Hostname: page-component-7c8c6479df-27gpq Total loading time: 0 Render date: 2024-03-28T17:53:11.954Z Has data issue: false hasContentIssue false

Tracing wedge-internal deformation by means of strontium isotope systematics of vein carbonates

Published online by Cambridge University Press:  23 February 2022

Armin Dielforder*
Affiliation:
Institut für Geologie, Leibniz Universität Hannover, Germany Institut für Geologie, Universität Bern, Switzerland
Igor M. Villa
Affiliation:
Institut für Geologie, Universität Bern, Switzerland
Alfons Berger
Affiliation:
Institut für Geologie, Universität Bern, Switzerland
Marco Herwegh
Affiliation:
Institut für Geologie, Universität Bern, Switzerland
*
Author for correspondence: Armin Dielforder, Email: dielforder@geowi.uni-hannover.de
Rights & Permissions [Opens in a new window]

Abstract

Radiogenic strontium isotopes (87Sr/86Sr) of vein carbonates play a central role in the tectonometamorphic study of fold-and-thrust belts and accretionary wedges and have been used to document fluid sources and fluxes, for example, along major fault zones. In addition, the 87Sr/86Sr ratios of vein carbonates can trace the diagenetic to metamorphic evolution of pore fluids in accreted sediments. Here we present 87Sr/86Sr ratios of vein carbonates from the Infrahelvetic flysch units of the central European Alps (Glarus Alps, Switzerland), which were accreted to the North Alpine fold-and-thrust belt during the early stages of continental collision. We show that the vein carbonates trace the Sr isotopic evolution of pore fluids from an initial seawater-like signature towards the Sr isotopic composition of the host rock with increasing metamorphic grade. This relationship reflects the progressive equilibration of the pore fluid with the host rock and allows us to constrain the diagenetic to low-grade metamorphic conditions of deformation events, including bedding-parallel shearing, imbricate thrusting, folding, cleavage development, tectonic mélange formation and extension. The strontium isotope systematics of vein carbonates provides new insights into the prograde to early retrograde tectonic evolution of the Alpine fold-and-thrust belt and helps to understand the relative timing of deformation events.

Type
FLUID FLOW AND MINERALIZATION IN FAULTS AND FRACTURES
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 (http://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), 2022. Published by Cambridge University Press

1. Introduction

Deformation at convergent plate margins depends on the presence of fluids and their strength-reducing effect. The build-up of pore fluid overpressures reduces the effective normal stresses, which brings the rock closer to failure and allows deformation at low differential stresses (e.g. Hubbert & Rubey, Reference Hubbert and Rubey1959). Such stress and strength conditions facilitate the formation of mineral veins and are common in fold-and-thrust belts and accretionary wedges (e.g. Fisher & Byrne, Reference Fisher and Byrne1987; Dahlen, Reference Dahlen1990; Labaume et al. Reference Labaume, Berty and Laurent1991; Sample & Kopf, Reference Sample, Kopf, Carson, Westbrook, Musgrave and Suess1995; Lacroix et al. Reference Lacroix, Buatier, Labaume, Travé, Dubois, Charpentier, Ventalon and Convert-Gaubier2011; Mittempergher et al. Reference Mittempergher, Cerchiari, Remitti and Festa2017; Ujiie et al. Reference Ujiie, Saishu, Fagereng, Nishiyama, Otsubo, Masuyama and Kagi2018).

Mineral veins are central to studying the tectonometamorphic evolution of convergent margins. From a mechanical perspective, mineral veins provide information on fracture modes, which can be used to constrain fault kinematics and the state of stress during fracturing (e.g. Sibson, Reference Sibson1998; Bons et al. Reference Bons, Elburg and Gomez-Rivas2012). Recently, this aspect has received renewed attention as mineral vein formation in accretionary settings can be linked to stress changes over the earthquake cycle, which allows the constraint of the effective confining pressures (e.g. Sibson, Reference Sibson2013; Takeshita et al. Reference Takeshita, Yamaguchi and Shigematsu2014; Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015; Cerchiari et al. Reference Cerchiari, Remitti, Mittempergher, Festa, Lugli and Cipriani2020). From a geochemical perspective, mineral veins can provide information, for example, on fluid sources, fluid–rock interaction, vein formation temperatures and absolute ages of veins (e.g. Vroliijk et al. Reference Vroliijk, Myers and Moore1988; Sharp & Kirschner, Reference Sharp and Kirschner1995; Tarantola et al. Reference Tarantola, Mullis, Vennemann, Dubessy and de Capitani2007; Sample et al. Reference Sample, Torres, Fisher, Hong, Destringeville, Defliese and Tripati2017; Beaudoin et al. Reference Beaudoin, Lacombe, Roberts and Koehn2018).

Over the past decades, vein formation has been successfully studied by stable isotope (δ13C, δ18O) and radiogenic strontium isotope (87Sr/86Sr) analysis of vein carbonates (e.g. Dietrich et al. Reference Dietrich, McKenzie and Song1983; Burkhard et al. Reference Burkhard, Kerrich, Maas and Fyfe1992; McCaig et al. Reference McCaig, Wayne, Marshall, Banks and Henderson1995; Travé et al. Reference Travé, Labaume, Calvet and Soler1997; Hilgers & Sindern, Reference Hilgers and Sindern2005; Sample, Reference Sample2010; Lacroix et al. Reference Lacroix, Buatier, Labaume, Travé, Dubois, Charpentier, Ventalon and Convert-Gaubier2011). Stable isotope analysis makes use of the temperature-dependent isotopic fractionation between the fluid and solid (e.g. water–calcite) and can provide information on processes such as the rock-buffering of fluids, carbonate diagenesis or the recrystallization of carbonates during metamorphism (e.g. Burkhard et al. Reference Burkhard, Kerrich, Maas and Fyfe1992; Voigt et al. Reference Voigt, Hathorne, Frank, Vollstaedt and Eisenhauer2015; Sample et al. Reference Sample, Torres, Fisher, Hong, Destringeville, Defliese and Tripati2017). By comparison, the temperature-dependent isotopic fractionation of Sr is very small, and this small source of bias is eliminated by normalizing 87Sr/86Sr ratios to a constant 86Sr/88Sr ratio of 0.1194 (Nier, Reference Nier1938; Krabbenhöft et al. Reference Krabbenhöft, Fietzke, Eisenhauer, Liebetrau, Böhm and Vollstaedt2009). The 87Sr/86Sr ratios of vein carbonates and host-rock cements can therefore be used as a tracer. In this respect, the Sr isotope systematics of carbonates has been used to constrain fluid sources, for example, along large thrust faults in the Pyrenees and European Alps and in the foreland of the Rocky Mountains (e.g. Burkhard et al. Reference Burkhard, Kerrich, Maas and Fyfe1992; McCaig et al. Reference McCaig, Wayne, Marshall, Banks and Henderson1995; Machel et al. Reference Machel, Cavell and Patey1996). These studies documented the influx of external fluids based on 87Sr/86Sr ratios that are clearly distinct from the Sr ratio of the host rock and require an external source.

The 87Sr/86Sr ratios of vein carbonates also trace the Sr isotopic evolution of intraformational pore fluids. Fold-and-thrust belts and accretionary wedges typically incorporate calcareous and siliciclastic marine sediments that comprise a mixture of Sr reservoirs. The Sr reservoirs interact with the pore fluid during diagenesis and metamorphism, which might be reflected in the 87Sr/86Sr ratios of different generations of veins carbonates. For example, Travé et al. (Reference Travé, Labaume, Calvet and Soler1997) analysed two groups of mineral veins from a small thrust fault developed within Eocene marlstones of the South Pyrenean foreland basin. While the first group of mineral veins showed 87Sr/86Sr ratios similar to Eocene seawater, the second group showed more radiogenic (that is, higher) 87Sr/86Sr ratios closer to the signature of the host rock. Travé et al. (Reference Travé, Labaume, Calvet and Soler1997) suggested that the difference in the Sr ratios might indicate the evolution of the pore fluid from a seawater-derived fluid towards a diagenetic fluid that interacted with the host rock. However, the impact of external metamorphic fluids could not be excluded. Dielforder et al. (Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015) analysed three groups of mineral veins that formed successively within marine marlstones of the Alpine fold-and-thrust belt. Similar to the veins from the Pyrenees, the 87Sr/86Sr ratios of the three vein groups documented an increase in the Sr ratios from a seawater-like signature towards the isotopic composition of the host rock. Dielforder et al. (Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015) interpreted this trend to reflect the diagenetic to low-grade metamorphic Sr isotopic evolution of the pore fluid that was essentially undisturbed by external fluids.

The finding that vein carbonates may record a systematic Sr isotopic evolution of pore fluids is promising, because it could help to relate deformation events within the larger-scale diagenetic to low-grade metamorphic evolution of fold-and-thrust belts and accretionary wedges. However, the longer-term and larger-scale isotopic evolution of pore fluids in accretionary settings remains insufficiently understood, also because many studies have focused on specific outcrops and/or distinct tectonic structures. Moreover, the Sr isotopic contrast and interaction between initial seawater-derived fluids, carbonates and the siliciclastic components of accreted sediments has not been systematically explored. The absolute differences in 87Sr/86Sr ratios of vein carbonates discussed, for example, by Travé et al. (Reference Travé, Labaume, Calvet and Soler1997) and Dielforder et al. (Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015) are more subtle than those discussed with respect to the influx of external fluids (e.g. McCaig et al. Reference McCaig, Wayne, Marshall, Banks and Henderson1995; Machel et al. Reference Machel, Cavell and Patey1996), and may have been overlooked in the past as scatter in a series of values close to host-rock values.

In this study, we address the Sr isotopic systematics of vein carbonates from marine foreland basin sediments of the North Alpine fold-and-thrust belt extending the dataset of Dielforder et al. (Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015). Our analysis involves several kinds of mineral veins that formed at different stages and in different structural contexts of the prograde to early retrograde tectonometamorphic evolution of the fold-and-thrust belt. In the following, we first provide a geological overview and introduce the different sampling sites located along a 30 km long transect across the fold-and-thrust belt (Section 2 below). Subsequently, we discuss the 87Sr/86Sr ratios in terms of pore fluid evolution and in terms of the structural and tectonic evolution of the marine foreland basin sediments (Section 5 below).

2. Geological background

2.a. Geology of the study area

The European Alps are a collisional orogen resulting from the subduction of the Alpine Tethys and the subsequent collision of the Adriatic microplate (upper plate) with southwestern Eurasia (‘Europe’, lower plate) (e.g. Frisch, Reference Frisch1979; Schmid et al. Reference Schmid, Pfiffner, Froitzheim, Schönborn and Kissling1996). Continental collision commenced in middle Eocene time with the partial subduction of the European continental margin and associated development of the North Alpine fold-and-thrust belt (e.g. Pfiffner, Reference Pfiffner, Allen and Homewood1986; Schmid et al. Reference Schmid, Pfiffner, Froitzheim, Schönborn and Kissling1996; Ford & Lickorish, Reference Ford, Lickorish, Joseph and Lomas2004; Kempf & Pfiffner, Reference Kempf and Pfiffner2004; Handy et al. Reference Handy, Schmid, Bousquet, Kissling and Bernoulli2010).

The study area is located in the central European Alps (Glarus Alps, Switzerland) and provides detailed insights into the evolution of the fold-and-thrust belt, which involved: (1) the Penninic–Austroalpine wedge that formed during the subduction of the Alpine Tethys, (2) the Infrahelvetic flysch units and Subalpine Molasse, (3) the Upper Helvetic nappes, and (4) the European crystalline and sedimentary basement (Fig. 1). The Infrahelvetic flysch units (IFUs) comprise three thrust slices of Upper Cretaceous (Cenomanian–Campanian) to Oligocene passive margin and marine foreland basin sediments, that is, the Ultrahelvetic, South-Helvetic and North-Helvetic flysch units (Milnes & Pfiffner, Reference Milnes and Pfiffner1977; Sinclair & Allen, Reference Sinclair and Allen1992; Jeanbourquin, Reference Jeanbourquin1994; Lihou, Reference Lihou1995, Reference Lihou1996 b) (Fig. 1). These flysch units were scraped off from the subducting European continental margin and accreted in-sequence to the Penninic–Austroalpine wedge (Pfiffner, Reference Pfiffner, Allen and Homewood1986; Dielforder et al. Reference Dielforder, Berger and Herwegh2016). Remnants of the Upper Penninic sediments are preserved in the northern part of the study area (Fig. 1). The subalpine Molasse comprises shallow marine to continental sediments that were deposited in the foreland basin from middle Oligocene time onwards, when the basin became filled (Sinclair & Allen, Reference Sinclair and Allen1992). The Molasse sediments were incorporated into the fold-and-thrust belt after the accretion of the IFUs (Pfiffner, Reference Pfiffner, Allen and Homewood1986; von Hagke et al. Reference von Hagke, Cederbom, Oncken, Stöckli, Rahn and Schlunegger2012; Mock et al. Reference Mock, von Hagke, Schlunegger, Dunkl and Herwegh2020).

Fig. 1. (a) Geological map of the study area and (b) geographical overview. The line A–A′ indicates the trace of the cross-section shown in (c). (c) Synthetic and simplified cross-section. The approximate sampling sites are indicated together with peak metamorphic temperatures. Temperatures based on Ebert et al. (Reference Ebert, Herwegh and Pfiffner2007), Lahfid et al. (Reference Lahfid, Beyssac, Deville, Negro, Chopin and Goffé2010) and Rahn et al. (Reference Rahn, Mullis and Erdelbrock1995). Geological map in (a) and cross-section in (c) based on Pfiffner (Reference Pfiffner2011).

The Upper Helvetic nappes represent the Palaeozoic to Mesozoic sedimentary cover of the European continental margin and were first underthrust and then underplated by duplex accretion to the base of the Alpine wedge. Subsequently, the nappes were thrust along large out-of-sequence thrusts on top of the accreted flysch units and Molasse sediments (see Glarus thrust in Fig. 1c). Thrusting of the Upper Helvetic nappes terminated around late Oligocene to early Miocene time and caused peak metamorphism in the fold-and-thrust belt, which reached zeolite-facies conditions in the north (c. 160–180 °C) and sub-greenschist-facies conditions in the south (c. 300–350 °C) (Hunziker et al. Reference Hunziker, Frey, Clauer, Dallmeyer, Friedrichsen, Flehmig, Hochstrasser, Roggwiler and Schwander1986; Pfiffner, Reference Pfiffner, Allen and Homewood1986; Frey, Reference Frey1988; Rahn et al. Reference Rahn, Mullis and Erdelbrock1995; Ebert et al. Reference Ebert, Herwegh and Pfiffner2007; Lahfid et al. Reference Lahfid, Beyssac, Deville, Negro, Chopin and Goffé2010; Akker et al. Reference Akker, Berger, Zwingmann, Todd, Schrank, Jones, Kewish, Schmid and Herwegh2021 b). Subsequently, deformation relocated into the crystalline basement, which initiated the exhumation of the Aar massif and the retrograde evolution of the overlying fold-and-thrust belt (Pfiffner, Reference Pfiffner, Allen and Homewood1986; Burkhard, Reference Burkhard1990; Glotzbach et al. Reference Glotzbach, Reinecker, Danišík, Rahn, Frisch and Spiegel2010; Herwegh, et al. Reference Herwegh, Berger, Baumberger, Wehrens and Kissling2017, Reference Herwegh, Berger, Glotzbach, Wangenheim, Mock, Wehrens, Baumberger, Egli and Kissling2020; Nibourel et al. Reference Nibourel, Berger, Egli, Luensdorf and Herwegh2018, Reference Nibourel, Berger, Egli, Heuberger and Herwegh2021). Note that the present-day arcuate geometry of the Glarus thrust evolved during the retrograde evolution of the belt after 20 Ma (Rahn et al. Reference Rahn, Hurford and Frey1997) (Fig. 1c).

2.b. Structural evolution of the flysch units and sampling sites

The IFUs constitute a central element of the North Alpine fold-and-thrust-belt and capture the entire collisional evolution from sedimentation in the foreland basin, to accretion, burial, peak metamorphism and exhumation (e.g. Pfiffner, Reference Pfiffner, Allen and Homewood1986; Lihou, Reference Lihou1996 b; Sinclair, Reference Sinclair1997; Herwegh et al. Reference Herwegh, Hürzeler, Pfiffner, Schmid, Abart and Ebert2008; Dielforder et al. Reference Dielforder, Berger and Herwegh2016). The structural evolution of the IFUs is classically distinguished in three deformation phases that approximately encompass the development during the accretion of the sediments, the emplacement of the Upper Helvetic nappes and the retrograde evolution during exhumation, although it should be noted that the deformation phases are diachronous (Schmid, Reference Schmid1975; Milnes & Pfiffner, Reference Milnes and Pfiffner1977; Gasser & den Brok, Reference Gasser and den Brok2008; Dielforder et al. Reference Dielforder, Berger and Herwegh2016). The two main structural elements that occur on a regional scale are a moderately to steeply SE-dipping pressure-solution cleavage and NW-vergent folds. The pressure-solution cleavage is most intensively developed in the central to southern part of the study area, where peak metamorphic temperatures exceeded c. 230 °C. This dependency on peak metamorphic temperatures suggests that the pressure-solution cleavage formed mainly during the higher-grade evolution of the IFUs (Dielforder et al. Reference Dielforder, Berger and Herwegh2016; Akker et al. Reference Akker, Berger, Schrank, Jones, Kewish, Klaver and Herwegh2021 a). In contrast, folding in the IFUs initiated already during the accretion of the flysch units by particulate flow, when the sediments were only weakly lithified and still deformed as soft sediments, but continued throughout the higher-grade evolution of the IFUs (Gasser & den Brok, Reference Gasser and den Brok2008; Dielforder et al. Reference Dielforder, Berger and Herwegh2016).

The structural evolution of the IFUs was accompanied by the widespread formation of mineral veins. This includes mineral veins that are structurally linked to distinct tectonic elements such as faults and folds, as well as mineral veins that occur independently from such elements. In the following, we describe the structural relationships of the mineral veins for the different sampling sites.

2.b.1. Deformation in the Ultrahelvetic flysch unit

The Ultrahelvetic flysch unit was deposited on the most distal part of the European continental margin and was the first flysch unit that was accreted to the Alpine wedge during incipient collision. The unit is folded into isoclinal folds that developed on the metre to hectometre scale (Lihou, Reference Lihou1996 a,b). Peak metamorphic temperatures reached c. 320 °C (Ebert et al. Reference Ebert, Herwegh and Pfiffner2007; Lahfid et al. Reference Lahfid, Beyssac, Deville, Negro, Chopin and Goffé2010) (Fig. 1c). The Ultrahelvetic flysch unit contains three groups of mineral veins (G1 to G3) that also occur in the other flysch units and can be found throughout the central and southern part of the study area (Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015). The first group of mineral veins (G1 veins) represents bedding-parallel calcite shear veins that record a top-to-the-NW sense of shear indicating contraction within the flysch units consistent with Alpine shortening directions. G1 veins are folded together with bedding planes, which indicates that these veins formed before folding (Fig. 2a, b). The second and third groups of mineral veins (G2 and G3 veins) comprise quartz-calcite veins, which cross-cut G1 veins and folds, and that show a mutually cross-cutting relationship with the pressure-solution cleavage, that is, the veins cross-cut the cleavage, but are overprinted by ongoing pressure solution (Fig. 2c–e). G2 veins form irregular-shaped lenses that can be several decimetres thick and a few metres long, and often comprise centimetre-sized clasts of brecciated host rock. In comparison, G3 veins form steeply dipping extension fractures that can be several metres long, but are typically thinner than G2 veins and do not include large clasts of brecciated host rocks. Both G2 and G3 veins record brief phases of dilation and extension within the IFUs (Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015).

Fig. 2. Examples of mineral veins sampled in the Globotruncana marl of the Ultrahelvetic flysch unit. (a, b) Bedding-parallel G1 calcite shear veins. G1 veins were folded together with bedding. (c) G2 quartz-calcite vein. The vein contains large clasts of brecciated host rock. (d, e) G3 quartz-calcite extension veins. G3 veins overprint G1 veins and the cleavage. (f) Mineralized tension gashes within fold hinges. The tension gashes record a brittle overprint of the folds. Sampling sites: 46.890° N, 9.153° E and 46.874° N, 9.126° E.

Samples of G1, G2 and G3 veins were taken from a c. 300 m thick marl unit (Globotruncana marl) with a stratigraphic age of Santonian to Campanian (c. 86–72 Ma; Lihou, Reference Lihou1996 a). The marlstone has a relatively homogeneous composition, although the carbonate content varies locally. In addition to G1 to G3 veins, we sampled calcite veins that are structurally linked to folding of the Globotruncana marl. The veins formed in the hinges of isoclinal folds and represent mineralized tension gashes that formed during the final stages of folding (Fig. 2f).

2.b.2. Imbrication of the Ultrahelvetic and South-Helvetic flysch units

The tectonic contacts between the flysch units represent the imbricate thrust faults that were active during accretion (Fig. 1c). The contact between the Ultrahelvetic and the South-Helvetic flysch units is exposed in the central part of the study area, where Ultrahelvetic sandstones are thrust on top of South-Helvetic marlstones (Figs 1c, 3a). The sandstones have a Maastrichtian to Thanetian stratigraphic age (c. 72–56 Ma), while the marlstones have a Lutetian to Priabonian stratigraphic age (c. 48–34 Ma) (Lihou, Reference Lihou1995, Reference Lihou1996 a). At the sampling site, the fault plane is defined by a cataclasite that is a few millimetres to centimetres thick. Calcite extension veins occur sporadically directly above the thrust plane, that is, within the lowermost 10 to 20 cm of the hanging wall (Fig. 3b). The veins contain fragments of the cataclasite suggesting that they formed during the activity of the imbricate thrust fault. We further documented steep-dipping calcite extension veins within the uppermost 10 to 15 m of the footwall, which can be some metres long and several centimetres thick (Fig. 3c, d). Some of the extension veins are cut by small thrust faults. The fault planes of these thrusts are partially mineralized by calcite and show a cataclastic reworking of bedding planes, similar to the imbricate thrust. The structural relationship between these smaller thrusts, the calcite extension veins and the imbricate thrust fault is, however, not evident in the field. To better constrain the sequence of formation of these structures, we sampled all kinds of mineralization, comprising the calcite veins from the foot- and hanging wall, as well as the mineralization on the thrust fault surfaces. We further sampled G1 calcite shear veins and G3 quartz-calcite extension veins, as well as mineralized fissures (that is, open fractures that are only partially mineralized and exhibit euhedral quartz and calcite crystals; Fig. 3e) from the footwall of the imbricate thrust.

Fig. 3. (a) Tectonic contact between Ultrahelvetic flysch (hanging wall) and South-Helvetic flysch (footwall). (b) Example of calcite extension veins formed in the direct hanging wall of the imbricate thrust fault shown in (a). The veins contain fragments of a cataclasite that formed along the imbricate thrust. (c) Example of small thrusts cross-cutting calcite extension veins in the footwall of the imbricate thrust. (d) Example of steep calcite extension veins in the footwall. (e) Mineralized fissure with euhedral quartz and calcite crystals. Sampling site: 46.9597° N, 9.1881° E.

2.b.3. South-Helvetic thrust slice

Thrust slices of South-Helvetic flysch occur in the northernmost part of the study area (Fig. 1a, c). These slices are only a few tens of metres thick and tectonically isolated from the South-Helvetic flysch unit in the central and southern parts of the study area. Peak metamorphic temperatures in the thrust slices reached only 160–180 °C (Rahn et al. Reference Rahn, Mullis and Erdelbrock1995), which indicates that the slices must have been cut off and thrust towards the north before peak temperatures of c. 300 °C were reached in the south. We documented one outcrop within intensively sheared South-Helvetic marlstones of Lutetian to Priabonian age (c. 48–34 Ma) near the village of Weesen (Figs 1a, 4). The marls are thrust on top of shallow marine Molasse sediments of Rupelian age (c. 34–28 Ma, lower marine Molasse) and tectonically overlain by Helvetic carbonates of Cenomanian to Turonian age (c. 100–90 Ma). Shearing in the marls is recorded by closely spaced anastomosing shear bands that define an intense foliation, which is steeply inclined at present (c. 90°; Fig. 4d). Moreover, the marls contain a set of calcite shear veins that formed at a low angle to the foliation (Fig. 4a, d). The veins indicate a sinistral sense of shear and often dip either steeply to the NW or to the SE, which suggests that they represent a set of synthetic Riedel shears. Shear planes within the veins show a tectonic lineation that records former top-to-the-NW thrusting (Dielforder et al. Reference Dielforder, Berger and Herwegh2016) (Fig. 4b). The sheared marlstones further show a typical block-in-matrix fabric, comprising isolated centimetre- to metre-sized intraformational boudins that float within the marlstone matrix (Fig. 4c, d). Some of the boudins are dissected by extensional calcite veins, which formed approximately perpendicular to the shear fabric of the matrix, but are restricted to the boudins. We sampled the calcite shear veins and the extensional veins.

Fig. 4. South-Helvetic thrust slice exposed in northernmost part of the study area. (a) The marl is intensively sheared and comprises long shear veins. (b) Striations (c. 130|45) on shear surfaces indicate top-to-NW shearing (cf. Dielforder et al. Reference Dielforder, Berger and Herwegh2016). (c, d) Examples of boudins dissected by calcite extension veins. Sampling site: 47.1407° N, 9.1073° E.

2.b.4. Thrust faulting in the North-Helvetic flysch unit

In the southern part of the study area, the North-Helvetic flysch unit is locally faulted by thrust faults that are hundreds of metres long and have displacements of a few tens of metres. Thrusting was to the NW (Fig. 5). In contrast to the imbricate thrust fault described above (Section 2.b.2), the hanging wall of these thrusts is intensively fractured in the lowermost 2 to 3 m and exhibits a dense network of extension veins, comprising quartz, calcite and minor amounts of chlorite and white mica. The veins overprint well-compacted and cemented host rocks (turbidites), and we interpret them to have formed during the activity of the thrusts (Fig. 5a, b). The veins are locally overprinted by a set of mineralized fissures that formed at a high angle to the extension veins and comprise euhedral quartz and calcite crystals (Fig. 5a, c). We sampled the extension veins and mineralized fissures.

Fig. 5. (a) Thrust fault in North-Helvetic flysch. The hanging wall of the thrust is intensively fractured. The mineral veins comprise quartz, calcite and minor amounts of chlorite and white mica. (b) Detail of mineral veins within the hanging wall of the thrust shown in (a). The veins overprint well-compacted and foliated rocks. (c) Detail of retrograde fissure overprinting the fault structure and related extension veins. Sampling site: 46.8877° N, 9.1273° E.

3. Methods

Samples of mineral veins were cleaned, trimmed of weathered surfaces and cut into small blocks. These blocks were crushed, washed and sieved. The 125–250 µm grain fraction was immersed in ethanol and c. 100–200 mg of calcite were handpicked under a stereomicroscope. Special care was taken to pick only calcite grains that were devoid of inclusions of host-rock fragments and minerals, such as white mica or chlorite. We further avoided ‘sugar-like’ agglomerations of fine-grained calcite crystals, as these agglomerates most likely comprise recrystallized calcite. The picked grains were washed in Milli-Q water, cleaned in an ultrasonic bath and dried at room temperature overnight. All calcite samples were dissolved in 1 M HCl.

Sr isotope measurements were performed using a TRITON Plus thermal ionization mass spectrometer (TIMS) (Thermo-Fisher) at the mass spectrometer facility of the Institute of Geological Sciences at the University of Bern. The samples were loaded on rhenium ribbon single filaments in combination with a Ta2O5 activator. The measurement commenced when a signal intensity of 5 V on mass 88 was achieved. The 87Sr/86Sr ratios of the samples were normalized to a 86Sr/88Sr ratio of 0.1194 (Nier, Reference Nier1938) using the exponential fractionation law. Samples were also corrected for the offset between the measured 87Sr/86Sr value of SRM987 of the individual session and the 87Sr/86Sr ratio of 0.71024 (Veizer et al. Reference Veizer, Ala, Azmy, Bruckschen, Buhl, Bruhn, Carden, Diener, Ebneth, Godderis, Jasper, Korte, Pawellek, Podlaha and Strauss1999). The external reproducibility (2 s.d.) estimated from the replicate analysis of the standard is ±0.00005 and is taken as uncertainty on the 87Sr/86Sr ratios.

4. Results

The 87Sr/86Sr ratios obtained for the vein carbonates are illustrated in Figure 6 and listed in online Supplementary Material Table S1. Overall, the 87Sr/86Sr ratios fall within the range of c. 0.7075 to 0.7095 (n = 84). In the following, we report the 87Sr/86Sr values obtained for the individual study sites. Note that the 87Sr/86Sr ratios of G1, G2 and G3 veins in the Ultrahelvetic flysch units were originally reported by Dielforder et al. (Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015) and are replicated here for convenience.

Fig. 6. Sr isotopic composition of vein carbonates (crosses) grouped for the different vein generations and sampling sites (a–d). Note the change in scale for 87Sr/86Sr values >0.71. See Figures 25 for examples of analysed mineral veins. The 2σ uncertainties on the 87Sr/86Sr ratios of vein carbonates are smaller than the symbols. Sr ratios of G1, G2 and G3 veins in (a) from Dielforder et al. (Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015). Seawater values at the time of sediment deposition are shown for comparison; data from McArthur et al. (Reference McArthur, Howarth and Bailey2001). Bulk host-rock values in (e) are recalculated to the time of metamorphism at 25 Ma; data compiled from Dielforder et al. (Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015) and Hilgers & Sindern (Reference Hilgers and Sindern2005).

Ultrahelvetic Flysch Unit (Fig. 6a). The 87Sr/86Sr ratios of vein carbonate from G1 calcite shear veins range between 0.7075 and 0.7079 (n = 13). For comparison, the vein carbonates from G2 and G3 quartz-calcite veins show more radiogenic 87Sr/86Sr ratios of 0.7080–0.7086 (n = 12) and 0.7081–0.7089 (n = 12), respectively. The vein carbonates of tension gashes preserved in the hinges of folds have 87Sr/86Sr ratios between 0.7082 and 0.7086 (n = 4), and overlap with the values obtained for G2 and G3 veins.

Imbricate Thrust Fault (Fig. 6b). The vein carbonates from the calcite extension veins in the hanging wall directly above the imbricate thrust fault have 87Sr/86Sr ratios that fall within a narrow range of 0.7082 to 0.7083 (n = 4). Similar values of c. 0.7082 to 0.7083 were obtained for the calcite extension veins (n = 2) and the small thrusts in the footwall (n = 2), respectively. For comparison, the G1 calcite shear veins in the footwall show slightly lower 87Sr/86Sr ratios of 0.7080–0.7081 (n = 2), while the G3 quartz-calcite veins have more radiogenic 87Sr/86Sr values of 0.7084–0.7086 (n = 3). Finally, calcite crystals from the fissures preserved in the footwall show 87Sr/86Sr ratios of 0.7091 to 0.7095 (n = 3).

South-Helvetic Thrust Slice (Fig. 6c). The vein carbonates from the shear veins in the marlstone matrix have 87Sr/86Sr ratios that fall within a narrow range of 0.7082 to 0.7083 (n = 15). Similar values of c. 0.7082 to 0.7084 were obtained for the extension veins from the boudins (n = 7).

Thrust Fault, North-Helvetic Flysch (Fig. 6d). The vein carbonates from the extension veins in the hanging wall directly above the thrust fault have 87Sr/86Sr ratios between 0.7083 and 0.7086 (n = 3). Slightly higher values of c. 0.7087 were obtained for the fissures overprinting the extension veins (n = 2).

5. Discussion

Figure 6 shows the 87Sr/86Sr ratios of vein carbonate together with the 87Sr/86Sr ratio of seawater (McArthur et al. Reference McArthur, Howarth and Bailey2001) at the time of sediment deposition and the 87Sr/86Sr ratios of samples of Ultrahelvetic and North-Helvetic host rocks. The host-rock values are compiled from the literature and back-calculated to 25 Ma, that is, the approximate time of metamorphism in the IFUs (Hunziker et al. Reference Hunziker, Frey, Clauer, Dallmeyer, Friedrichsen, Flehmig, Hochstrasser, Roggwiler and Schwander1986; Hilgers & Sindern, Reference Hilgers and Sindern2005; Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015; Akker et al. Reference Akker, Berger, Zwingmann, Todd, Schrank, Jones, Kewish, Schmid and Herwegh2021 b). The host-rock values are listed in online Supplementary Material Table S2. Overall, the 87Sr/86Sr ratios of vein carbonate plot between the seawater values and the host-rock values. In more detail, we find that the 87Sr/86Sr ratios of vein carbonate correlate with the relative age of the mineral veins, where more radiogenic isotope ratios tend to be associated with mineral veins that formed at a later stage of the structural evolution. This relationship is best expressed by the Sr isotope systematics of the G1 to G3 veins and mineralized fissures, as discussed in the following section. The 87Sr/86Sr ratios of all other vein carbonates are subsequently discussed in Section 5.b below.

5.a. Strontium isotope ratios and pore fluid evolution

G1 veins formed during the earliest structural evolution of the IFUs and are overprinted by all other structural elements, including folds, the pressure-solution cleavage and other mineral veins (Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015, Reference Dielforder, Berger and Herwegh2016) (Fig. 2a, b, d). The Sr isotope ratios of G1 veins are the lowest ratios obtained for all samples and generally closest to the Sr isotope ratios of seawater at the time of sediment deposition (Fig. 6a, b). For the Ultrahelvetic flysch unit, the 87Sr/86Sr ratios of G1 veins partly overlap with the seawater values. By comparison, G2 and G3 veins formed after G1 veins and show more radiogenic 87Sr/86Sr ratios that range between the values of G1 veins and the host rocks (Fig. 6a). Finally, we find that the mineralized fissures, which record the latest stages of vein formation in the IFUs, show the most radiogenic 87Sr/86Sr ratios of all vein carbonates. In detail, the three fissures sampled in the footwall of the imbricate thrust fault that emplaced the Ultrahelvetic flysch on the South-Helvetic flysch show 87Sr/86Sr ratios that overlap with the range of host-rock values and are distinctly higher than the values of nearby G3 veins (Fig. 6b). Similarly, the 87Sr/86Sr ratios of the two fissures sampled along the thrust fault in the North-Helvetic flysch show the highest values obtained for this outcrop (Fig. 6d). Taken together, the Sr isotope ratios of G1 to G3 veins and mineralized fissures indicate an overall increase in 87Sr/86Sr ratios of vein carbonate with the structural evolution of the IFUs, from a seawater-like signature towards the signature of the host rock. We interpret this trend to reflect the diagenetic to metamorphic evolution of pore fluids within the IFUs, as outlined in the following.

The IFUs were deposited on the rifted continental margin of Europe and consisted of biogenic marine carbonates and terrigenous sediments comprising quartz, micas, clays, alkali-feldspars and plagioclase (Frey, Reference Frey1988; Tarantola et al. Reference Tarantola, Mullis, Vennemann, Dubessy and de Capitani2007, Reference Tarantola, Mullis, Guillaume, Dubessy, de Capitani and Abdelmoula2009; Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015; Mullis et al. Reference Mullis, Mählmann and Wolf2017; Akker et al. Reference Akker, Berger, Zwingmann, Todd, Schrank, Jones, Kewish, Schmid and Herwegh2021 b). At the time of sedimentation, the initial pore fluid derived from seawater and thus had a similar 87Sr/86Sr ratio. Likewise, the carbonates consisting of skeletal remnants of marine organisms had a Sr isotopic composition similar to the one of contemporaneous seawater (Milliman, Reference Milliman1974; Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015; El Meknassi et al. Reference El Meknassi, Dera, Cardone, De Rafélis, Brahmi and Chavagnac2018). In contrast, the terrigenous components had a Sr isotopic composition that was on average more radiogenic than the one of seawater (Fig. 6e). Accordingly, the bulk Sr isotopic composition of the flysch sediments varies with the carbonate and silicate content, such that marlstones show less radiogenic Sr ratios than siltstones and shales (Fig. 6e). We further find that the range of 87Sr/86Sr ratios is similar for Ultrahelvetic and North-Helvetic siltstones and shales, which suggests that the Sr isotopic signature of the terrigenous components remained similar throughout the deposition of the three flysch units. In more detail, the Sr isotopic composition of the terrigenous components will have differed for the individual mineral phases. In particular, rubidium-bearing silicates, such as alkali-feldspars, illite, muscovite or biotite, inherited from old continental crust typically contain radiogenic Sr, due to the beta decay of 87Rb to 87Sr (McLennan et al. Reference McLennan, Taylor, McCulloch and Maynard1990; Clauer & Chaudhuri, Reference Clauer and Chaudhuri1995). Thus, the elevated 87Sr/86Sr ratios of the host rocks should mainly come from the sheet silicates and feldspars. In summary, the initial pore fluid was not in isotopic equilibrium with the bulk sediments after the deposition of the IFUs.

With the onset of diagenesis and related sediment alteration, the pore fluid in the IFUs evolved. Microstructural observations from the IFUs indicate that the earliest tectonic deformation affected soft sediments, which suggests that diagenesis occurred mainly syntectonically during incorporation of the flysch units into the Alpine wedge (Dielforder et al. Reference Dielforder, Berger and Herwegh2016). The early diagenetic alteration of calcareous sediments includes the recrystallization of biogenic calcite to secondary calcite (Milliman, Reference Milliman1974; Elderfield et al. Reference Elderfield, Gieskes, Baker, Oldfield, Hawkesworth and Miller1982). Carbonate diagenesis typically takes place at temperatures below 100 °C and involves the release of Sr into the pore fluid from the biogenic carbonates, whose Sr concentration is one to three orders of magnitude higher than in seawater (Richter & Liang, Reference Richter and Liang1993; Fantle & DePaolo, Reference Fantle and DePaolo2006; Voigt et al. Reference Voigt, Hathorne, Frank, Vollstaedt and Eisenhauer2015). The Sr release buffers the isotopic composition of the pore fluid during carbonate diagenesis. We therefore suggest that the seawater-like 87Sr/86Sr ratios of G1 veins record a formation of these veins during carbonate diagenesis and early tectonic burial.

Another early diagenetic process is the transformation of smectite to illite, which occurs between 80 and 150 °C in most geological settings (Pytte & Reynolds, Reference Pytte, Reynolds, Naeser and McCulloh1989; Awwiller, Reference Awwiller1993; Moore & Saffer, Reference Moore and Saffer2001). Although young authigenic smectites may lack relevant amounts of radiogenic 87Sr, the smectite-to-illite transformation is accompanied by alkali-feldspar albitization, which releases Sr with elevated 87Sr/86Sr ratios into the pore fluid (Awwiller, Reference Awwiller1993; Baccar et al. Reference Baccar, Fritz and Made1993; Clauer et al. Reference Clauer, Środoń, Aubert, Uysal and Toulkeridis2020). Additionally, the recrystallization of detrital mixed-layer illite-smectite is likely to release more 87Sr into the pore fluid. We therefore expect that the Sr isotopic signature of the pore fluid becomes more radiogenic during the smectite-to-illite transformation. It remains difficult to quantify the change in Sr isotopic composition, which would require detailed information on fluid volumes, fluid fluxes, Sr concentrations and isotope ratios, but the process may be constrained for the IFUs by the following considerations. First, sediment compaction within the upper 4 to 5 km of accretionary wedges results in a strong reduction in porosity and expulsion of pore fluids (e.g. Wang, Reference Wang1994; Moore & Saffer, Reference Moore and Saffer2001). Dielforder et al. (Reference Dielforder, Berger and Herwegh2016) modelled the compaction and diagenetic fluid release for the IFUs and found that the smectite-to-illite transformation occurred after the main compaction of the sediments, when much of the initial pore fluid was already expelled. This suggests that the smectite-to-illite transformation and related fluid liberation resulted in a substantial alteration of the initial pore fluid. Second, the formation temperatures of G2 and G3 quartz-calcite veins from the Ultrahelvetic flysch unit have been previously constrained by oxygen isotope thermometry to c. 210–290 °C (Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015) (Fig. 7). Thus, the quartz-calcite veins formed after the smectite-to-illite transformation, which suggests that vein carbonates with 87Sr/86Sr ratios lower than those of G2 and G3 veins but higher than those for seawater reflect the approximate Sr isotopic signature of the pore fluid during illitization.

Fig. 7. Diagram showing the positive correlation between the 87Sr/86Sr ratios of G2 and G3 veins from the Ultrahelvetic flysch unit and the formation temperature of these veins as constrained by oxygen isotope thermometry (data from Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015). R is the correlation coefficient. The range of 87Sr/86Sr ratios obtained for nearby tension gashes overlaps with the values of G2 and G3 veins, which suggests an approximate formation temperature of 230–260 °C for the tension gashes as indicated by dashed lines. See Section 5.b.1 for details.

After the smectite-to-illite transformation, the Sr isotopic composition of the pore fluid continued to develop towards the signature of the host rock, as illustrated by the 87Sr/86Sr ratios of G2 and G3 veins from the Ultrahelvetic flysch unit, which increase with the vein formation temperature (Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015) (Fig. 7). This development towards more radiogenic values could in principle reflect two processes: the release of Sr into the pore fluid due to continuous recrystallization of minerals (e.g. Glodny et al. Reference Glodny, Kühn and Austrheim2008), and temperature-dependent diffusion of Sr (e.g. Dodson, Reference Dodson1973). Because of the high Ar and Sr retentivity of micas (Villa, Reference Villa1998) and because the dissolution-reprecipitation rate at 200 °C is around ten orders of magnitude faster than volume diffusion (Villa, Reference Villa2016), we are forced to conclude that the increase in the 87Sr/86Sr ratios was not primarily controlled by Sr diffusion. In contrast, the recrystallization of micas at low-grade metamorphic conditions is well documented also for the flysch units. Akker et al. (Reference Akker, Berger, Schrank, Jones, Kewish, Klaver and Herwegh2021 a, b) showed that the development of the pressure-solution cleavage in the IFUs at temperatures above 230 °C involved the partial recrystallization of white micas. Tarantola et al. (Reference Tarantola, Mullis, Vennemann, Dubessy and de Capitani2007, Reference Tarantola, Mullis, Guillaume, Dubessy, de Capitani and Abdelmoula2009) documented the chloritization of detrital biotites in the North-Helvetic flysch at temperatures of c. 270 °C. After the smectite-illite transformation and temperatures above 200 °C, illite is successively transformed into muscovite (Hunziker et al. Reference Hunziker, Frey, Clauer, Dallmeyer, Friedrichsen, Flehmig, Hochstrasser, Roggwiler and Schwander1986; van de Kamp, Reference van de Kamp2008). We therefore argue that the increase in 87Sr/86Sr ratios primarily reflects a release of 87Sr into the pore fluid due to the recrystallization of minerals and associated mass transfer processes.

Taken together, the diagenetic and low-grade metamorphic processes discussed above contributed to the Sr isotopic equilibration of the flysch units. Recently, Akker et al. (Reference Akker, Berger, Zwingmann, Todd, Schrank, Jones, Kewish, Schmid and Herwegh2021 b) evaluated the grain-size dependent isotopic resetting of the K–Ar system for white micas from the IFUs and found that the system is only partially reset, except for the southernmost part of the study area (T max ≈ 330 °C), where the grain-size fraction of ≤0.8 µm is nearly reset. We therefore propose that the Rb–Sr system was also only partially reset and that the late-stage Sr isotopic evolution of the pore fluid was dominated by the interaction with silicates rather than with the bulk rock, which would have required a higher metamorphic grade and full recrystallization of the sediments. Independently, the bulk Sr isotopic evolution of the pore fluid reflects the progressive isotopic equilibration and is recorded by the vein carbonates. It is important to note that the pore fluid evolution was not synchronous within the different flysch units, as the flysch units were accreted successively to the Alpine wedge (that is, the Ultrahelvetic flysch unit was already accreted and experienced syntectonic diagenesis, while the North-Helvetic flysch was still being deposited in the foreland basin). Moreover, the detailed Sr isotopic evolution will have varied between the units and on smaller spatial scales, owing to differences in the initial Sr isotopic composition of seawater and sediments (Fig. 6). The absolute 87Sr/86Sr values of vein carbonates from different units can therefore not be directly compared. Likewise, small differences in the 87Sr/86Sr ratio of nearby mineral veins do not necessarily reflect the sequence of vein formation, as evident from the scatter in Figure 7. Independent of these restrictions, the overall Sr isotope systematics of the vein carbonates allow us to constrain the approximate diagenetic to low-grade metamorphic conditions of vein formation and can therefore allow a better understanding of aspects of the structural and tectonic evolution, as discussed in the following section.

5.b. Implications for the structural and tectonic evolution of the Infrahelvetic flysch units

In addition to the G1 to G3 veins and mineralized fissures discussed in the previous section, we sampled mineral veins that relate to distinct structural and tectonic features (Section 2.b above) and help to understand their genesis and sequence of formation.

5.b.1. Folding

The tension gashes in the hinge zone of isoclinal folds in the Ultrahelvetic flysch unit document a brittle overprint of the folds and thus formed after or during the latest stage of folding. The vein carbonates from the tension gashes have 87Sr/86Sr ratios that overlap with Sr ratios of nearby G2 and G3 veins, which indicates that the gashes and veins formed at similar diagenetic-metamorphic conditions (Fig. 6a). In addition, the 87Sr/86Sr ratios of the G2 and G3 veins exhibit a strong positive correlation with the vein formation temperatures and increase from c. 0.7080 and 210 °C to c. 0.7089 and 290 °C (Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015) (Fig. 7). Transferred to the 87Sr/86Sr ratios obtained for the tension gashes, this correlation between the Sr ratios and temperature suggests that the tension gashes formed at c. 230–260 °C (Fig. 7). This finding is consistent with the structural constraints, which indicate that folding in the IFUs initiated during the early structural evolution and at temperatures below 160 °C, and that the folds were subsequently tightened and overprinted by the pressure-solution cleavage, which mainly developed above 230 °C (Dielforder et al. Reference Dielforder, Berger and Herwegh2016). Instead, the classical interpretation was that folding and cleavage development in the IFUs represent one (and the main) deformation event (e.g. Schmid, Reference Schmid1975; Milnes & Pfiffner, Reference Milnes and Pfiffner1977).

5.b.2. Imbricate thrusting

The calcite extension veins that formed in the immediate hanging wall of the imbricate thrust emplacing the Ultrahelvetic flysch on the South-Helvetic flysch have similar 87Sr/86Sr ratios as the vein carbonate from the extension vein and the small thrusts in the footwall of the imbricate fault (Fig. 6b). This relationship suggests that the structures formed at similar times and that the small thrusts may represent splays from the main fault. Likewise, the calcite extension veins in the footwall may record damage in the surrounding of the imbricate thrust. Notably, all 87Sr/86Sr ratios of vein carbonate related to imbrication are more radiogenic than the Sr ratios of nearby G1 veins, which suggests that the G1 veins record shearing before imbrication (Fig. 6b). Moreover, the 87Sr/86Sr ratios are distinctly more radiogenic than seawater at the time of sedimentation (note that the seawater values are similar for the sediments in the hanging wall and footwall despite their different ages; Fig. 6b), but less radiogenic than the 87Sr/86Sr ratios of nearby G3 quartz-calcite veins and mineralized fissures. Thus, the Sr ratios suggest that the imbrication of the units occurred at c. 100–150 °C, that is, after carbonate diagenesis but before quartz cementation, which typically becomes active in accretionary settings at 150–200 °C (e.g. Moore et al. Reference Moore, Rowe, Meneghini, Dixon and Moore2007). This temperature constraint is consistent with the cataclasis along the thrust faults and the off-fault damage, which both suggest that the rocks were already consolidated and had some elastic strength during faulting (e.g. Sibson, Reference Sibson1977; Fagereng & Toy, Reference Fagereng, Toy, Fagereng and Toy2011).

5.b.3. South-Helvetic thrust slice

Vein carbonates from the South-Helvetic thrust slice preserved in the northernmost part of the study area document similar 87Sr/86Sr ratios for the calcite shear veins and the calcite extension veins dissecting the boudins (Fig. 6c). We interpret these Sr ratios to indicate that the shearing and formation of the tectonic mélange occurred simultaneously and was related to the transport of the thrust slice, which agrees with concepts for mélange formation within shear zones (Remitti et al. Reference Remitti, Bettelli and Vannucchi2007; Fagereng & Sibson, Reference Fagereng and Sibson2010; Festa et al. Reference Festa, Pini, Dilek and Codegone2010). Interestingly, the veins in the South-Helvetic thrust slice have similar Sr ratios to vein carbonates related to imbricate thrusting at the Ultrahelvetic flysch – South-Helvetic flysch contact. As the sediments in the thrust slice have a similar age and composition to the sediments in the footwall of the imbricate thrust (Lihou, Reference Lihou1995, Reference Lihou1996 a), we expect that the pore fluid evolution was similar for both localities. The similarity in the Sr ratios therefore suggests that both imbricate thrusting and the tectonic transport of the South-Helvetic thrust slice occurred at diagenetic conditions, which is consistent with the peak temperatures of c. 160–180 °C inferred for the South-Helvetic thrust slice (Rahn et al. Reference Rahn, Mullis and Erdelbrock1995).

5.b.4. Late-stage thrust faulting

Vein carbonates from the thrust fault in the North-Helvetic flysch show elevated 87Sr/86Sr ratios (Fig. 6d). In detail, the Sr ratios of the mineralized fissures are slightly more radiogenic than the Sr ratios of the quartz-calcite extension veins that are overprinted by the fissures, which is consistent with the general pore fluid evolution towards a more radiogenic signature. The radiogenic signature of the vein carbonates suggests that the thrust was active at or close to peak metamorphic conditions, which would be consistent with structural constraints indicating that the thrust overprints well-compacted and foliated rocks (Dielforder et al. Reference Dielforder, Berger and Herwegh2016). The intriguing aspect of this activity is that the thrust documents late-stage frictional faulting in the North-Helvetic flysch. The activity of the thrust may therefore be related to the relocation of the deformation into the crystalline basement of the Aar massif and faulting near the basement–cover contact (Pfiffner, Reference Pfiffner, Allen and Homewood1986; Nibourel et al. Reference Nibourel, Berger, Egli, Heuberger and Herwegh2021) (Fig. 1). In this case, thrust faulting took place during the earliest retrograde evolution of the IFUs and was potentially related to the early doming and exhumation of the Aar massif.

5.c. External fluids

We interpret the trend in the 87Sr/86Sr ratios of vein carbonates to document the diagenetic to metamorphic evolution of the pore fluid due to local fluid–rock interaction. This pore fluid evolution could have been disturbed by an influx of external fluids with a different Sr isotopic composition.

In fold-and-thrust belts and accretionary wedges, external fluids can migrate along major faults, including the decollement or megathrust, and major splay faults such as imbricate or out-of-sequence thrusts (McCaig et al. Reference McCaig, Wayne, Marshall, Banks and Henderson1995; Machel et al. Reference Machel, Cavell and Patey1996; Travé et al. Reference Travé, Labaume, Calvet and Soler1997; Lauer & Saffer, Reference Lauer and Saffer2015; Cerchiari et al. Reference Cerchiari, Remitti, Mittempergher, Festa, Lugli and Cipriani2020). Fluid flow along the Glarus thrust has been studied by means of radiogenic Sr isotope and stable isotope (δ13C, δ18O) systematics (Burkhard et al. Reference Burkhard, Kerrich, Maas and Fyfe1992; Abart et al. Reference Abart, Badertscher, Burkhard and Povoden2002; Badertscher et al. Reference Badertscher, Abart, Burkhard and McCaig2002). These studies inferred a syntectonic flow of external, basement-derived metamorphic fluids along the Glarus thrust for the southernmost part of the study area, where the thrust overlies Mesozoic marine carbonates. The metamorphic fluid has been related to high 87Sr/86Sr ratios of >0.71 recorded in calcmylonites that formed along the thrust plane. In addition, fluid infiltration into the uppermost metres of the footwall caused significant 18O depletion in the limestones. In contrast, where the Glarus thrust overlies flysch sediments, fluid flow along the thrust was found to be consistent with fluids sourced from the flysch units, suggesting that there was a net flux of fluids from the flysch towards the Glarus thrust (Burkhard et al. Reference Burkhard, Kerrich, Maas and Fyfe1992; Abart et al. Reference Abart, Badertscher, Burkhard and Povoden2002; Badertscher et al. Reference Badertscher, Abart, Burkhard and McCaig2002). This finding is consistent with the stable isotope systematics of calcite cements and vein carbonates from the Ultrahelvetic flysch unit, which indicate that the calcite cements bear a marine equilibrated stable isotope signature, and that the vein carbonates precipitated from a local, rock-buffered fluid (Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015). It has been further shown that the deuterium isotope values of fluid inclusions in quartz from mineralized fissures from the North-Helvetic flysch unit are in equilibrium with sheet silicates of the host rocks and that the fluid inclusions contain locally sourced methane (Tarantola et al. Reference Tarantola, Mullis, Vennemann, Dubessy and de Capitani2007; Mangenot et al. Reference Mangenot, Tarantola, Mullis, Girard, Le and Eiler2021) These findings also indicate vein precipitation from a rock-buffered fluid. We therefore suppose that the 87Sr/86Sr ratios of vein carbonates reported in this study do not record external fluids that infiltrated from the Glarus thrust into the flysch units.

The imbricate thrust emplacing the Ultrahelvetic flysch on the South-Helvetic flysch likely rooted in the decollement and may have channelled diagenetic to metamorphic fluids migrating upwards along the basal fault (e.g. Machel et al. Reference Machel, Cavell and Patey1996; Lauer & Saffer, Reference Lauer and Saffer2015; Cerchiari et al. Reference Cerchiari, Remitti, Mittempergher, Festa, Lugli and Cipriani2020). The 87Sr/86Sr ratios of vein carbonates related to imbricate thrusting are much less radiogenic than the values obtained, for example, by Burkhard et al. (Reference Burkhard, Kerrich, Maas and Fyfe1992) for the calcmylonites. We therefore exclude that the imbricate thrust channelled relevant amounts of metamorphic fluids. On the other hand, migrating diagenetic fluids may have had Sr ratios similar to the one recorded in the vein carbonates, as the decollement was located within flysch sediments of similar composition and stratigraphic age (Pfiffner, Reference Pfiffner, Allen and Homewood1986). We can therefore not exclude the migration of nearby diagenetic fluids along the imbricate thrust. Independently, the general increase towards more radiogenic Sr ratios documented by the different vein generations that formed before (G1 veins) and after imbricate thrusting (G3 veins, fissures) would still reflect the overall diagenetic to metamorphic evolution of the pore fluid, similar to the development in the Ultrahelvetic flysch.

The thrust slice of South-Helvetic marls in the northern part of the study area is tectonically intercalated between Oligocene Molasse sediments and Upper Cretaceous marine limestones, providing both units as potential sources for external fluids. Fluids derived from the Cretaceous limestones in the hanging wall must be expected to show 87Sr/86Sr ratios close to the one of contemporaneous seawater, that is, c. 0.7073–0.7074 (McArthur et al. Reference McArthur, Howarth and Bailey2001). The values are significantly lower than for contemporaneous seawater of the South-Helvetic marls (Fig. 6c). A significant influx of fluids from the Cretaceous limestone should therefore result in low 87Sr/86Sr ratios, which is not recorded in the Sr isotopes of the mineral veins. We therefore exclude the limestones as the fluid source. In contrast, the 87Sr/86Sr ratios of seawater during the deposition of the shallow marine Molasse sediments are only slightly higher and partly overlap with the one for the South-Helvetic marls (Fig. 6c). An influx of diagenetic fluids from the Molasse sediments may therefore be difficult to detect in the Sr ratios of vein carbonates. However, the shearing and mélange formation should have mainly occurred during the tectonic transport of the thrust slice and before its emplacement on the Molasse sediments. We therefore exclude fluids expelled from the Molasse as a relevant source for the vein carbonates.

6. Summary and conclusions

The 87Sr/86Sr ratios of vein carbonates from the IFUs indicate a consistent increase from a seawater-like signature recorded by early veins towards the signature of the host rock recorded by late veins. We interpret these values as tracing the bulk diagenetic to low-grade metamorphic pore fluid evolution, which was governed by carbonate diagenesis (≤100 °C), the smectite-to-illite transformation and alkali-feldspar albitization (c. 80–150 °C), and the progressive recrystallization of sheet silicates (≥200 °C). Figure 8 shows the inferred pore fluid evolution for the Ultrahelvetic and South-Helvetic flysch units as a function of temperature and diagenetic-metamorphic processes. Notably, the trends for the flysch units are not identical, which most likely reflects differences in seawater signature and sedimentological composition. Independently from this variation, the bulk pore fluid evolution appears to be unaffected by the influx of external metamorphic fluids. A local mixture of diagenetic fluids with similar 87Sr/86Sr ratios, for example, during imbricate thrusting, can, however, not be excluded.

Fig. 8. Diagram illustrating the Sr isotopic evolution of pore fluids in the Ultrahelvetic flysch and South-Helvetic flysch units. The increase from the seawater-like signature towards more radiogenic values is interpreted to relate to diagenetic and metamorphic processes, including carbonate diagenesis, smectite-to-illite transformation, alkali-feldspar albitization, illite-to-muscovite transformation and recrystallization (recryst.) of detrital white mica and biotite. See Sections 5.a and 6 for details.

Our analysis shows that the Sr isotope systematics of vein carbonates sampled along a 30 km long transect and from a variety of outcrops, rock units and structures can provide consistent information that helps to understand the evolution of the thrust belt and the sequence of formation of distinct deformation structures, which was not sufficiently clear from field observations. In particular, our findings indicate that the emplacement of the Ultrahelvetic flysch onto the South-Helvetic flysch as well as the transport of the South-Helvetic thrust slice and related mélange formation occurred at diagenetic conditions of c. 100–150 °C. This suggests that the Ultrahelvetic and South-Helvetic flysch units were first underthrust and then accreted to the base of the wedge at some kilometres depth (Fig. 9a). Shearing during underthrusting is probably recorded by G1 calcite shear veins that document a less-evolved fluid than the veins related to imbricate thrusting. After or during imbrication, slices of South-Helvetic flysch were displaced towards the foreland, probably by dragging the slices along the Penninic basal thrust that accommodated the underthrusting of the European continental margin until early Miocene time (e.g. Cardello et al. Reference Cardello, Di Vincenzo, Giorgetti, Zwingmann and Mancktelow2019). Out-of-sequence thrusting along the Glarus thrust and the related emplacement of the Helvetic nappes on top of the IFUs resulted in peak metamorphic conditions (Fig. 9b). At this time, the flysch units were already folded and the folds experienced a late-stage brittle overprint, as indicated by the tension gashes. Otherwise, deformation was mainly related to the development of the pressure-solution cleavage and the formation of quartz-calcite veins (G2 and G3, veins, fissures). Finally, small thrusts developed in the North-Helvetic flysch during the late prograde to early retrograde evolution of the fold-and-thrust belt, which was potentially related to incipient deformation in the Aar massif.

Fig. 9. Schematic cross-section illustrating the (a) diagenetic to (b) metamorphic evolution of the Infrahelvetic flysch units during the Alpine orogeny. LFM – Lower Freshwater Molasse; LMM – Lower Marine Molasse; NHF – North-Helvetic flysch; SHF – South-Helvetic flysch; UHF – Ultrahelvetic flysch. The blue arrows indicate fluid flow; the blue spirals indicate internal fluid evolution. Based on Burkhard et al. (Reference Burkhard, Kerrich, Maas and Fyfe1992), Dielforder et al. (Reference Dielforder, Berger and Herwegh2016) and Pfiffner (Reference Pfiffner, Allen and Homewood1986). Not to scale.

Taken together, we have demonstrated that the Sr isotope systematics of vein carbonates can help to constrain the relative timing of deformation events within the tectonic and structural evolution of a fold-and-thrust belt. Our study benefited from good outcrop conditions and availability of data about the diagenetic to low-grade metamorphic evolution of the thrust belt. The presented approach may also be applicable to other thrust belts or to geological settings, with less structural control, such as active geothermal reservoirs or accretionary wedges. The Sr isotope systematics of vein carbonates recovered from drill cores from such settings may provide an easy means to disentangle deformation sequences and to identify fluid sources.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0016756821001357

Acknowledgements

Funding for this project was provided by the Swiss National Science Foundation (No. 144381). We thank H. Vollstaedt for discussion and support in the laboratory. The thoughtful and constructive reviews by Johannes Glodny and one anonymous reviewer are gratefully acknowledged. The authors declare that there is no conflict of interest.

References

Abart, R, Badertscher, N, Burkhard, M and Povoden, E (2002) Oxygen, carbon and strontium isotope systematics in two profiles across the Glarus thrust: implications for fluid flow. Contributions to Mineralogy and Petrology 143, 192208.Google Scholar
Akker, IV, Berger, A, Schrank, CE, Jones, MWM, Kewish, CM, Klaver, J and Herwegh, M (2021a) The evolution of slate microfabrics during progressive accretion of foreland basin sediments. Journal of Structural Geology 150, 104404. doi: 10.1016/j.jsg.2021.104404.CrossRefGoogle Scholar
Akker, IV, Berger, A, Zwingmann, H, Todd, A, Schrank, CE, Jones, MWM, Kewish, CM, Schmid, TC and Herwegh, M (2021b) Structural and chemical resetting processes in white mica and their effect on K–Ar data during low temperature metamorphism. Tectonophysics 800, 228708. doi: 10.1016/j.tecto.2020.228708.CrossRefGoogle Scholar
Awwiller, DN (1993) Illite/smectite formation and potassium mass transfer during burial diagenesis of mudrocks: a study from the Texas gulf coast Paleocene–Eocene. Journal of Sedimentary Petrology 63, 501–12.Google Scholar
Baccar, MB, Fritz, B and Made, B (1993) Diagenetic albitization of K-feldspar and plagioclase in sandstone reservoirs: thermodynamic and kinetic modeling. Journal of Sedimentary Research 63, 1100–9.Google Scholar
Badertscher, NP, Abart, R, Burkhard, M and McCaig, A (2002) Fluid flow pathways along the Glarus overthrust derived from stable and Sr-isotope patterns. American Journal of Science 302, 517–47.CrossRefGoogle Scholar
Beaudoin, N, Lacombe, O, Roberts, NMW and Koehn, D (2018) U–Pb dating of calcite veins reveals complex stress evolution and thrust sequence in the Bighorn Basin, Wyoming, USA. Geology 46, 1015–18.CrossRefGoogle Scholar
Bons, PD, Elburg, MA and Gomez-Rivas, E (2012) A review of the formation of tectonic veins and their microstructures. Journal of Structural Geology 43, 3362.CrossRefGoogle Scholar
Burkhard, M (1990) Aspects of the large-scale Miocene deformation in the most external part of the Swiss Alps (Subalpine Molasse to Jura fold belt). Eclogae Geologicae Helvetiae 83, 559–83.Google Scholar
Burkhard, M, Kerrich, R, Maas, R and Fyfe, WS (1992) Stable and Sr-isotope evidence for fluid advection during thrusting of the Glarus nappe (Swiss Alps). Contributions to Mineralogy and Petrology 112, 293311.CrossRefGoogle Scholar
Cardello, GL, Di Vincenzo, G, Giorgetti, G, Zwingmann, H and Mancktelow, N (2019) Initiation and development of the Pennine Basal Thrust (Swiss Alps): a structural and geochronological study of an exhumed megathrust. Journal of Structural Geology 129, 338–56.CrossRefGoogle Scholar
Cerchiari, A, Remitti, F, Mittempergher, S, Festa, A, Lugli, F and Cipriani, A (2020) Cyclical variations of fluid sources and stress state in a shallow megathrust-zone mélange. Journal of the Geological Society, London 177, 647–59.CrossRefGoogle Scholar
Clauer, N and Chaudhuri, S (1995) Clays in Crustal Environments: Isotope Dating and Tracing. New York: Springer, 369 pp.CrossRefGoogle Scholar
Clauer, N, Środoń, J, Aubert, A, Uysal, IT and Toulkeridis, T (2020) K–Ar and Rb–Sr dating of nanometer-sized smectite-rich mixed layers from bentonite beds of the Campos basin (Rio de Janeiro State, Brazil). Clays and Clay Minerals 68, 446–64.CrossRefGoogle Scholar
Dahlen, FA (1990) Critical taper model of fold-and-thrust belts and accretionary wedges. Annual Review of Earth and Planetary Sciences 18, 5599.CrossRefGoogle Scholar
Dielforder, A, Berger, A and Herwegh, M (2016) The accretion of foreland basin sediments during early stages of continental collision and similarities to accretionary wedge tectonics. Tectonics 35, 2216–38.CrossRefGoogle Scholar
Dielforder, A, Vollstaedt, H, Vennemann, T, Berger, A and Herwegh, M (2015) Linking megathrust earthquakes to brittle deformation in a fossil accretionary complex. Nature Communications 6, 7504. doi: 10.1038/ncomms8504.CrossRefGoogle Scholar
Dietrich, D, McKenzie, J and Song, H (1983) Origin of calcite in syntectonic veins as determined from carbon isotope ratios. Geology 11, 547–51.2.0.CO;2>CrossRefGoogle Scholar
Dodson, MH (1973) Closure temperature in cooling geochronological and petrological systems. Contributions to Mineralogy and Petrology 40, 259–74.CrossRefGoogle Scholar
Ebert, A, Herwegh, M and Pfiffner, A (2007) Cooling induced strain localization in carbonate mylonites within a large-scale shear zone (Glarus thrust, Switzerland). Journal of Structural Geology 29, 1164–84.CrossRefGoogle Scholar
El Meknassi, S, Dera, G, Cardone, T, De Rafélis, M, Brahmi, C and Chavagnac, V (2018) Sr isotope ratios of modern carbonate shells: good and bad news for chemostratigraphy. Geology 46, 1003–6.CrossRefGoogle Scholar
Elderfield, H, Gieskes, JM, Baker, PA, Oldfield, RK, Hawkesworth, CJ and Miller, R (1982) 87Sr/86Sr and 18O/16O ratios, interstitial water chemistry and diagenesis in deep-sea carbonate sediments of the Ontong Java Plateau. Geochimica et Cosmochimica Acta 46, 2259–68.CrossRefGoogle Scholar
Fagereng, Å and Sibson, RH (2010) Mélange rheology and seismic style. Geology 38, 751–4.CrossRefGoogle Scholar
Fagereng, Å and Toy, VG (2011) Geology of the earthquake source: an introduction. In Geology of the Earthquake Source (eds Fagereng, Å and Toy, VG), pp. 116. Geological Society of London, Special Publication no. 359.Google Scholar
Fantle, MS and DePaolo, DJ (2006) Sr isotopes and pore fluid chemistry in carbonate sediment of the Ontong Java Plateau: calcite recrystallization rates and evidence for a rapid rise in seawater Mg over the last 10 million years. Geochimica et Cosmochimica Acta 70, 3883–904.CrossRefGoogle Scholar
Festa, A, Pini, GA, Dilek, Y and Codegone, G (2010) Mélanges and mélange-forming processes: a historical overview and new concepts. International Geology Review 52, 1040–105.CrossRefGoogle Scholar
Fisher, D and Byrne, T (1987) Structural evolution of underthrusted sediments, Kodiak Islands, Alaska. Tectonics 6, 775–93.CrossRefGoogle Scholar
Ford, M and Lickorish, H (2004) Foreland basin evolution around the western Alpine Arc. In Deep-water Sedimentation in the Alpine Basin of SE France: New Perspectives on the Gres d’Annot and Related Systems (eds Joseph, P and Lomas, SA), pp. 3963. Geological Society of London, Special Publication no. 221.Google Scholar
Frey, M (1988) Discontinuous inversed metamorphic zonation, Glarus Alps, Switzerland: evidence from illite crystallinity data. Schweizerische Mineralogische und Petrographische Mitteilungen 86, 171–84.Google Scholar
Frisch, W (1979) Tectonic progradation and plate tectonic evolution of the Alps. Tectonophysics 60, 121–39.CrossRefGoogle Scholar
Gasser, D and den Brok, B (2008) Tectonic evolution of the Engi Slates, Glarus Alps, Switzerland. Swiss Journal of Geosciences 101, 311–22.CrossRefGoogle Scholar
Glodny, J, Kühn, A and Austrheim, H (2008) Diffusion versus recrystallization processes in Rb–Sr geochronology: isotopic relics in eclogite facies rocks, Western Gneiss Region, Norway. Geochimica et Cosmochimica Acta 72, 506–25.CrossRefGoogle Scholar
Glotzbach, C, Reinecker, J, Danišík, M, Rahn, M, Frisch, W and Spiegel, C (2010) Thermal history of the central Gotthard and Aar massifs, European Alps: evidence for steady state long-term exhumation. Journal of Geophysical Research: Solid Earth 115, F03017. doi: 10.1029/2009JF001304.CrossRefGoogle Scholar
Handy, MR, 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, 121–58.CrossRefGoogle Scholar
Herwegh, M, Berger, A, Baumberger, R, Wehrens, P and Kissling, E (2017) Large-scale crustal-block-extrusion during late Alpine collision. Scientific Reports 7, 413. doi: 10.1038/s41598-017-00440-0.CrossRefGoogle ScholarPubMed
Herwegh, M, Berger, A, Glotzbach, C, Wangenheim, C, Mock, S, Wehrens, P, Baumberger, R, Egli, D and Kissling, E (2020) Late stages of continent-continent collision: timing, kinematic evolution, and exhumation of the Northern rim (Aar Massif) of the Alps. Earth-Science Reviews 200, 102959. doi: 10.1016/j.earscirev.2019.102959.CrossRefGoogle Scholar
Herwegh, M, Hürzeler, J-P, Pfiffner, OA, Schmid, SM, Abart, R and Ebert, A (2008) The Glarus thrust: excursion guide and report of a field trip of the Swiss Tectonic Studies Group (Swiss Geological Society, 14. – 16.09.2006). Swiss Journal of Geosciences 101, 323–40.CrossRefGoogle Scholar
Hilgers, C and Sindern, S (2005) Textural and isotopic evidence on the fluid source and transport mechanism of antitaxial fibrous microstructures from the Alps and the Appalachians. Geofluids 5, 239–50.CrossRefGoogle Scholar
Hubbert, MK and Rubey, WW (1959) Role of fluid pressure in mechanics of overthrust faulting. 1. Mechanics of fluid-filled porous solids and its application to overthrust faulting. Bulletin of the Geological Society of America 70, 115–66.Google Scholar
Hunziker, JC, Frey, M, Clauer, N, Dallmeyer, RD, Friedrichsen, H, Flehmig, W, Hochstrasser, K, Roggwiler, P and Schwander, H (1986) The evolution of illite to muscovite; mineralogical and isotopical data from the Glarus Alps, Switzerland. Contributions to Mineralogy and Petrology 92, 157–80.CrossRefGoogle Scholar
Jeanbourquin, P (1994) Early deformation of Ultrahelvetic mélanges in the Helvetic nappes (western Swiss Alps). Journal of Structural Geology 16, 1367–83.CrossRefGoogle Scholar
Kempf, O and Pfiffner, OA (2004) Early Tertiary evolution of the North Alpine Foreland Basin of the Swiss Alps and adjoining areas. Basin Research 16, 549–67.CrossRefGoogle Scholar
Krabbenhöft, A, Fietzke, J, Eisenhauer, A, Liebetrau, V, Böhm, F and Vollstaedt, H (2009) Determination of radiogenic and stable strontium isotope ratios (87Sr/86Sr; δ88/86Sr) by thermal ionization mass spectrometry applying an 87Sr/86Sr double spike. Journal of Analytical Atomic Spectrometry 24, 1267–71.CrossRefGoogle Scholar
Labaume, P, Berty, C and Laurent, Ph (1991) Syn-diagenetic evolution of shear structures in superficial nappes: an example from the Northern Apennines (NW Italy). Journal of Structural Geology 18, 385–98.CrossRefGoogle Scholar
Lacroix, B, Buatier, M, Labaume, P, Travé, A, Dubois, M, Charpentier, D, Ventalon, S and Convert-Gaubier, D (2011) Microtectonic and geochemical characterization of thrusting in a foreland basin: example of the South-Pyrenean orogenic wedge (Spain). Journal of Structural Geology 33, 1359–77.CrossRefGoogle Scholar
Lahfid, A, Beyssac, O, Deville, E, Negro, F, Chopin, C and Goffé, B (2010) Evolution of the Raman spectrum of carbonaceous material in low-grade metasediments of the Glarus Alps (Switzerland). Terra Nova 22, 354–60.CrossRefGoogle Scholar
Lauer, RM and Saffer, DM (2015) The impact of splay faults on fluid flow, solute transport, and pore pressure distribution in subduction zones: a case study offshore the Nicoya Peninsula, Costa Rica. Geochemistry, Geophysics, Geosystems 16, 1089–104.CrossRefGoogle Scholar
Lihou, JC (1995) A new look at the Blattengrat unit of eastern Switzerland: Early Tertiary foreland basin sediments from the South Helvetic realm. Eclogae Geologicae Helvetiae 88, 91114.Google Scholar
Lihou, JC (1996a) Stratigraphy and sedimentology of the Sardona unit, Glarus Alps: Upper Cretaceous/middle Eocene deep-marine flysch sediments from the Ultrahelvetic realm. Eclogae Geologicae Helvetiae 89, 721–52.Google Scholar
Lihou, JC (1996b) Structure and deformational history of the Infrahelvetic flysch units, Glarus Alps, eastern Switzerland. Eclogae Geologicae Helvetiae 89, 439–60.Google Scholar
Machel, HG, Cavell, PA and Patey, KS (1996) Isotopic evidence for carbonate cementation and recrystallization, and for tectonic expulsion of fluids into the Western Canada Sedimentary Basin. Geological Society of America Bulletin 108, 1108–19.2.3.CO;2>CrossRefGoogle Scholar
Mangenot, X, Tarantola, A, Mullis, J, Girard, JP, Le, VH and Eiler, JM (2021) Geochemistry of clumped isotopologues of CH4 within fluid inclusions in Alpine tectonic quartz fissures. Earth and Planetary Science Letters 561, 116792. doi: 10.1016/j.epsl.2021.116792.CrossRefGoogle Scholar
McArthur, JM, Howarth, RJ and Bailey, TR (2001) Strontium isotope stratigraphy: LOWESS Version 3: best fit to the marine Sr-isotope curve for 0–509 Ma and accompanying look-up table for deriving numerical age. The Journal of Geology 109, 155–70.CrossRefGoogle Scholar
McCaig, AM, Wayne, DM, Marshall, JD, Banks, D and Henderson, I (1995) Isotopic and fluid inclusion studies of fluid movement along the Gavarnie thrust, central Pyrenees: reaction fronts in carbonate mylonites. American Journal of Science 295, 309–43.CrossRefGoogle Scholar
McLennan, SM, Taylor, SR, McCulloch, MT and Maynard, JB (1990) Geochemical and Nd–Sr isotopic composition of deep-sea turbidites: crustal evolution and plate tectonic associations. Geochimica et Cosmochimica Acta 54, 2015–50.CrossRefGoogle Scholar
Milliman, JD (1974) Marine Carbonates: Recent Sedimentary Carbonates. Part 1. New York: Springer, 375 pp.CrossRefGoogle Scholar
Milnes, AG and Pfiffner, OA (1977) Structural development of the Infrahelvetic complex, eastern Switzerland. Eclogae Geologicae Helvetiae 70, 8395.Google Scholar
Mittempergher, S, Cerchiari, A, Remitti, F and Festa, A (2017) From soft sediment deformation to fluid assisted faulting in the shallow part of a subduction megathrust analogue: the Sestola Vidicatico tectonic unit (Northern Apennines, Italy). Geological Magazine 155, 438–50.CrossRefGoogle Scholar
Mock, S, von Hagke, C, Schlunegger, F, Dunkl, I and Herwegh, M (2020) Long-wavelength late-Miocene thrusting in the north Alpine foreland: implications for late orogenic processes. Solid Earth 11, 1823–47.CrossRefGoogle Scholar
Moore, JC, Rowe, C and Meneghini, F (2007) How accretionary prisms elucidate seismogenesis in subduction zones. In The Seismogenic Zone of Subduction Thrust Faults (eds Dixon, TH and Moore, JC), pp. 288315. New York: Columbia University Press.CrossRefGoogle Scholar
Moore, JC and Saffer, DM (2001) Updip limit of the seismogenic zone beneath the accretionary prism southwest Japan: an effect of diagenetic to low-grade metamorphic processes and increasing effective stress. Geology 29, 183–6.2.0.CO;2>CrossRefGoogle Scholar
Mullis, J, Mählmann, RF and Wolf, M (2017) Fluid inclusion microthermometry to calibrate vitrinite reflectance (between 50 and 270°C), illite Kübler-Indes data and the diagenesis/anchizone boundary in the external part of the Central Alps. Applied Clay Science 143, 307–19.CrossRefGoogle Scholar
Nibourel, L, Berger, A, Egli, D, Heuberger, S and Herwegh, M (2021) Structural and thermal evolution of the eastern Aar Massif: insights from structural field work and Raman thermometry. Swiss Journal of Geosciences 114, 9. doi: 10.1186/s00015-020-00381-3.CrossRefGoogle ScholarPubMed
Nibourel, L, Berger, A, Egli, D, Luensdorf, NK and Herwegh, M (2018) Large vertical displacement of a crystalline massif recorded by Raman thermometry. Geology 46, 879–82.CrossRefGoogle Scholar
Nier, AO (1938) The isotopic constitution of strontium, barium, bismuth, thallium and mercury. Physical Reviews 54, 275–8.CrossRefGoogle Scholar
Pfiffner, OA (1986) Evolution of the north Alpine foreland basin in the central Alps. In Foreland Basins (eds Allen, PA and Homewood, P), pp. 219–28. Special Publications of the International Association of Sedimentologists no 8. Oxford: Blackwell Scientific.CrossRefGoogle Scholar
Pfiffner, OA (2011) Structural Map of the Helvetic Zone of the Swiss Alps, including Vorarlberg (Austria) and Haute Sovie (France), 1:100000. Geological Special Map 128. Explanatory Notes. Wabern: Federal Office of Topography Swisstopo.Google Scholar
Pytte, AM and Reynolds, RC (1989) The thermal transformation of smectite to illite. In Thermal History of Sedimentary Basins (eds Naeser, ND and McCulloh, TH), pp. 133–40. Berlin: Springer.CrossRefGoogle Scholar
Rahn, MK, Hurford, AJ and Frey, M (1997) Rotation and exhumation of a thrust plane: apatite fission-track data from the Glarus thrust, Switzerland. Geology 25, 599602.2.3.CO;2>CrossRefGoogle Scholar
Rahn, M, Mullis, J and Erdelbrock, K (1995) Alpine metamorphism in the North Helvetic Flysch of the Glarus Alps, Switzerland. Eclogae Geologicae Helvetiae 88, 157–78.Google Scholar
Remitti, F, Bettelli, G and Vannucchi, P (2007) Internal structure and tectonic evolution of an underthrust tectonic mélange: the Sestola-Vidiciatico tectonic unit of the Northern Apennines, Italy. Geodinamica Acta 20, 3751.CrossRefGoogle Scholar
Richter, FM and Liang, Y (1993) The rate and consequences of Sr diagenesis in deep-sea carbonates. Earth and Planetary Science Letters 117, 553–65.CrossRefGoogle Scholar
Sample, JC (2010) Stable isotope constraints on vein formation and fluid evolution along a recent thrust fault in the Cascadia accretionary wedge. Earth and Planetary Science Letters 293, 300–12.CrossRefGoogle Scholar
Sample, JC and Kopf, A (1995) Isotope geochemistry of syntectonic carbonate cements and veins from the Oregon margin: implications for the hydrogeologic evolution of the accretionary wedge. In Proceedings of the Ocean Drilling Program, Scientific Results, vol. 146 (eds Carson, B, Westbrook, GK, Musgrave, RJ and Suess, E), pp. 137–48. College Station, Texas.Google Scholar
Sample, JC, Torres, ME, Fisher, A, Hong, W-L, Destringeville, C, Defliese, WF and Tripati, AE (2017) Geochemical constraints on the temperature and timing of carbonate formation and lithification in the Nankai Trough, NanTroSEIZE transect. Geochimica et Cosmochimica Acta 198, 92114.CrossRefGoogle Scholar
Schmid, SM (1975) The Glarus overthrust: field evidence and mechanical model. Eclogae Geologicae Helvetiae 68, 247–80.Google Scholar
Schmid, SM, Pfiffner, OA, Froitzheim, N, Schönborn, G and Kissling, E (1996) Geophysical-geological transect and tectonic evolution of the Swiss-Italian Alps. Tectonics 15, 1036–64.CrossRefGoogle Scholar
Sharp, ZD and Kirschner, DL (1995) Quartz-calcite oxygen isotope thermometry: a calibration based on natural isotope variations. Geochimica et Cosmochimica Acta 58, 4491–501.CrossRefGoogle Scholar
Sibson, RH (1977) Fault rocks and fault mechanisms. Journal of the Geological Society, London 133, 191213.CrossRefGoogle Scholar
Sibson, RH (1998) Brittle failure mode plots for compressional and extensional tectonic regimes. Journal of Structural Geology 20, 655–60.CrossRefGoogle Scholar
Sibson, RH (2013) Stress switching in subduction forearcs: implications for overpressure containment and strength cycling on megathrusts. Tectonophysics 600, 142–52.CrossRefGoogle Scholar
Sinclair, HD (1997) Flysch to molasse transition in peripheral foreland basins: the role of the passive margin breakoff. Geology 25, 1123–6.2.3.CO;2>CrossRefGoogle Scholar
Sinclair, HD and Allen, PA (1992) Vertical versus horizontal motions in the Alpine orogenic wedge: stratigraphic response in the foreland basin. Basin Research 4, 215–32.CrossRefGoogle Scholar
Takeshita, T, Yamaguchi, A and Shigematsu, N (2014) Stress reversal recorded in calcite vein cuttings from the Nankai accretionary prism, southwest Japan. Earth, Planets and Space 66, 144. doi: 10.1186/s40623-014-0144-4.CrossRefGoogle Scholar
Tarantola, A, Mullis, J, Guillaume, D, Dubessy, J, de Capitani, C and Abdelmoula, M (2009) Oxidation of CH4 to CO2 and H2O by chloritization of detrital biotite at 270 ± 5 °C in the external part of the Central Alps, Switzerland. Lithos 112, 497510.CrossRefGoogle Scholar
Tarantola, A, Mullis, J, Vennemann, T, Dubessy, J and de Capitani, C (2007) Oxidation of methane at the CH4/H2O–(CO2) transition zone in the external part of the Central Alps, Switzerland: evidence from stable isotope investigations. Chemical Geology 237, 329–57.CrossRefGoogle Scholar
Travé, A, Labaume, P, Calvet, F and Soler, A (1997) Sediment dewatering and pore fluid migration inferred from isotopic and elemental geochemical analyses (Eocene southern Pyrenees, Spain). Tectonophysics 282, 375–98.CrossRefGoogle Scholar
Ujiie, K, Saishu, H, Fagereng, Å, Nishiyama, N, Otsubo, M, Masuyama, H and Kagi, H (2018) An explanation of episodic tremor and slow slip constrained by crack-seal veins and viscous shear in subduction mélange. Geophysical Research Letters 45, 5371–9.CrossRefGoogle Scholar
van de Kamp, PC (2008) Smectite-illite-muscovite transformations, quartz dissolution, and silica release in shales. Clays and Clay Minerals 56, 6681.CrossRefGoogle Scholar
Veizer, J, Ala, D, Azmy, K, Bruckschen, P, Buhl, D, Bruhn, F, Carden, GAF, Diener, A, Ebneth, S, Godderis, Y, Jasper, T, Korte, C, Pawellek, F, Podlaha, O and Strauss, H (1999) 87Sr/86Sr, δ13C and δ18O evolution of Phanerozoic seawater. Chemical Geology 161, 5988.CrossRefGoogle Scholar
Villa, IM (1998) Isotopic closure. Terra Nova 10, 42–7.CrossRefGoogle Scholar
Villa, IM (2016) Diffusion in mineral geochronometers: present and absent. Chemical Geology 420, 110.CrossRefGoogle Scholar
Voigt, J, Hathorne, ED, Frank, M, Vollstaedt, H and Eisenhauer, A (2015) Variability of carbonate diagenesis in equatorial Pacific sediments deduced from radiogenic and stable Sr isotopes. Geochimica et Cosmochimica Acta 148, 360–77.CrossRefGoogle Scholar
von Hagke, C, Cederbom, CE, Oncken, O, Stöckli, DF, Rahn, MK and Schlunegger, F (2012) Linking the northern Alps with their foreland: the latest exhumation history resolved by low-temperature thermochronology. Tectonics 31, TC5010. doi: 10.1029/2011TC003078.CrossRefGoogle Scholar
Vroliijk, P, Myers, G and Moore, JC (1988) Warm fluid migration along tectonic melanges in the Kodiak accretionary complex, Alaska. Journal of Geophysical Research 93, 10313–24.CrossRefGoogle Scholar
Wang, K (1994) Kinematic models of dewatering accretionary prisms. Journal of Geophysical Research 99, 4429–38.CrossRefGoogle Scholar
Figure 0

Fig. 1. (a) Geological map of the study area and (b) geographical overview. The line A–A′ indicates the trace of the cross-section shown in (c). (c) Synthetic and simplified cross-section. The approximate sampling sites are indicated together with peak metamorphic temperatures. Temperatures based on Ebert et al. (2007), Lahfid et al. (2010) and Rahn et al. (1995). Geological map in (a) and cross-section in (c) based on Pfiffner (2011).

Figure 1

Fig. 2. Examples of mineral veins sampled in the Globotruncana marl of the Ultrahelvetic flysch unit. (a, b) Bedding-parallel G1 calcite shear veins. G1 veins were folded together with bedding. (c) G2 quartz-calcite vein. The vein contains large clasts of brecciated host rock. (d, e) G3 quartz-calcite extension veins. G3 veins overprint G1 veins and the cleavage. (f) Mineralized tension gashes within fold hinges. The tension gashes record a brittle overprint of the folds. Sampling sites: 46.890° N, 9.153° E and 46.874° N, 9.126° E.

Figure 2

Fig. 3. (a) Tectonic contact between Ultrahelvetic flysch (hanging wall) and South-Helvetic flysch (footwall). (b) Example of calcite extension veins formed in the direct hanging wall of the imbricate thrust fault shown in (a). The veins contain fragments of a cataclasite that formed along the imbricate thrust. (c) Example of small thrusts cross-cutting calcite extension veins in the footwall of the imbricate thrust. (d) Example of steep calcite extension veins in the footwall. (e) Mineralized fissure with euhedral quartz and calcite crystals. Sampling site: 46.9597° N, 9.1881° E.

Figure 3

Fig. 4. South-Helvetic thrust slice exposed in northernmost part of the study area. (a) The marl is intensively sheared and comprises long shear veins. (b) Striations (c. 130|45) on shear surfaces indicate top-to-NW shearing (cf. Dielforder et al.2016). (c, d) Examples of boudins dissected by calcite extension veins. Sampling site: 47.1407° N, 9.1073° E.

Figure 4

Fig. 5. (a) Thrust fault in North-Helvetic flysch. The hanging wall of the thrust is intensively fractured. The mineral veins comprise quartz, calcite and minor amounts of chlorite and white mica. (b) Detail of mineral veins within the hanging wall of the thrust shown in (a). The veins overprint well-compacted and foliated rocks. (c) Detail of retrograde fissure overprinting the fault structure and related extension veins. Sampling site: 46.8877° N, 9.1273° E.

Figure 5

Fig. 6. Sr isotopic composition of vein carbonates (crosses) grouped for the different vein generations and sampling sites (a–d). Note the change in scale for 87Sr/86Sr values >0.71. See Figures 2–5 for examples of analysed mineral veins. The 2σ uncertainties on the 87Sr/86Sr ratios of vein carbonates are smaller than the symbols. Sr ratios of G1, G2 and G3 veins in (a) from Dielforder et al. (2015). Seawater values at the time of sediment deposition are shown for comparison; data from McArthur et al. (2001). Bulk host-rock values in (e) are recalculated to the time of metamorphism at 25 Ma; data compiled from Dielforder et al. (2015) and Hilgers & Sindern (2005).

Figure 6

Fig. 7. Diagram showing the positive correlation between the 87Sr/86Sr ratios of G2 and G3 veins from the Ultrahelvetic flysch unit and the formation temperature of these veins as constrained by oxygen isotope thermometry (data from Dielforder et al.2015). R is the correlation coefficient. The range of 87Sr/86Sr ratios obtained for nearby tension gashes overlaps with the values of G2 and G3 veins, which suggests an approximate formation temperature of 230–260 °C for the tension gashes as indicated by dashed lines. See Section 5.b.1 for details.

Figure 7

Fig. 8. Diagram illustrating the Sr isotopic evolution of pore fluids in the Ultrahelvetic flysch and South-Helvetic flysch units. The increase from the seawater-like signature towards more radiogenic values is interpreted to relate to diagenetic and metamorphic processes, including carbonate diagenesis, smectite-to-illite transformation, alkali-feldspar albitization, illite-to-muscovite transformation and recrystallization (recryst.) of detrital white mica and biotite. See Sections 5.a and 6 for details.

Figure 8

Fig. 9. Schematic cross-section illustrating the (a) diagenetic to (b) metamorphic evolution of the Infrahelvetic flysch units during the Alpine orogeny. LFM – Lower Freshwater Molasse; LMM – Lower Marine Molasse; NHF – North-Helvetic flysch; SHF – South-Helvetic flysch; UHF – Ultrahelvetic flysch. The blue arrows indicate fluid flow; the blue spirals indicate internal fluid evolution. Based on Burkhard et al. (1992), Dielforder et al. (2016) and Pfiffner (1986). Not to scale.

Supplementary material: File

Dielforder et al. supplementary material

Tables S1-S2

Download Dielforder et al. supplementary material(File)
File 31.1 KB