verbano Zone, Sesia and Sessera valleys (Northern-Italy). Journal of Geodynamics 30(1-2), 147–66.CrossRefGoogle Scholar
Mégard-Galli, J and Baud, A (1977) Le Trias moyen et supérieur des Alpes nord-occidentales et occidentales: données nouvelles et corrélations stratigraphiques. Bulletin du Bureau de Recherche Géologiques et Minières (BRGM) 2(3), 233–50.Google Scholar
Ménot, RP, Peucat, JJ, Scarenzi, D and Piboule, M (1988) 496 My age of plagiogranites in the Chamrousse ophiolite complex (external crystalline massifs in the French Alps): evidence of a Lower Paleozoic oceanization. Earth and Planetary Science Letters 88(1-2), 8292.CrossRefGoogle Scholar
Mercier, D and Beaudoin, B (1987) Révision du Carbonifère briançonnais: stratigraphie et évolution du bassin. Géologie Alpine 13, 2531.Google Scholar
Merlo, C and Malaroda, SR (1990) Il complesso carbonifero del Pinerolese (Zona Dora-Maira) in val Lemina (Pinerolo). Rendiconti Lincei 1, 313.CrossRefGoogle Scholar
Michard, A. (1967) Etudes géologiques dans les zones internes des Alpes cottiennes. PhD thesis, Centre National de la Recherche Scientifique, Paris, France. Published thesis.Google Scholar
Michard, A, Henry, C and Chopin, C (1993) Compression versus extension in the exhumation of the Dora-Maira coesite-bearing unit, Western Alps, Italy. Tectonophysics 221(2), 173–93. https://doi.org/10.1016/0040-1951(93)90331-D CrossRefGoogle Scholar
Michard, A, Henry, C and Chopin, C (1995) Structures in UHPM rocks: A case study from the Alps. In Ultrahigh Pressure Metamorphism (eds Coleman, R. G., & Wang, X., pp. 206–43. Cambridge University Press.Google Scholar
Michard, A, Schmid, SM, Lahfid, A, Ballèvre, M, Manzotti, P, Chopin, C, Iaccarino, S & Dana, D (2022) The Maira-Sampeyre and Val Grana Allochthons (South Western Alps): review and new data on the tectonometamorphic evolution of the Briançonnais distal margin. Swiss Journal of Geosciences 115 (1), 143. https://doi.org/10.1186/s00015-022-00419-8 CrossRefGoogle Scholar
Michard, A and Vialon, P (1966) Permo-Trias, Permien s. 1. Et Permo-Carbonifère metamorphises des Alpes Cottiennes internes: Les facies “Verrucano” et les series volcano-detritiques du Massif Dora-Maira. In Atti Symposium Verrucano, Pisa, Società Toscana di Scienze Naturali (pp. 116–35).Google Scholar
Middlemost, E. A. (1994) Naming materials in the magma/igneous rock system. Earth-Science Reviews 37(3-4), 215–24.CrossRefGoogle Scholar
Mohn, G, Manatschal, G, Müntener, O, Beltrando, M and Masini, E (2010) Unravelling the interaction between tectonic and sedimentary processes during lithospheric thinning in the Alpine Tethys margins. International Journal of Earth Sciences 99, 75101.CrossRefGoogle Scholar
Neubauer, F (2002) Evolution of late Neoproterozoic to early Paleozoic tectonic elements in Central and Southeast European Alpine mountain belts: review and synthesis. Tectonophysics 352(1-2), 87103.CrossRefGoogle Scholar
Neubauer, F, Chang, R, Dong, Y, Genser, J and Liu, Y (2025) Unravelling the history of mountain belts through U–Pb and Lu–Hf dating of zircon and 40Ar/39Ar dating of detrital white mica: a case study from the Eastern Alps. Isotopes in Environmental and Health Studies 61(1), 114–32.CrossRefGoogle Scholar
Nosenzo, F, Manzotti, P, Krona, M, Ballèvre, M and Poujol, M (2024) Tectonic architecture of the northern Dora-Maira Massif (Western Alps, Italy): field and geochronological data. Swiss Journal of Geosciences 117(1), 6.CrossRefGoogle ScholarPubMed
Nosenzo, F, Manzotti, P, Poujol, M, Ballèvre, M and Langlade, J (2022) A window into an older orogenic cycle: P–T conditions and timing of the pre-Alpine history of the Dora-Maira Massif (Western Alps). Journal of Metamorphic Geology 40(4), 789821.CrossRefGoogle Scholar
Nosenzo, F, Manzotti, P and Robyr, M (2023) H2O budget and metamorphic re-equilibration in polycyclic rocks as recorded by garnet textures and chemistry. Lithos 452, 107230.CrossRefGoogle Scholar
Novarese, V (1895) Sul rilevamento eseguito nel 1894 in valle della Germanasca (Alpi Cozie). Bollettino del Reale Comitato Geologico d’Italia 26, 253–82.Google Scholar
Novarese, V (1896) Rilevamento geologico del 1895 nella Val Pellice (Alpi Cozie). Bollettino del Reale Comitato Geologico d’Italia VII, 231–67.Google Scholar
Novarese, V (1898) I giacimenti di grafite delle Alpi Cozie. Bollettino del Reale Comitato Geologico d’Italia 29, 335.Google Scholar
Pantet, A., Epard, J. L., Masson, H., Baumgartner-Mora, C., Baumgartner, P. O. and Baumgartner, L. (2023) Schistes Lustrés in a hyper-extended continental margin setting and reinterpretation of the limit between the Mont Fort and Tsaté nappes (Middle and Upper Penninics, Western Swiss Alps). Swiss Journal of Geosciences 116(1), 2.CrossRefGoogle Scholar
Paoli, G, Stokke, HH, Rocchi, S, Sirevaag, H, Ksienzyk, AK, Jacobs, J and Košler, J (2017) Basement provenance revealed by U–Pb detrital zircon ages: A tale of African and European heritage in Tuscany, Italy. Lithos 277, 376–87.CrossRefGoogle Scholar
Paquette, JL, Montel, JM and Chopin, C (1999) U-Th-Pb dating of the Brossasco ultrahigh-pressure metagranite, Dora-Maira massif, Western Alps. European Journal of Mineralogy 11, 6978. https://doi.org/10.1127/ejm/11/1/0069 CrossRefGoogle Scholar
Passchier, C. W. and Trouw, R. A. (2005) Microtectonics. Berlin, Heidelberg: Springer Berlin Heidelberg.Google Scholar
Paton, C, Hellstrom, J, Paul, B, Woodhead, J and Hergt, J (2011) Iolite: Freeware for the visualisation and processing of mass spectrometric data. Journal of Analytical Atomic Spectrometry 26(12), 2508–18.CrossRefGoogle Scholar
Paton, C, Woodhead, JD, Hellstrom, JC, Hergt, JM, Greig, A and Maas, R (2010) Improved laser ablation U-Pb zircon geochronology through robust downhole fractionation correction. Geochemistry, Geophysics, Geosystems 11(3), Q0AA06.CrossRefGoogle Scholar
Pawlig, S and Baumgartner, LP (2001) Geochemistry of a talc-kyanite-chloritoid shear zone within the Monte Rosa granite, Val d’Ayas, Italy. Schweizerische mineralogische und petrographische Mitteilungen 81, 329–46.Google Scholar
Pearce, JA, Harris, NB and Tindle, AG (1984) Trace element discrimination diagrams for the tectonic interpretation of granitic rocks. Journal of Petrology 25(4), 956–83.CrossRefGoogle Scholar
Pearce, JA (1996) A User’s Guide to Basalt Discrimination Diagrams. In Trace Element Geochemistry of Volcanic Rocks: Applications for Massive Sulphide Exploration: Geological Association of Canada, Short Course Notes, (ed Wyman, D.A.), Vol. 12, pp. 79–113. St. John’s.Google Scholar
Petri, B, Mohn, G, Skrzypek, E, Mateeva, T, Galster, F and Manatschal, G (2017) U–Pb geochronology of the Sondalo gabbroic complex (Central Alps) and its position within the Permian post-Variscan extension. International Journal of Earth Sciences 106, 2873–93.CrossRefGoogle Scholar
Pezzotta, F and Pinarelli, L (1994) The magmatic evolution of Ordovician metagranitoids of the Serie dei Laghi (Southern Alps): Inferences from petrological, geochemical, and Sr and Nd isotope data. Periodico di Mineralogia. 63, 127–47.Google Scholar
Piana, F, Fioraso, G, Irace, A, Mosca, P, d’Atri, A. Barale, L, Falletti, P, Monegato, G, Morelli, M, Tallone, S & Vigna, GB (2017) Geology of Piemonte region (NW Italy, Alps–Apennines interference zone). Journal of Maps 13(2), 395405. https://doi.org/10.1080/17445647.2017.1316218 CrossRefGoogle Scholar
Pieruccioni, D, Vezzoni, S and Petrelli, M (2018) A petrographic and U-Pb geochronological approach to the reconstruction of the pre-alpine history of Alpi Apuane (Tuscany). Atti Società. Toscana Scienze Naturali, Memorie, Serie A 125, 6980.Google Scholar
Pillola, GL, Leone, F and Loi, A (1998) The Cambrian and early Ordovician of SW Sardinia. Giornale di Geologia 60, 2538 Google Scholar
Poli, G and Tommasini, S (1991) A geochemical approach to the evolution of granitic plutons: a case study, the acid intrusions of Punta Falcone (northern Sardinia, Italy). Chemical geology 92(1-3), 87105.CrossRefGoogle Scholar
Poller, U. (1997) U–Pb single-zircon study of gabbroic and granitic rocks in the Val Barlasch (Silvretta nappe, Switzerland). Schweizerische Mineralogische und Petrographische Mitteilungen 77, 351–60.Google Scholar
Ring, U, Collins, AS and Kassem, OK (2005) U-Pb SHRIMP data on the crystallization age of the Gran Paradiso augengneiss, Italian Western Alps: Further evidence for Permian magmatic activity in the Alps during break-up of Pangea. Eclogae Geologicae Helvetiae 98, 363–70.CrossRefGoogle Scholar
Rubatto, D (2017) Zircon: the metamorphic mineral. Reviews in Mineralogy and Geochemistry 83(1), 261–95.CrossRefGoogle Scholar
Sacchi, R., Balestro, G., Cadoppi, P., Carraro, F., Delle Piane, L., Di Martino, L., Enrietti, M., Gallarà, F., Gattiglio, M., Martinotti, G. & Perello, P. (2004) - Studi geologici in Val di Susa finalizzati ad un nuovo collegamento ferroviario Torino-Lione. Museo Regionale Sci. Nat. Torino, Monografie, 41, 117 pp.Google Scholar
Sandrone, R, Barla, G, Bianco, G, Compagnoni, R, Giani, GP (1982) Planning of quarrying operations in the “Pietra di Luserna” basin (Western Italian Alps). Proceedings 4th International Congress I.A.E.G., New Delhi, 10–15 December 1982, 413–424.Google Scholar
Sandrone, R, Borghi, A, Carosso, G, Morsetti, C, Tagliano, C and Zucchetti, S (1990) Geometry of the talc deposit of Fontane, and structural evolution of the area (Dora Maira Massif, Western Alps). Bollettino Associazione Mineraria Subalpina 27, 4562.Google Scholar
Sandrone, R, Cadoppi, P, Sacchi, R and Vialon, P (1993) The Dora-Maira Massif. In Pre-Mesozoic Geology in the Alps (ed von Raumer J.F., N.F.). pp. 317–25. Berlin: Springer. https://doi.org/10.1007/978-3-642-84640-3_18 CrossRefGoogle Scholar
Sandrone, R, Cordola, M and Fontan, D (1986) Pre-Variscan granitoids in the Dora-Maira Massif (Cottian Alps). Atti della Accademia delle Scienze di Torino 120, 179–89.Google Scholar
Sandrone, R, Sacchi, R, Cordola, M, Fontan, D and Villa, IM (1988) Metadiorites in the Dora-Maira polymetamorphic basement (Cottian Alps). Rendiconti della Società Italiana di Mineralogia e Petrologia 43(2), 593608.Google Scholar
Sartori, M, Gouffon, Y and Marthaler, M (2006) Harmonisation et définition des unités lithostratigraphiques briançonnaises dans les nappes penniques du Valais. Eclogae Geologicae Helvetiae 99, 363407.CrossRefGoogle Scholar
Schaltegger, U (1997) Magma pulses in the Central Variscan Belt: episodic melt generation and emplacement during lithospheric thinning. Terra Nova 9(5-6), 242–45.CrossRefGoogle Scholar
Schaltegger, U and Brack, P (2007) Crustal-scale magmatic systems during intracontinental strike-slip tectonics: U, Pb and Hf isotopic constraints from Permian magmatic rocks of the Southern Alps. International Journal of Earth Sciences 96, 1131–51.CrossRefGoogle Scholar
Schaltegger, U and Gebauer, D (1999) Pre-Alpine geochronology of the central, western and southern Alps. Schweizerische mineralogische und petrographische Mitteilungen 79(1), 7987.Google Scholar
Scheiber, T, Adrian Pfiffner, O and Schreurs, G (2013) Upper crustal deformation in continent-continent collision: a case study from the Bernard nappe complex (Valais, Switzerland). Tectonics 32(5), 1320–42.CrossRefGoogle Scholar
Scheiber, T, Berndt, J, Mezger, K and Pfiffner, OA (2014) Precambrian to Paleozoic zircon record in the Siviez-Mischabel basement (western Swiss Alps). Swiss Journal of Geosciences 107, 4964.CrossRefGoogle Scholar
Schmid, SM, Fügenschuh, B, Kissling, E and Schuster, R (2004) Tectonic map and over-all architecture of the Alpine orogen. Eclogae Geologicae Helvetiae 97(1), 93117. https://doi.org/10.1007/s00015-004-1113-x CrossRefGoogle Scholar
Schmid, SM, Kissling, E, Dichl, T, van Hinsbergen, DJ and Molli, G (2017) Ivrea mantle wedge, arc of the Western Alps, and kinematic. Swiss Journal of Geoscience 110, 581612. https://doi.org/10.1007/s00015-016-0237-0 CrossRefGoogle Scholar
Schuster, R, Scharbert, S, Abart, R and Frank, W (2001) Permo-Triassic extension and related HT/LP metamorphism in the Austroalpine-Southalpine realm. Mitteilungen der Österreichischen Geologischen Gesellschaft 45, 111–41.Google Scholar
Schuster, R and Stüwe, K (2008) Permian metamorphic event in the Alps. Geology 36(8), 603–06.CrossRefGoogle Scholar
Scotese, CR (2021) An atlas of Phanerozoic paleogeographic maps: the seas come in and the seas go out. Annual Review of Earth and Planetary Sciences 49(1), 679728.CrossRefGoogle Scholar
Siegesmund, S, Oriolo, S, Broge, A, Hueck, M, Lammerer, B, Basei, MA and Schulz, B. (2023) Cadomian to Cenerian accretionary orogenic processes in the Alpine basement: the detrital zircon archive. International Journal of Earth Sciences 112(4), 1157–74.CrossRefGoogle Scholar
Siegesmund, S, Oriolo, S, Schulz, B, Heinrichs, T, Basei, MAS and Lammerer, B (2021) The birth of the Alps: Ediacaran to Paleozoic accretionary processes and crustal growth along the northern Gondwana margin. International Journal of Earth Sciences 110(4), 1321–48.CrossRefGoogle Scholar
Simonetti, M, Carosi, R, Montomoli, C, Cottle, JM and Law, RD (2020) Transpressive deformation in the southern european variscan belt: new insights from the Aiguilles Rouges Massif (Western Alps). Tectonics 39(6), e2020TC006153.CrossRefGoogle Scholar
Simonetti, M, Carosi, R, Montomoli, C, Langone, A, D’Addario, E and Mammoliti, E (2018) Kinematic and geochronological constraints on shear deformation in the Ferriere-Mollières shear zone (Argentera-Mercantour Massif, Western Alps): implications for the evolution of the Southern European Variscan Belt. International Journal of Earth Sciences 107(6), 2163–89.CrossRefGoogle Scholar
Simonetti, M, Carosi, R, Montomoli, C, Law, RD and Cottle, JM (2021) Unravelling the development of regional-scale shear zones by a multidisciplinary approach: The case study of the Ferrière-Mollières Shear Zone (Argentera Massif, Western Alps). Journal of Structural Geology 149, 104399.CrossRefGoogle Scholar
Sláma, J, Košler, J, Condon, DJ, Crowley, JL, Gerdes, A, Hanchar, JM, … & Whitehouse, MJ (2008) Plešovice zircon—a new natural reference material for U–Pb and Hf isotopic microanalysis. Chemical Geology 249(1-2), 135.CrossRefGoogle Scholar
Spalla, M. I., Zanoni, D., Marotta, A. M., Rebay, G., Roda, M., Zucali, M. and Gosso, G. (2014) The transition from Variscan collision to continental break-up in the Alps: insights from the comparison between natural data and numerical model predictions. Geological Society, London, Special Publications 405, 363400, https://doi.org/10.1144/SP405.11 CrossRefGoogle Scholar
Spencer, CJ, Roberts, NMW and Santosh, M (2017) Growth, destruction and preservation of Earth’s continental crust. Earth-Science Reviews 172, 87106.CrossRefGoogle Scholar
Stampfli, G. M., Hochard, C., Vérard, C. and Wilhem, C. (2013) The formation of Pangea. Tectonophysics 593, 119.CrossRefGoogle Scholar
Stampfli, GM (1993) Le Briançonnais, terrain exotique dans les Alpes?. Eclogae Geologicae Helvetiae 86, 145.Google Scholar
Stampfli, GM and Borel, GD (2002) A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and restored synthetic oceanic isochrons. Earth and Planetary Science Letters 196(1-2), 1733.CrossRefGoogle Scholar
Stella, A (1895) Sul rilevamento geologico eseguito nel 1894 in Valle Varaita (Alpi Cozie). Bollettino del Regio Comitato Geologico d’Italia 26, 283313.Google Scholar
Stella, A (1896) Sul rilevamento geologico eseguito in Valle Po (Alpi Cozie) nel 1895. Bollettino del Regio Comitato Geologico d’Italia 27, 268–96.Google Scholar
Stipp, M, StuÈnitz, H, Heilbronner, R Schmid, SM (2002a) 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), 1861–84.CrossRefGoogle Scholar
Stipp, M, Stünitz, H, Heilbronner, R and Schmid, SM (2002b) Dynamic recrystallization of quartz: correlation between natural and experimental conditions. Geological Society, London, Special Publications 200, 171–90.CrossRefGoogle Scholar
Sun, SS and McDonough, WF (1989) Chemical and isotopic systematics of oceanic basalts: implications for mantle composition and processes. Geological Society, London, Special Publications 42(1), 313–45.CrossRefGoogle Scholar
Thélin, P, Sartori, M, Burri, M, Gouffon, Y and Chessex, R (1993) The pre-Alpine basement of the briançonnais (Wallis, Switzerland). In Pre-Mesozoic geology in the Alps (pp. 297315). Springer, Berlin, Heidelberg.CrossRefGoogle Scholar
Thélin, P, Sartori, M, Lengeler, R and Schaerer, J-P (1990) Eclogites of Paleozoic or early Alpine age in the basement of the Penninic Siviez-Mischabel nappe, Wallis, Switzerland. Lithos 25, 7188. https://doi.org/10.1016/0024-4937(90)90007-N CrossRefGoogle Scholar
Thiéblemont, D, Jacob, JB, Lach, P, Guerrot, C and Leguérinel, M (2023) First report of an Ediacarian basement in the Western Alps: the Serre Chevalier crystalline unit (Briançonnais domain, France). BSGF-Earth Sciences Bulletin 194, 16.CrossRefGoogle Scholar
Thöny, WF, Tropper, P, Schennach, F, Krenn, E, Finger, F, Kaindl, R, … & Hoinkes, G. (2008) The metamorphic evolution of migmatites from the Ötztal Complex (Tyrol, Austria) and constraints on the timing of the pre-Variscan high-T event in the Eastern Alps. Swiss Journal of Geosciences 101, 111–26.CrossRefGoogle Scholar
Vavra, G, Gebauer, D, Schmidt, R and Compston, W (1996) Multiple zircon growth and recrystallization during polyphase Late Carboniferous to Triassic metamorphism in granulites of the Ivrea Zone (southern Alps): An ion microprobe (SHRIMP) study. Contributions to Mineralogy and Petrology 122, 337–58CrossRefGoogle Scholar
Vavra, G, Schmid, R. and Gebauer, D. (1999) Internal morphology, habit and U-Th-Pb microanalysis of amphibolite-to-granulite facies zircons: Geochronology of the Ivrea Zone (Southern Alps). Contributions to Mineralogy and Petrology 134, 380404.CrossRefGoogle Scholar
Vermeesch, P (2018) IsoplotR: A free and open toolbox for geochronology. Geoscience Frontiers 9(5), 1479–93.CrossRefGoogle Scholar
Vialon, P (1966) Etude géologique du massif cristallin Dora-Maira: Alpes cottiennes internes: Italie. PhD thesis, Université de Grenoble, Grenoble, France. Published thesis.Google Scholar
Villaseca, C, Martínez, EM, Orejana, D, Andersen, T and Belousova, E (2016) Zircon Hf signatures from granitic orthogneisses of the Spanish Central System: Significance and sources of the Cambro-Ordovician magmatism in the Iberian Variscan Belt. Gondwana Research 34, 6083.CrossRefGoogle Scholar
Vola, G, Ardit, M, Frijia, G, Cavallo, A, Natali, C, Mion, CB, Lugli, F & Primavori, P (2022) Archaeometric characterization of historical-contemporaneous marbles from Waldensian Valleys of Piedmont (Dora-Maira Massif, Cottian Alps, Italy). Journal of Archaeological Science: Reports 45, 103562.Google Scholar
von Gümbel, CW (1891) Geognostiche Beschreibung des bayerischen Alpengebirges und seines Vorlandes, Perthes, Gotha, 970 pp.Google Scholar
Von Raumer, J, Stampfli, G, Borel, G and Bussy, F (2002) Organization of pre-Variscan basement areas at the north-Gondwanan margin. International Journal of Earth Sciences 91, 3552.CrossRefGoogle Scholar
Von Raumer, JF (1998) The Palaeozoic evolution in the Alps: from Gondwana to Pangea. Geologische Rundschau 87, 407–35.CrossRefGoogle Scholar
Von Raumer, JF, Bussy, F, Schaltegger, U, Schulz, B and Stampfli, GM (2013) Pre-Mesozoic Alpine basements—their place in the European Paleozoic framework. GSA Bulletin 125(1-2), 89108.CrossRefGoogle Scholar
Von Raumer, JF and Neubauer, F (1993) Late Precambrian and Palaeozoic evolution of the Alpine basement—an overview. In Pre-Mesozoic Geology in the Alps. (eds von Raumer, J.F., Neubauer, F.), pp. 625–639. Berlin, HeidelbergSpringer. https://doi.org/10.1007/978-3-642-84640-3_37.CrossRefGoogle Scholar
Von Raumer, JF and Stampfli, GM (2008) The birth of the Rheic Ocean—Early Palaeozoic subsidence patterns and subsequent tectonic plate scenarios. Tectonophysics 461(1-4), 920.CrossRefGoogle Scholar
Von Raumer, JF, Stampfli, GM, Arenas, R and Sánchez Martínez, S (2015) Ediacaran to Cambrian oceanic rocks of the Gondwana margin and their tectonic interpretation. International Journal of Earth Sciences 104, 1107–21.CrossRefGoogle Scholar
Voshage, H, Hunziker, JC, Hofmann, AW and Zingg, A (1987) A Nd and Sr isotopic study of the Ivrea zone, Southern Alps, N-Italy. Contributions to Mineralogy and Petrology 97, 3142.CrossRefGoogle Scholar
Wheeler, J (1991) Structural evolution of a subducted continental sliver: the northern Dora Maira massif, Italian Alps. Journal of the Geological Society 148(6), 1101–13.CrossRefGoogle Scholar
Whitney, DL and Evans, BW (2010) Abbreviations for names of rock-forming minerals. American Mineralogist 95(1), 185–87.CrossRefGoogle Scholar
Wiedenbeck, MAPC, Alle, P, Corfu, FY, Griffin, WL, Meier, M, Oberli, FV, … & Spiegel, W (1995) Three natural zircon standards for U-Th-Pb, Lu-Hf, trace element and REE analyses. Geostandards newsletter 19(1), 123.CrossRefGoogle Scholar
Žák, J., Verner, K., Janoušek, V., Holub, F. V., Kachlík, V., Finger, F., … & Trubač, J. (2014) A plate-kinematic model for the assembly of the Bohemian Massif constrained by structural relationships around granitoid plutons. Geological Society, London, Special Publications 405, 169–96.CrossRefGoogle Scholar
Zanettin-Lorenzoni, E (1967) Su un fenomeno di granodioritizzazione degli gneiss biotitico-epidotici di Albarea (Val Chisone – Alpi Cozie). Atti e Memorie dell’Accademia Patavina di Scienze, Lettere ed Arti, Classe Scienze Matematiche Naturali 79, 129–69.Google Scholar
Ziegler, PA (1993) Late Paleozoic-Early Mesozoic plate reorganization: Evolution and demise of the Variscan fold belt. In Pre-Mesozoic geology in the Alps (eds Raumer, JF & Neubauer, F), Vol. 2, pp. 203–16. Berlin: Springer.CrossRefGoogle Scholar
Zurbriggen, R (2017) The Cenerian orogeny (early Paleozoic) from the perspective of the Alpine region. International Journal of Earth Sciences 106(2), 517–29.CrossRefGoogle Scholar
Zurbriggen, R (2020) Banded amphibolites in the Alps: a new interpretation in relation to early Paleozoic peraluminous magmatism. Swiss Journal of Geosciences 113(1), 10.CrossRefGoogle Scholar
Figure 0

Figure 1. Tectonic sketch of the Western Alps (modified from Dana, 2024). The geolocation of the map is shown in the upper left insert. Magmatic protoliths with ages around 500–525 Ma are reported (1: Liati et al., 2001; 2: Sartori et al., 2006; 3: Scheiber et al., 2014; 4: Bussy et al., 1996; 5: Guillot et al., 1991; 6: Bertrand et al., 2000b, 7: Bertrand et al.,2000a; 8: Ménot et al., 1988; 9: Thiéblemont et al., 2023; 10: Balestro et al., 2022; 11: Filippi et al., 2024). The tectonic subdivision of the Western Alps is largely based on Schmid et al. (2004, 2017) and Gouffon et al. (2024b). The red box highlights the position of Fig. 2. Abbreviations of the tectonic units cited in the text: AM = Ambin Massif, DM = Dora-Maira, GP = Gran Paradiso, MR = Monte Rosa, R = Ruitor, SM = Siviez-Mischabel, ZH = Zone Houillère.

Figure 1

Figure 2. Tectonic sketch of the Dora-Maira Massif (compiled from Henry, 1990; Sandrone et al., 1993; Piana et al., 2017; Michard et al., 2022; Nosenzo et al., 2024; Groppo et al., 2025 and our new observations). Magmatic photolith ages from the literature are given (a: Bussy & Cadoppi, 1996; b: Nosenzo et al., 2024; c: Nosenzo et al., 2022; d: Paquette et al., 1999; e: Chen et al., 2017; f: Balestro et al., 2022; g: Gebauer et al., 1997). Tectonic units abbreviations: VG = Val Grana, VMS = Val Maira-Sampeyre, CLSZ = Cima Lubin Shear Zone, DU = Dronero, VSZ = Valmala Shear Zone, RU = Ricordone, SE = Serre, MU = Muret, RS = Rocca Solei, GU = Grimbassa, CU = Chasteiran, SC = San Chiaffredo, BIU = Brossasco-Isasca, PSU = Pinerolo–Sanfront, MV = Monviso, OR = Orsiera-Rocciavrè, SU = Susa, LA = Lanzo. The yellow stars show the location of the studied samples; the black square is the area covered by the Geological Map of Italy sheet 172 ‘Pinerolo’.

Figure 2

Table 1. Meta-intrusive rocks of the Dora-Maira Massif and associated regional shear zones (excluding meta-mafic rocks), with their main features (MR = magmatic relics, MM = metamorphic minerals, ACC = accessory minerals) and the proposed or published absolute ages available in the literature prior to this study. Reference: (1) Franchi (1898), Franchi & Novarese (1895), Novarese (1895; 1896; 1898) and Stella (1895; 1896); (2) Vialon (1966); (3) Borghi et al. (1984); (4) Sandrone et al. (1986); (5) Sandrone et al. (1988); (6) Henry (1990); (7) Cadoppi (1990); (8) Wheeler (1991); (9) Bussy & Cadoppi (1996); (10) Gebauer et al. (1997); (11) Paquette et al (1999); (12) Chen et al. (2017); (13) Balestro et al. (2022); (14) Nosenzo et al. (2022); (15) Nosenzo et al. (2024)

Figure 3

Table 2. Meta-intrusive samples studied in this contribution. Coordinates are given in WGS84 system. Mineral abbreviations after Whitney & Evans (2010), except for Wm = white mica

Figure 4

Figure 3. Selected field images and photomicrographs of the studied lithologies. (a) Outcrop of Malanaggio metadiorite with the typical banded appearance and microgranular mafic enclaves. (b) Crossed polarized light (CPL hereafter) microphoto of the Malanaggio metadiorite. (c) Fine-grained facies of the Freidour orthogneiss. (d) Dyke of Freidour orthogneiss cross-cutting the Malanaggio metadiorite. (e) CPL photomicrograph of a large K-feldspar porphyroclast wrapped by the phengite foliation in the Freidour orthogneiss. (f) Field image of the Ferrera orthogneiss felsic facies, W of Bobbio Pellice village. (g) Field image of the Ferrera orthogneiss leucocratic facies (100 m from the outcrop shown in f). (h) CPL photomicrograph of the Ferrera orthogneiss (sample SD2) showing the spaced foliation defined by white mica.

Figure 5

Figure 4. Selected field images and photomicrographs of the studied lithologies. (a) Ghiandone orthogneiss showing the distinctive large polycrystalline aggregates porphyroclasts and microgranular mafic enclaves. (b) CPL photomicrograph of the Ghiandone orthogneiss. (c) Field image of the Granero orthogneiss with cm-sized K-feldspar porphyroclast. (d) CPL photomicrograph of Granero orthogneiss (sample DD43) showing K-feldspar porphyroclast enveloped by the main foliation defined by white mica and quartz-felspathic levels. (e) Typical banding of the Cialancia orthogneiss, with alternating dark and leucocratic levels. (f) Aplitic dyke associated to the Cialancia orthogneiss intruding the host micaschist. (g) CPL photomicrograph of Cialancia orthogneiss (sample Y6) showing a K-feldspar porphyroclast surrounded by a matrix of plagioclase, quartz and phengite. (h) Giasset metadiorite dyke cross-cutting the Ghiandone orthogneiss, suggesting a younger protolith age for the metadiorites.

Figure 6

Figure 5. Selected field images and photomicrographs of the studied lithologies. (a) Giasset metadiorite containing cm-sized mafic microgranular enclaves. (b) CPL photomicrograph of the Giasset metadiorite showing fine-grained Alpine mineral crystallization (sample SC58). (c) Field image of the Luserna orthogneiss, with associated quartz-phengitic micaschist. (d) CPL photomicrograph of the Luserna orthogneiss (sample DD109) showing a K-feldspar porphyroclast, wrapped by the main foliation. (e) Field view of the Giulian orthogneiss body sampled within the Cima Lubin Shear Zone, view from the Pellice side of the Germanasca-Pellice watershed ridge. (f) CPL photomicrograph of the Giulian orthogneiss sample (AB21) showing quartz-rich microlithons and a continuous to spaced foliation defined by white mica.

Figure 7

Figure 6. Results of whole-rock bulk-rock geochemistry. (a) TAS diagram for intrusive rocks (Middlemost, 1994). (b) R1–R2 classification diagram of the De la Roche et al. (1980). R1 (millications) = 4Si–11(Na + K)–2(Fe + Ti) and R2 (millications) = 6Ca + 2 Mg + Al. (c) Aluminium Saturation Index (Al/Ca-1.67P+N+K) versus silica plot of Frost and Frost (2008). (d) Zr/TiO2 -Nb/Y classification of granitic rocks after Pearce (1996), note that fewer data points are shown due to the absence of complete trace element analyses in several literature sources. Literature data are from: Vialon (1966), Sandrone et al. (1982); Borghi (1983); Sandrone et al. (1986); Sandrone et al. (1988), Cadoppi (1990); Bussy & Cadoppi (1996); Chen et al. (2016); Nosenzo et al. (2022) and Nosenzo et al. (2024).

Figure 8

Figure 7. Summary of REE patterns of the studied samples. REE patterns are normalized to Chondrite values (Sun & McDonough, 1989) and compared to data from previous studies (grey areas). Where Pr, Tb and Tm were not analysed, dashed lines are used to join Ce and Nd, Gd and Dy and Er and Yb, respectively.

Figure 9

Figure 8. U–Pb zircon geochronology of orthogneiss samples: (a) Freidour orthogneiss; (b) Ferrera orthogneiss; (c) and d) Cialancia orthogneiss. Cathodoluminescence images of zircon crystals are shown on the left side, with the LA-ICP-MS spot position (red circles) and the relative date indicated. The Concordia diagrams on the right side report Concordia ages. Empty dashed ellipses represent dates excluded from the calculation.

Figure 10

Figure 9. U–Pb zircon geochronology of orthogneiss samples: (a) Cialancia orthogneiss; (b) Giasset metadiorite; (c) Luserna orthogneiss; (d) Giulian orthogneiss (tectonic slice in Cima Lubin shear zone). Cathodoluminescence images of zircon crystals are shown on the left side, with the LA-ICP-MS spot position (red circles) and the relative date indicated. The Concordia diagrams on the right side report Concordia ages. Empty dashed ellipses represent dates excluded from the calculation.

Figure 11

Figure 10. Summary of the pre-Alpine evolution of the Dora-Maira Massif comparing the ‘Basement Complex’ units with the Pinerolo–Sanfront unit, modified from Nosenzo et al. (2022). Discontinuous lines indicate hypothetical time periods that are not yet constrained by geochronology. Data are from: 1 – Chen et al. (2024); 2 – Bussy & Cadoppi (1996); 3 – Manzotti et al. (2025); 4 – Nosenzo et al. (2022); 5 – Paquette et al. (1999); 6 – Chen et al. (2017); 7 - Nosenzo et al. (2024); 8 – Cadoppi & Tallone (1992), Sacchi et al. (2004); 9 – Merlo & Malaroda (1990); 10 – Manzotti et al. (2016); 11 – Balestro et al. (2022); 12 – Bonnet et al. (2022); 13 – This work.

Supplementary material: File

Dana et al. supplementary material 1

Dana et al. supplementary material
Download Dana et al. supplementary material 1(File)
File 56.3 KB
Supplementary material: File

Dana et al. supplementary material 2

Dana et al. supplementary material
Download Dana et al. supplementary material 2(File)
File 407.1 KB
Supplementary material: File

Dana et al. supplementary material 3

Dana et al. supplementary material
Download Dana et al. supplementary material 3(File)
File 55.5 KB