Hostname: page-component-7bb8b95d7b-fmk2r Total loading time: 0 Render date: 2024-09-28T01:18:59.997Z Has data issue: false hasContentIssue false

How the geochemistry of syn-kinematic calcite cement depicts past fluid flow and assists structural interpretations: a review of concepts and applications in orogenic forelands

Published online by Cambridge University Press:  15 February 2023

Nicolas E. Beaudoin*
Affiliation:
Universite de Pau et des Pays de l’Adour, E2S UPPA, LFCR, CNRS, Total Energies, Pau, France
Olivier Lacombe
Affiliation:
Sorbonne Université, CNRS-INSU, Institut des Sciences de la Terre de Paris – ISTeP, UMR7193, Paris, France
Guilhem Hoareau
Affiliation:
Universite de Pau et des Pays de l’Adour, E2S UPPA, LFCR, CNRS, Total Energies, Pau, France
Jean-Paul Callot
Affiliation:
Universite de Pau et des Pays de l’Adour, E2S UPPA, LFCR, CNRS, Total Energies, Pau, France
*
Author for correspondence: Nicolas E. Beaudoin, Email: nicolas.beaudoin@univ-pau.fr
Rights & Permissions [Opens in a new window]

Abstract

Orogenic forelands host interactions between deformation and static or migrating fluids. Given their accessibility and dimensions, these areas are not only historic landmarks for structural geology, but they are also areas of prime interest for georesource exploration and geological storage, and loci of potential geohazards. Geochemical techniques applied on cements filling tectonic structures and associated trapped fluids can constrain the temperature, pressure, origin and pathways of fluids during deformation and allow the characterization of the past fluid system. In this review focused on calcite cements, we first present and critically discuss the most used geochemical techniques to appraise specific parameters of the fluid system. Second, we summarize the outcomes of selected case studies where the past fluid system was reconstructed with consideration of tectonics, either at the scale of the individual fold/thrust or at the scale of the fold-and-thrust belt. At first order, the past fluid system evolves in a similar way with respect to the considered stage of deformation, being rather closed to external fluids when deformation is bounded to mesoscale structure development, and opening to vertical flow when thrust and folds develop. In a more detailed view, it seems that the past fluid system evolves and distributes under the influence of the structural style, of the geometry of the major faults and of the lithology of the sedimentary succession. Through this review, we illustrate the concept of geochemistry-assisted structural geology through case studies where the geochemistry of calcite veins constrained subsurface geometries and structural developments in orogenic forelands.

Type
FLUID FLOW AND MINERALIZATION IN FAULTS AND FRACTURES
Copyright
© The Author(s), 2023. Published by Cambridge University Press

1. Introduction

Orogenic forelands are regions where fluid migration intimately entwined with the development of deformational structures. In both fold-and-thrust belt – foreland basin systems (FTB hereinafter) or broken forelands (BF hereinafter), fluid migrations occur at the large scale (in relation with e.g. major faults) and at the mesoscale (in relation with e.g. mesoscale faults, joints, bands and stylolites (Figs 1, 2; Roure et al. Reference Roure, Swennen, Schneider, Faure, Ferket, Guilhaumou, Osadetz, Robion and Vandeginste2005; Groshong et al. Reference Groshong, Kronenberg, Couzens-Schultz and Newman2014; Lacombe et al. Reference Lacombe, Swennen and Caracausi2014; Agosta et al. Reference Agosta, Luetkemeyer, Lamarche, Crider and Lacombe2016; Lacombe & Rolland, Reference Lacombe and Rolland2016). The mesoscale structures directly affect the fluid transport by altering the porosity, either enhancing its connectivity with, for instance, fracturing, or reducing it with, for instance, cataclasis, consequently changing the permeability of a given rock volume. In turn, fluids interact with the rock to enhance the porosity locally by dissolution (e.g. Szymczak & Ladd, Reference Szymczak and Ladd2014) or to reduce it locally by either mineral precipitation (e.g. Bons et al. Reference Bons, Elburg and Gomez-Rivas2012) or mass transfer by pressure solution (e.g. Toussaint et al. Reference Toussaint, Aharonov, Koehn, Gratier, Ebner, Baud, Rolland and Renard2018). Studies of the interplays between fluid flow and deformation processes are numerous (e.g. Mourgues & Cobbold, Reference Mourgues and Cobbold2003; Laubach et al. Reference Laubach, Eichhubl, Hilgers and Lander2010, Reference Laubach, Lander, Criscenti, Anovitz, Urai, Pollyea, Hooker, Narr, Evans, Kerisit, Olson, Dewers, Fisher, Bodnar, Evans, Dove, Bonnell, Marder and Pyrak-Nolte2019; Cobbold et al. Reference Cobbold, Zanella, Rodrigues and Løseth2013; Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015). Besides the reconstruction of the diagenetic evolution of rocks, a large literature has been dedicated to the reconstruction of past fluid flow associated to deformation by studying syn-kinematic mineralization associated to the development of large-scale fault zones (e.g. Lacroix et al. Reference Lacroix, Trave, Buatier, Labaume, Vennemann and Dubois2014; de Graaf et al. Reference de Graaf, Nooitgedacht, Goff, van der Lubbe, Vonhof and Reijmer2019; Smeraglia et al. Reference Smeraglia, Fabbri, Choulet, Buatier, Boulvais, Bernasconi and Castorina2019), deformation bands (Parry et al. Reference Parry, Chan and Beitler2004; Fossen et al. Reference Fossen, Schultz, Shipton and Mair2007; Zuluaga et al. Reference Zuluaga, Rotevatn, Keilegavlen and Fossen2016; del Sole et al. Reference del Sole, Antonellini, Soliva, Ballas, Balsamo and Viola2020), and vein networks at the scale of the fold and at the scale of the entire FTB/BF (e.g. Ferket et al. Reference Ferket, Roure, Swennen and Ortuño2000; Roure et al. Reference Roure, Swennen, Schneider, Faure, Ferket, Guilhaumou, Osadetz, Robion and Vandeginste2005; Sibson, Reference Sibson2005; Fischer et al. Reference Fischer, Higuera-Díaz, Evans, Perry and Lefticariu2009; Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011, Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a, Reference Beaudoin, Huyghe, Bellahsen, Lacombe, Emmanuel, Mouthereau and Ouanhnon2015; Fitz-Diaz et al. Reference Fitz-Diaz, Hudleston, Siebenaller, Kirschner, Camprubí, Tolson and Puig2011 a; Evans & Fischer, Reference Evans and Fischer2012; Crognier et al. Reference Crognier, Hoareau, Aubourg, Dubois, Lacroix, Branellec, Callot and Vennemann2018; de Graaf et al. Reference de Graaf, Nooitgedacht, Goff, van der Lubbe, Vonhof and Reijmer2019; Dielforder et al. Reference Dielforder, Villa, Berger and Herwegh2022).

Fig. 1. Sketches representing a fold-and-thrust belt – foreland basin system (a) and a broken foreland (b). Stress regime, fluid system, source of fluids and engine of migrations, and mesoscale structures related to the folding event, from layer-parallel shortening to late stage of fold tightening, are reported.

Fig. 2. Field and microscope photographs of mesoscale structures studied for past fluid system reconstruction. (a, b) Veins in a fracture network in the Apennines, Italy (a) and in the Laramide province, USA (b). (c, d) Examples of vein textures: blocky calcite from the Apennines, Italy (c); elongated blocky calcite from southern Pyrenees, Spain (d). (e) Multiple fluid flow events and related cement precipitation in a crack seal, Laramide province, USA. (f, g) Cathodoluminescence images of a single phase of cementation in a vein whose borders were affected by dolomitization, Laramide province, USA (f), and of multiple diagenetic phases affecting a vein, Southeast Basin, France (g). Note pressure-solution (stylolite) along the vein border in (g). (h, i) Striated calcite steps along a fault surface, Laramide province USA (h), and fault breccia, associated to local overpressure, Jaca Basin, Spain (i).

This approach implies the reconstruction of the past fluid system (Evans & Fischer, Reference Evans and Fischer2012) from studying the syn-kinematic cement properties. The fluid system is characterized by (1) the nature of the fluid (i.e. formational, meteoric, basement-derived) from which the cement precipitated, (2) the temperature of precipitation (hotter or cooler than, or at equilibrium with, the surrounding rock), (3) the pressure regime at which the cement precipitated (under hydrostatic to sub-lithostatic), (4) the amount of interaction the fluid phase had with the host rock, and (5) the timing of the cement precipitation with consideration as to whether this age may also be the age of deformation, and (6) the hydraulic structure of the reservoir rock. A reservoir can be closed, i.e. with no lateral or vertical migration of fluids; it can be open and compartmentalized with stratified lateral migrations (i.e. no fluid external to the reservoir is involved), or with vertical migration limited by tectonic barriers like faults; finally it can be open and homogenized (i.e. with drains connecting it to other reservoirs or to the surface). One of the prime means of access to the past fluid system lies in the study of cements associated to the sequence of deformation that affected a given rock. Mineral precipitation in major fault zones, along mesoscale fault planes, in veins, and to a lesser extent associated to stylolite development, is the main and most studied marker of past fluids (Fig. 2). In this review, we focus on the mesoscale structures including fractures (faults, veins), bands and stylolite-related veins (e.g. tension gashes, reopened stylolites). For the sake of simplicity, we refer to all these mesoscale structures as the fracture network, as (1) all these develop following patterns (i.e. repetitive distribution) that arrange together as networks (Tavani et al. Reference Tavani, Storti, Lacombe, Corradetti, Muñoz and Mazzoli2015) and (2) they all are host to past fluid-derived mineralization related to deformation. The fracture network comprises sets of structures that can develop either locally (e.g. curvature at the fold hinge) or regionally (e.g. far field contraction). The cement that can fill a mesoscale structure can thus provide a good picture of the past fluid system at the time the structure developed, providing the cement is indeed coeval with the structure development, and that its chemistry remained unaltered. In mesostructures developed in FTB and BF, quartz and calcite are the most common minerals encountered. In this review, we will focus on syn-kinematic calcite cement, because it is the most frequently encountered in orogenic forelands as a majority of deformed rocks are carbonates in type. Calcite also yields more information about the past fluid system, even when considering only carbonate host rock. Calcite veins, striated calcite steps or calcite slickenfibres on mesoscale fault surfaces, and calcite veins around stylolites are ubiquitous in deformed sedimentary rocks in FTB/BF. Understanding the evolution of the past fluid system from calcite cements is bound to both the evolution of concepts and the continuous development of geochemical tools applied to calcite cements. Studying the syn-kinematic calcite cement geochemistry, some, if not all, of the parameters that constitute the fluid system can be understood for a given reservoir, providing a picture of the conduits used by the fluids during their migration, the so-called plumbing system. It also hints at the processes required to trigger and maintain the flow on a larger scale than the studied reservoir (Bjørlykke, Reference Bjørlykke and Farrell1994; Andresen, Reference Andresen2012). For instance, deciphering the past fluid system in FTB such as the Canadian Rocky Mountains has helped identify the main drains (faults/sedimentary interfaces) for hydrocarbon-rich fluids and the relative timing of the fluid migration along with the topographic-driven migration engine (Roure et al. Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010). This kind of information provides boundary conditions to basin-scale fluid flow simulations that predict the migration pathways and rates, and ultimately the distribution of plays (Roure et al. Reference Roure, Swennen, Schneider, Faure, Ferket, Guilhaumou, Osadetz, Robion and Vandeginste2005, Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010; Callot et al. Reference Callot, Sassi, Roure, Hill, Wilson and Divies2017).

This review aims at defining the concept of geochemistry-assisted structural geology, by focusing on the various case studies that report syn-kinematic calcite cement geochemical analysis in carbonate host rocks. We wish to illustrate what geochemical analysis can be conducted on syn-kinematic calcite cements to reconstruct the fluid temperature and pressure at the time of precipitation, the fluid origin, the age of precipitation, or the migration pathways. In turn, we question how this information can be used to qualitatively infer the distribution and activation time of the conduits, leading to deciphering structural information such as extension of faults, the nature of decoupling level, and the timescale of deformation. Hereinafter, we provide an overview of the geochemical techniques that can be used to characterize the conditions of precipitation of calcite cements (and to a lesser extent, quartz), and the chemistry of the fluid it precipitated from. Among others, we present classical techniques such as δ18O, δ13C, 87Sr/86Sr, elementary content, and fluid inclusion microthermometry, along with others under development such as Δ47CO2 and U–Pb geochronology. The prerequisite to a proper geochemical investigation of past fluid systems, i.e. a petrologic and diagenetic characterization of samples, will be briefly introduced, including vein petrography and techniques such as cathodoluminescence. We then present detailed examples of how the past fluid system can be reconstructed from calcite cements precipitated in major faults and/or in fracture networks at the scale of individual folds and at the scale of FTB (Canadian Rockies, Central Apennines, southern Pyrenees) or at the scale of the BF (Laramide province). More cases are then summarized, based on the large-scale structural style, in order to draw a complete picture of the relationship between the past fluid system, its evolution during deformation, and the structural geology of the subsurface.

Beyond the study of syn-kinematic cements in fault rocks from major faults, this review reports numerous studies focusing on the mesoscale structures (comprising faults and veins, but extended to stylolites and bands). Studying the fracture network in strata potentially allows access to the past fluids during a deformation history, from the weak deformation related to foreland flexure to the involvement of the strata in folds and/or thrusts as the FTB/BF developed (Fig. 1). This review excludes past fluid systems reconstructed from (1) magmatic fluids (e.g. MacDonald et al. Reference MacDonald, Faithfull, Roberts, Davies, Holdsworth, Newton, Williamson, Boyce and John2019) and related ore deposits (e.g. Jin et al. Reference Jin, Zhao, Feng, Hofstra, Deng, Zhao and Li2021), (2) hydrocarbon-dedicated geochemical techniques, and (3) geochemical techniques dedicated to non-carbonate veins, e.g. sulphur in gypsum cement (Machel, Reference Machel1985 b; Moragas et al. Reference Moragas, Martínez, Baqués, Playà, Travé, Alías and Cantarero2013). This review also does not examine how the understanding of a past fluid system can refine fluid flow simulation. As such, this contribution is to be seen as a complement to recent reviews dedicated to (1) how FTBs accommodate deformation through the development of mesostructures that damage reservoirs (Tavani et al. Reference Tavani, Storti, Lacombe, Corradetti, Muñoz and Mazzoli2015); (2) the specific relationship between fault zones and permeability (Faulkner et al. Reference Faulkner, Jackson, Lunn, Schlische, Shipton, Wibberley and Withjack2010) and more generally fault zone hydrogeology (Bense et al. Reference Bense, Gleeson, Loveless, Bour and Scibek2013); (3) the interplay between deformation and fluid flow at the fold scale (Evans & Fischer, Reference Evans and Fischer2012); (4) the prediction of past fluid flow in FTBs (Roure et al. Reference Roure, Swennen, Schneider, Faure, Ferket, Guilhaumou, Osadetz, Robion and Vandeginste2005, Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010); (5) the petrophysical evolution of carbonates (Archie, Reference Archie1952; Regnet et al. Reference Regnet, David, Robion and Menéndez2019); and (6) the interplay between chemistry and deformation in carbonates (Laubach et al. Reference Laubach, Lander, Criscenti, Anovitz, Urai, Pollyea, Hooker, Narr, Evans, Kerisit, Olson, Dewers, Fisher, Bodnar, Evans, Dove, Bonnell, Marder and Pyrak-Nolte2019).

2. Fault zones and fracture networks: witnesses of past fluid flow

2.a. Development of fault zones and fracture networks

Orogenic forelands, that can be either fold-and-thrust belt – foreland basin systems or broken forelands, are the external parts of orogens where the deformation of sedimentary rocks occurred under diagenetic to low-grade metamorphic conditions of pressure and temperature, and in that sense are opposed to the higher-grade metamorphic hinterland (Fig. 1). The depth and degree of the mechanical decoupling level within the continental crust activated during collision has been widely debated for many years. One view states that the sedimentary cover is detached from the underlying basement along a shallow, low-strength décollement and deformed by thrusts with ramp-flat geometries rooting into the décollement level (thin-skinned tectonic style). This thin-skinned tectonic style supposes large-scale displacements and duplication of the sedimentary sequence, the underlying basement remaining undeformed (Lacombe & Bellahsen, Reference Lacombe and Bellahsen2016; Pfiffner, Reference Pfiffner2017). The low-strength décollement where thrusts are rooting at depth can be either viscous (e.g. salt, as in the Potwar Basin, Pakistan) or frictional (e.g. shales, as in the Alberta Rocky Mountains, Canada), or it can be associated with salt-related tectonic structures (e.g. the Sivas Basin, Turkey; Hudec & Jackson, Reference Hudec and Jackson2007; Callot et al. Reference Callot, Trocmé, Letouzey, Albouy, Jahani and Sherkati2012). The alternative view states that shortening involves a significant part of the crust along crustal-scale ramps above a deep ductile detachment (thick-skinned tectonic style). These are two end-members, and superimposed thin-skinned and basement-involved deformation may occur in various parts of the belt (Lacombe & Bellahsen, Reference Lacombe and Bellahsen2016; Pfiffner, Reference Pfiffner2017). Broken forelands (Fig. 1b) are characterized by basement uplifts, forming arches in a possibly erratic sequence of basement fault reactivation. These uplifts segment the former basin (Jordan & Allmendinger, Reference Jordan and Allmendinger1986; Horton et al. Reference Horton, Capaldi, Mackaman-Lofland, Perez, Bush, Fuentes and Constenius2022). Examples of broken forelands are the Laramide province in the western USA and the Sierras Pampeanas in Argentina.

The porosity and permeability of carbonate rocks are modified very early by diagenesis and later by structural damage. The latter develops mesoscale structures: (1) surfaces/bands that create porosity/permeability by fracturing, with a displacement either perpendicular to their surfaces (Fig. 2a–g; joints, veins and dilation bands) or along/oblique to their surfaces (Fig. 2h, i; faults and shear bands); and (2) surfaces/bands that reduce porosity/permeability by pressure-solution/cataclasis, with a motion either perpendicular to their surfaces (stylolites and compaction bands) or along/oblique to their surfaces (slickolites/shear bands). The distribution of these mesoscale structures varies spatially and temporally according to the position of the considered strata in the FTB/BF (Fig. 1). Considering carbonates, the formation of the fracture network starts during burial with the development of along-strike and across-strike joints and veins and of bedding-parallel sedimentary stylolites, the latter potentially developing at very shallow depths (<400 m; Toussaint et al. Reference Toussaint, Aharonov, Koehn, Gratier, Ebner, Baud, Rolland and Renard2018). During foreland flexure, along-strike and across-strike joints and veins oriented at high angle to bedding develop, along with normal faults, in the sedimentary series. These structures form in response to forebulge development and along-foredeep stretching (in blue in Fig. 1a). The folding event (Lacombe et al. Reference Lacombe, Beaudoin, Hoareau, Labeur, Pecheyran and Callot2021) starts with the layer-parallel shortening (LPS). LPS produces a network of structures comprising (1) joints and veins either at high angle to bedding and striking parallel to the shortening direction or parallel to bedding, (2) pressure solution cleavages, including tectonic stylolites, of which planes are oriented at high angle to bedding and perpendicular to the shortening direction, and (3) conjugate strike-slip or reverse faults. The LPS may still affect strata at the onset of fold growth (Fig. 3) up to c. 30° of bed tilting (Tavani et al. Reference Tavani, Storti, Lacombe, Corradetti, Muñoz and Mazzoli2015), and it can also develop bedding-parallel veins. In the orogenic foreland, fold growth is accommodated by flexural slip in the fold limbs and tangential longitudinal strain (e.g. outer-arc extension) at the fold hinge, the latter giving birth to along-strike joints at high angle to bedding as well as normal faults (Figs 1, 3). The fold ‘locks’ when limb rotation and/or kink-band migration cannot accommodate shortening anymore. At that stage, strata tilting is over but continuous horizontal shortening leads to late-stage fold tightening (LSFT), accommodated by mesoscale structures developing irrespective of bedding dip, such as (1) vertical conjugate strike-slip faults, (2) vertical joints striking parallel to the LPS-related ones, and (3) vertical tectonic stylolites oriented perpendicular to the shortening direction. At all stages of fold growth, earlier-formed fractures, either inherited from an ancient contractional event or from the foreland flexure preceding folding and thrusting, may also be reactivated (Guiton et al. Reference Guiton, Leroy and Sassi2003; Bergbauer & Pollard, Reference Bergbauer and Pollard2004; Bellahsen et al. Reference Bellahsen, Fiore and Pollard2006 b; Callot et al. Reference Callot, Robion, Sassi, Guiton, Faure, Daniel, Mengus and Schmitz2010 a; Sassi et al. Reference Sassi, Guiton, Leroy, Daniel and Callot2012). The folding event, encompassing deformation from the LPS to the LSFT, spans from a few Myr to a few tenths of Myr, according to the structural style and to the distance between the front of the range and the location of the (future) fold (Lacombe et al. Reference Lacombe, Beaudoin, Hoareau, Labeur, Pecheyran and Callot2021). It is important to note that in some specific cases this sequence will see development of fractures geometrically identical at various times of the deformation history, such as outer-arc extension related to the forebulge and to fold growth, or fold-axis perpendicular, LSFT and LPS fractures in fold limbs. In such cases, the origin of the fracture can only be assessed by robust relative chronology observations and/or might be assessed by geochemical, temperature and absolute age constraints on the filling cements precipitated during deformation.

Fig. 3. (a) Expected distribution of the mesoscale fracture pattern at the fold scale, including faults, joints and stylolites. Colours show the relation between the feature and the stage of deformation: brown relates to burial, grey to pre-orogenic, pink to outer arc extension during flexure, black to the layer-parallel shortening, green to fold growth, red to late stage of fold tightening. (b–e) Zoom on the expected fluid flow in the structures depicted in (a): (b) matrix-scale reactive fluid migration pathways; (c) migration pathways in joint network; (d) migration pathways in stylolites; (e) migration pathways in fault zones and cores, with the static state and the dynamic state. See text for details and references.

2.b. Fluid sources and fluid evolution during migration

FTB/BF develop with various structural styles, that affect (1) the availability of different fluid sources and (2) the triggers for the fluid migration (Fig. 1). Excluding magmatic fluids we can distinguish between various parent fluids at the time of precipitation: (1) surficial fluids such as meteoric fluids and formational water (seawater or fresh water); (2) crustal fluids, either described as basement fluids or as metamorphic fluids. The metamorphic fluids are either produced during retrograde metasomatic reactions (Yardley & Graham, Reference Yardley and Graham2002) or during prograde dehydration reactions (e.g. Hollocher, Reference Hollocher1991) occurring deeper in the crust. The fluids flowing in the basement rocks, beyond magmatic fluids, are proven to be usually derived from surficial fluids (Yardley & Graham, Reference Yardley and Graham2002). Similarly, formational water can evolve during fluid migration by interacting with the host rock in the basin, making so-called brines (or basinal fluids) if interacting with host rocks.

The fluid migration is triggered by various mechanisms (Fig. 1). The most common is the water table migration, or topographically driven flow, based on topographic/burial difference of a given stratum, that triggers pressure gradients making fluids migrate from the deepest towards the shallowest/highest part of the strata (Bjørlykke, Reference Bjørlykke and Farrell1994, Reference Bjørlykke2015), usually along the slope between the hinterland and the forebulge, or between synclines and anticlines. Migration can also be triggered by the stress gradient of units thrusting over others, expelling the fluid forelandwards. This mechanism, called squeegee, channelizes fluids along permeable horizons (Oliver, Reference Oliver1986) where fluid can migrate at a speed of 10 km Myr−1 over several million years (Ge & Garven, Reference Ge and Garven1994), with pulses up to 100 km Myr−1 (Michael & Bachu, Reference Michael and Bachu2001). Local stress accumulation in fault zones may also trigger ‘hot flashes’, an along-plane fast upward migration of hot fluids (Fig. 1; Machel & Cavell, Reference Machel and Cavell1999). Fluids can also flow downwards in fault zones following pressure gradients between the sealed and the broken domains during the seismic cycle (fault valve behaviour: Henderson & McCaig, Reference Henderson and McCaig1996; Sibson, Reference Sibson2000), or by enhancing permeability in the fault zone by local overpressure (Ortiz et al. Reference Ortiz, Person, Mozley, Evans and Bilek2019). Local chemical gradients between fluids and rocks can also enhance fluid migration, typically by carbonate dissolution, developing wormholes due to the flow of meteoric water undersaturated with respect to carbonates (Szymczak & Ladd, Reference Szymczak and Ladd2009; Petrus & Szymczak, Reference Petrus and Szymczak2016). This mechanism is efficient at a very shallow crustal level (Fig. 1).

2.c. Fault zones and fluid migration

At the scale of fault zones, the link between tectonics and fluid dynamics has been extensively investigated. For instance, the hydrological behaviour of fault zones, first introduced by Caine et al. (Reference Caine, Evans and Forster1996), has received a lot of attention since, with a compelling review proposed by Faulkner et al. (Reference Faulkner, Jackson, Lunn, Schlische, Shipton, Wibberley and Withjack2010). When displacement exceeds 1 m (Micarelli et al. Reference Micarelli, Benedicto and Wibberley2006), a fault can be divided into two main zones characterized by very different hydraulic behaviours (Fig. 3): the fault core, usually clay-rich, is a non-permeable zone, acting as a transverse barrier to the fluid which is forced to flow along the fault plane. In the surrounding damage zone, fractures develop (Fig. 2i), progressively increasing the connected porosity (Bense et al. Reference Bense, Gleeson, Loveless, Bour and Scibek2013). However, fast cement precipitation after hydrofracturing can potentially reduce the permeability, leading to an overpressure–failure cycle bound to the seismic cycle (Sibson, Reference Sibson and Knipe1990; Vass et al. Reference Vass, Koehn, Toussaint, Ghani and Piazolo2014). Regardless of the stress regime under which the fault developed, the damage zone is observed as being the main drain for the fluid, being especially efficient for a flow parallel to the fault plane (Agosta, Reference Agosta2008; Baietto et al. Reference Bau and Möller2008; Sutherland et al. Reference Sutherland, Toy, Townend, Cox, Eccles, Faulkner, Prior, Norris, Mariani, Boulton, Carpenter, Menzies, Little, Hasting, De Pascale, Langridge, Scott, Reid Lindroos, Fleming and Kopf2012).

2.d. Fracture network and fluid migration

At the scale of the mesostructures, and leaving aside deformation bands that are rare in carbonates (Tondi et al. Reference Tondi, Antonellini, Aydin, Marchegiani and Cello2006), two kinds of mesostructures exert a strong control on the fluid migration (Fig. 3c, d): the fractures, either opened in mode I (referred to as joints: Oliver & Bons, Reference Oliver and Bons2001; Sachau et al. Reference Sachau, Bons and Gomez-Rivas2015; Wennberg et al. Reference Wennberg, Casini, Jonoud and Peacock2016) or shear/hybrid in type (mode II or I + II), and the stylolites (Braithwaite, Reference Braithwaite1989). Regardless of the nature of the rock volume, the hydraulic efficiency of fractures is directly controlled by (1) the fracture aperture, (2) the orientation of the fracture plane with respect to the ambient stress field, (3) its vertical/lateral extent and (4) the density and connectivity of the fracture network. In carbonates, fracture density and permeability are linked to the diagenetic state of the reservoir and to the mechanical properties of the rock, namely stiffness and tensile strength. There are reciprocal controls between the diagenetic state of the rock and the fracture development, leading to the concept of mechanical stratigraphy (Ortega et al. Reference Ortega, Gale and Marrett2010; Barbier et al. Reference Barbier, Leprêtre, Callot, Gasparrini, Daniel, Hamon, Lacombe and Floquet2012 a; Wennberg et al. Reference Wennberg, Casini, Jonoud and Peacock2016). An efficient vertical fluid flow is observed in various folds (Evans & Fischer, Reference Evans and Fischer2012), probably related to an enhanced vertical extension of the joint network during folding. In contrast, joints developed during the LPS allowed efficient stratified, lateral migration of fluids with no vertical mixing in various case studies (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014a, Reference Beaudoin, Labeur, Lacombe, Koehn, Billi, Hoareau, Boyce, John, Marchegiano, Roberts, Millar, Claverie, Pecheyran and Callot2020 c). These observations of different hydraulic behaviour for the same kind of joints led to the proposition that this difference is not due just to the different geometry, but probably also to the stress regime active, influencing fracture aperture (and its evolution during time) (Fig. 3). Indeed, joint networks developed during LPS under a strike-slip stress regime favour lateral fluid flow and fluid system stratification while joint networks developed in response to strata curvature under a local extensional stress regime favour vertical fluid flow and fluid homogenization between reservoirs. Although less efficient for fluid flow than fault zones, the mesoscale structures are also an important part of the palaeohydrology at the scale of the FTB (Beaudoin et al. Reference Beaudoin, Lacombe, Bellahsen and Emmanuel2013), and the veins can reliably be used to reconstruct the fluid system related to both large-scale and local tectonics.

The role of stylolites on the fluid system remains vastly overlooked in that context, except to constrain the diagenesis related to the burial phase or the early LPS (Swennen et al. Reference Swennen, Muskha and Roure2000; Roure et al. Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010; Vandeginste et al. Reference Vandeginste, Swennen, Allaeys, Ellam, Osadetz and Roure2012). Stylolites are mainly considered as efficient barriers for fluid flow, since they accumulate non-soluble, non-permeable material along their dissolution plane, and redistribute cement around, clogging the porosity (Nelson, Reference Nelson1981). This consideration makes stylolites interpreted as structures able to compartmentalize a reservoir (Roure et al. Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010). Yet, a growing number of natural observations and experimental case studies reveal that stylolites can be efficient drains for fluid flow (Koehn et al. Reference Koehn, Rood, Beaudoin, Chung, Bons and Gomez-Rivas2016; Heap et al. Reference Heap, Reuschlé, Baud, Renard and Iezzi2018; Martín-Martín et al. Reference Martín-Martín, Gomez-Rivas, Gómez-Gras, Travé, Ameneiro, Koehn and Bons2018; Toussaint et al. Reference Toussaint, Aharonov, Koehn, Gratier, Ebner, Baud, Rolland and Renard2018; Bruna et al. Reference Bruna, Lavenu, Matonti and Bertotti2019). The impact of stylolites on the fluid flow depends on the orientation of the stylolite with respect to the flow, i.e. guiding a fluid flow parallel to the dissolution plane, and blocking the flow perpendicular to the dissolution plane (Fig. 3). This behaviour, similar to that of a fault zone, was observed in nature with a channelization of the reactive fluid flow along a stylolite plane (Martín-Martín et al. Reference Martín-Martín, Gomez-Rivas, Gómez-Gras, Travé, Ameneiro, Koehn and Bons2018; Humphrey et al. Reference Humphrey, Gomez-Rivas, Koehn, Bons, Neilson, Martín-Martín and Schoenherr2019) and in experiments (Heap et al. Reference Heap, Reuschlé, Baud, Renard and Iezzi2018). Yet it is not the only controlling factor of permeability to fluid flow. Indeed, Koehn et al. (Reference Koehn, Rood, Beaudoin, Chung, Bons and Gomez-Rivas2016) have observed ore-bearing stylolites in the Zechstein carbonate units (Germany), which demonstrates a strong control of the stylolite morphology over the fluid flow. They proposed a new classification in which the morphology of the stylolite, i.e. rectangular, suture and sharp peak, seismogram and simple wave-like, itself controlled by the growth rate of the stylolite, controls its impact over the fluid flow. In the classification, the stylolites that have high-amplitude peaks undergo a growth that locally shifts the non-permeable residue layer, allowing a fluid flow across the stylolite plane. It is important to note that because of local stress perturbations (Aharonov & Karcz, Reference Aharonov and Karcz2019), cracks can open during pressure solution and be filled with cement that will record information in relation to the stylolite development. Considering the complexity of stylolite development and its potential impact on fluid flow, it appears necessary to further investigate stylolites and reappraise their role in past fluid systems.

Though beyond the scope of this review, it is worth mentioning that the fluid chemistry may itself impact fluid flow, for instance through dissolution or fluid-mediated replacement, that will affect the fluid dynamics down to the crystal scale. Classic examples of the latter in carbonates are dolomitization or apatitization processes, and recent experimental works (Jonas et al. Reference Jonas, John, King, Geisler and Putnis2014; Pedrosa et al. Reference Pedrosa, Boeck, Putnis and Putnis2017; Weber et al. Reference Weber, Cheshire, Bleuel, Mildner, Chang, Ievlev, Littrell, Ilavsky, Stack and Anovitz2021) and natural observations (Centrella et al. Reference Centrella, Beaudoin, Derluyn, Motte, Hoareau, Lanari, Piccoli, Pecheyran and Callot2021) highlighted the prominent role of grain boundaries and mineralogical defects like cleavage planes on the fluid flow at that scale (Fig. 3).

2.e. Past fluid flow markers

Reconstructing fluid flow in orogenic forelands requires identification of markers of past fluid flow. These witnesses can be indirect, like the contrasted colour halo in the host rock around a fracture related to the migration of a fluid with a different redox state than the host (e.g. Eichhubl et al. Reference Eichhubl, Taylor, Pollard and Aydin2004; Stel, Reference Stel2009; Missenard et al. Reference Missenard, Bertrand, Vergely, Benedicto, Cushing and Rocher2014). In such cases, quantitative information remains very limited. More interestingly, there is a wealth of direct markers of past fluid flow, like cements precipitating in breccia during hydrofracturing (Fig. 2i), even though the link between breccia texture and hydrofracturing is nowadays questioned (Centrella et al. Reference Centrella, Beaudoin, Koehn, Motte, Hoareau and Callot2022). Mineralogical phase transformation such as dolomitization is another example of a well-studied marker that allows reconstruction of the past fluid system (e.g. Martín-Martín et al. Reference Martín-Martín, Gomez-Rivas, Bover-Arnal, Travé, Salas, Moreno-Bedmar, Tomás, Corbella, Teixell, Vergés and Stafford2013; Koeshidayatullah et al. Reference Koeshidayatullah, Corlett, Stacey, Swart, Boyce, Robertson, Whitaker and Hollis2020 a, b; Centrella et al. Reference Centrella, Beaudoin, Derluyn, Motte, Hoareau, Lanari, Piccoli, Pecheyran and Callot2021; Motte et al. Reference Motte, Hoareau, Callot, Revillon, Piccoli, Calassou and Gaucher2021). Yet, the most studied witness of the past fluid flow in contractional settings consists in the cements filling fractures (veins and faults), including the fluid inclusions they host (Bons et al. Reference Bons, Elburg and Gomez-Rivas2012; Laubach et al. Reference Laubach, Lander, Criscenti, Anovitz, Urai, Pollyea, Hooker, Narr, Evans, Kerisit, Olson, Dewers, Fisher, Bodnar, Evans, Dove, Bonnell, Marder and Pyrak-Nolte2019). In FTB and BF, the mineralogy of such filling, beyond calcite and quartz, can include fluorite, ores or gypsum.

3. Bounding cement precipitation from past fluid to deformation: the role of petrography

In order to reconstruct the past fluid properties during deformation, it is important to (1) unambiguously link cement precipitation to the timing of development of the studied structure (fault, vein, stylolite), and (2) check whether the cement has been altered or not by later diagenetic events. This is a prerequisite to any robust geochemical analysis of the past fluid system. In the case of cements associated to faults, the observation of either cement in cataclasites (Fig. 2i) or, better, of striated calcite steps (Fig. 2h) is a reliable way to ensure that the cement precipitated from a fluid involved during the deformation, provided the breccia cement is not a simple matter of recrystallization (Centrella et al. Reference Centrella, Beaudoin, Koehn, Motte, Hoareau and Callot2022). In the case of veins, a robust petrographic study is systematically required prior to any geochemical analysis being performed. First, the nature of the filling is an obvious yet ambiguous indicator of the nature of the fluid. For example, the occurrence of gypsum veins indicates the involvement of fluids that were in contact with evaporitic layers in the reservoir (Pichat et al. Reference Pichat, Hoareau, Callot, Legeay, Kavak, Révillon, Parat and Ringenbach2018), but it cannot argue in favour of the evaporites being the source of the fluids (dehydration during the tectonic remobilization of a salt layer) or part of the migration pathway of the fluid (a meteoric fluid percolating within a gypsum level; e.g. Travé et al. Reference Travé, Labaume and Vergés2007). Another example is the content of the crystal fluid inclusions (halite crystals, hydrocarbons) that can also be informative about either the fluid origin or its pathways. In their review, Bons et al. (Reference Bons, Elburg and Gomez-Rivas2012) illustrate what vein cement texture can tell us about the mode of deformation, and about the common timing between vein opening and cement precipitation (Fig. 2c–e). The coevality between precipitation of cement and deformation is obvious when the cement precipitates as fibrous crystals in opened veins (Bons et al. Reference Bons, Elburg and Gomez-Rivas2012), or when the vein growth is driven by crystallization forces during precipitation (Gratier et al. Reference Gratier, Frery, Deschamps, Røyne, Renard, Dysthe, Ellouz-Zimmerman and Hamelin2012; Cobbold et al. Reference Cobbold, Zanella, Rodrigues and Løseth2013). In the case of a syntaxial texture supporting the growth-induced opening of a vein (Fig. 2d), experimental data suggest that a very high fluid flow is required to keep the crystal growth rate sufficiently high (Lee & Morse, Reference Lee and Morse1999; Hilgers & Urai, Reference Hilgers and Urai2002 b). In the case where the timing of precipitation is ambiguous with respect to the development of the fracture (e.g. in presence of blocky calcite; Fig. 2c), the process that triggered precipitation must be questioned. In the case of calcite, a mineral with a retrograde solubility (Segnit et al. Reference Segnit, Holland and Biscardi1962), oversaturation for a given fluid is reached when temperature increases or when pCO2 decreases. Precipitation can be triggered by mixing between hydrothermal migrating fluids and local fluids. In cases where fluid mixing is not supported by data, authors have considered that the opening of the fracture triggers the precipitation by locally reducing pCO2 (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a; Cruset et al. Reference Cruset, Cantarero, Vergés, John, Muñoz-López and Travé2018). It is interesting to note that the recent study of Roberts and Holdsworth (Reference Roberts and Holdsworth2022) suggests that veins developed as crack seal are optimal to access the past fluid at the time of deformation, even though this needs to be considered with care as Bons et al. (Reference Bons, Elburg and Gomez-Rivas2012) clearly showed that a crack seal can be related to multiple events of deformation.

When a cement is considered to be linked to the fracture development, a second mandatory petrographic step is to check that the cement was not altered by diagenesis. Indeed, calcite crystals can keep their initial chemical signature pristine at low-grade metamorphism (e.g. Dielforder et al. Reference Dielforder, Villa, Berger and Herwegh2022), so it is important to discard the idea that dissolution/reprecipitation or recrystallization processes affected the cement. A classical tool used for that purpose is cathodoluminescence (Fig. 2f, g) ( Machel, Reference Machel1985 a; Reference Machel1997). The natural luminescence of calcite (but also of quartz, fluorite, gypsum) in a vein is informative about (1) the number of cement phases in the fracture, and/or alteration events affecting the vein cements, (2) the redox state of the fluid, itself arguably relatable to the fluid origin (Machel, Reference Machel1997), and (3) the rate of precipitation of calcite cement, variation of which can create luminescence zoning (Dromgoole & Walter, Reference Dromgoole and Walter1990; Boggs & Krinsley, Reference Boggs and Krinsley2006). For the latter, the relative part of the growth rate with regard to the surface growth preference has been debated (Reeder & Grams, Reference Reeder and Grams1987). Above all, ensuring that the cement filling the vein precipitated in a single phase (homogeneous luminescence; Fig. 2f), and that it has not been subsequently altered (Fig. 2g) is key to validate that the geochemical analysis of the cement actually yields information about the past fluid (e.g. van Geet et al. Reference van Geet, Swennen, Durmishi, Roure and Muchez2002; Hanks et al. Reference Hanks, Parris and Wallace2006; Barbier et al., Reference Barbier, Leprêtre, Callot, Gasparrini, Daniel, Hamon, Lacombe and Floquet2012 a). That preliminary step is mandatory to allow further interpretation of a geochemical dataset with respect to the tectonic history of an area.

4. Geochemical proxies to reconstruct the past fluid system

In this section we review the basic concepts and pitfalls of the geochemical techniques that are used for, but not limited to, the reconstruction of part of the past fluid system when applied to calcite cement in tectonic structures. We first focus on the most classical techniques, then summarize the less developed/applied ones. Figure 4 presents an ideal workflow of how these techniques can be applied to a sample, with emphasis on what (combination of) analysis can be used to access specific parameters of the fluid system. For reviews of the use of geochemistry for past fluid beyond its relation to tectonics (e.g. diagenesis), we refer the reader to the existing literature (Emery & Robinson, Reference Emery and Robinson1993; Swart, Reference Swart2015).

Fig. 4. Concept of geochemistry-assisted structural geology. Left-hand side: representation of all parameters (bold) that can be reconstructed from geochemical analysis (framed) on calcite, organized in a suggested workflow. The suggestion is to pick at least one of the analyses of each box to reconstruct the past fluid system with the least ambiguity. Right-hand side: structural implication of the outcome of the past fluid system reconstruction, including a direct appraisal of the reservoir hydrological structure, and an inferred model of the fluid conduit distribution in time and space.

4.a. Fluid inclusions

Fluid inclusions are micrometre- to millimetre-scale inclusions – filled with fluids, or a combination of fluid, vapour and solid – that are ubiquitous in crystals (Goldstein & Reynolds, Reference Goldstein and Reynolds1994). Three types of fluid inclusions are classically described (Roedder, Reference Roedder1984; Goldstein & Reynolds, Reference Goldstein and Reynolds1994): (1) the primary fluid inclusions, that are defects related to the crystal growth itself. Primary fluid inclusions mainly occur along the crystal growth rims, and mostly have a shape corresponding to the crystal habitus. (2) Pseudo-secondary fluid inclusions are not directly distributed in a pattern linked to the crystal growth, yet they are lined onto deformation planes that developed during the crystal growth, being limited to the crystal. (3) Secondary fluid inclusions appear as trails, the orientation of which is not related to the crystal growth; they can be considered as post-growth mode I joints at the crystal scale (e.g. André et al. Reference André, Sausse and Lespinasse2001). Regardless of the type of inclusions, the mineralizing fluid is trapped at set pressure and temperature. Assuming that (1) the volume of the inclusion remains constant, (2) fluid–solid chemical exchanges are null and thus (3) the fluid chemistry is constant, the pressure and temperature decrease during the subsequent exhumation of the crystal triggers phase changes potentially leading to the development of a vapour bubble (e.g. Bourdet et al. Reference Bourdet, Pironon, Levresse and Tritlla2008). Fluid inclusions might be one-phase, two-phase with vapour/liquid or liquid/solid, or three-phase with coexistence in the laboratory conditions of solid, vapour and liquid phases, the former being typical of highly saline fluids (Goldstein & Reynolds, Reference Goldstein and Reynolds1994). In the following paragraphs, we do not mention techniques that rely on hydrocarbon-bearing fluid inclusions. For more information than is summarized hereinafter, please refer to the detailed technical review by Chi et al. (Reference Chi, Diamond, Lu, Lai and Chu2020).

The original analysis conducted with fluid inclusions consists in reconstructing the temperature and pressure condition of fluid entrapment, i.e. at the time the crystal precipitated, by means of microthermometry. A thick section of mineral is heated up under the microscope using a temperature-controlled microthermometric stage in order to measure the temperature at which the vapour phase becomes homogeneous with the liquid one. This is called the temperature of homogenization (T h), and corresponds to the minimal temperature of the fluids when it was trapped in the inclusion (Goldstein & Reynolds, Reference Goldstein and Reynolds1994). Note that if the fluid temperature is lower than 50 °C at the time of entrapment, the vapour bubble will usually be absent or may not be visible in laboratory conditions, but it can be nucleated or enlarged by freezing the fluid. With the knowledge of, or by assuming, the fluid chemical system and the pressure of entrapment (i.e. depth), one can reconstruct the absolute temperature of precipitation. By comparing it to the expected temperature at which the host rock was at the time of precipitation, by considering a geothermal gradient and a past depth, it is then possible to assess if the fluid is hydrothermal (hotter than the local temperature), geothermal (equal to the local temperature) or hydrofrigid (colder than the local temperature) sensu Machel and Lonnee (Reference Machel and Lonnee2002). Microthermometry is also used to estimate the type of salts and the salinity of the system. The protocol is to freeze the studied fluid inclusion, and to heat it back up slowly to measure (1) the temperature(s) at which the ice starts to melt (i.e. eutectic temperature(s) T e), which relates to the fluid composition (Davis et al. Reference Davis, Lowenstein and Spencer1990), and (2) the temperature at which the last cube of ice disappears, i.e. the ice melting temperature (Tm ice). In a simple H2O–NaCl system, the temperature at which a frozen fluid inclusion becomes completely liquid is directly related to the fluid salinity.

Two major pitfalls affect fluid inclusion microthermometry in calcite cement. The first is that the approach is valid only under the assumption that the volume and chemistry of the inclusion remained constant. In weak crystals such as calcite (Goldstein, Reference Goldstein1986; Bodnar, Reference Bodnar, Samson, Anderson and Marshall2003), volume might easily vary (1) during the analysis and (2) throughout the crystal history, especially in deformed areas. The variation in volume of the inclusion will usually have a petrographic effect on the shape of the inclusion (e.g. stretching), or a visible effect on the vapour bubble membrane, making it impossible to consider the measured T h in the dataset of measurements. In order to avoid misleading temperature data, the norm is to conduct a rigorous petrographic study to distinguish between the different generations of fluid inclusions in a given cement, then first to measure the homogenization temperature of fluid inclusion populations before the eutectic and ice melting temperatures (as freezing likely will stretch the inclusions), and then to use the modal value of population distribution as representative of the T h. When data is available, the comparison of the distribution with current geothermal temperature might also indicate a thermal re-equilibration of the fluid inclusion population with the current temperature, involving volume variations within the studied fluid inclusions (Bodnar, Reference Bodnar, Samson, Anderson and Marshall2003). The second pitfall is that, considering the usual size of inclusions in calcite (<10 µm), the eutectic and melting temperatures are usually very hard to constrain because of technical limitation related to their limited size.

Fluid inclusions can also grant direct access to the fluid composition. One method consists in evaporating the liquid phase following sample crushing (e.g. Blamey, Reference Blamey2012; Dassié et al. Reference Dassié, Genty, Noret, Mangenot, Massault, Lebas, Duhamel, Bonifacie, Gasparrini, Minster and Michelot2018) and analysing its elemental or isotopic composition by either mass spectrometry or gas chromatography. In spite of recent progress, allowing data to be obtained from down to 0.1 µL of fluid by crushing, the main limitation is related to the fact that it would be required to crush only primary/pseudo-secondary fluid inclusions, which appear to be complicated, even with a solid petrographic characterization. Böhlke and Irwin (Reference Böhlke and Irwin1992) showed that it was possible to analyse the fluid inclusions directly, but at the time it worked only for 10−11 L of fluid, in inclusions of very large dimensions (hundreds of µm). Laser absorption spectroscopy allows the reduction of these dimensions to a few tens of µm (Affolter et al. Reference Affolter, Fleitmann and Leuenberger2014). Yet future development in mass spectrometry on the one hand and in spotted sampling such as laser ablation on the other hand might lead to exciting developments enabling study of the content of selected fluid inclusion related to specific tectonic stages without crushing. A second method to access the fluid trapped in inclusions is to perform Raman spectroscopy on the fluid inclusion (Rosasco et al. Reference Rosasco, Roedder and Simmons1975). Raman spectroscopy is usually coupled to microthermometry to gradually heat up the inclusion during the measurement, and so to characterize and quantify the nature and amount of chemical components trapped in the fluid inclusion. This technique is mainly developed for oil-bearing fluid inclusions (e.g. Guillaume et al. Reference Guillaume, Teinturier, Dubessy and Pironon2003) and allows, when combined with other techniques, for proper computation of the chemical composition of the fluid inclusion, i.e. the isochore and isopleth curves (Bakker, Reference Bakker2003), or for direct determination of the salinity of aqueous fluid inclusions (Caumon et al. Reference Caumon, Dubessy, Robert and Tarantola2014). When aqueous fluid inclusion microthermometry is combined with coeval oil-bearing fluid inclusion microthermometry, it becomes possible to access the exact pressure and temperature of entrapment of the fluid (Pironon & Bourdet., Reference Pironon and Bourdet2008).

4.b. Established isotope-based techniques

4.b.1. Oxygen

Oxygen stable isotope measurements consist in measuring the ratio between the 18O and the 16O of the oxygen-bearing mineral (carbonates, but also quartz or clays), dividing it by the measured ratio between the 18O and 16O of a standard itself calibrated to a reference (Standard Mean Ocean Water (SMOW) for liquid phase, PeeDee Belemnite (PDB) for carbonates), as follows:

(1) $$ {{\rm{\delta }}^{18}}{\rm{O}} = \left( {{\raise0.7ex\hbox{${{{({}_{}^{18}{\rm{O}}/{}_{}^{16}{\rm{O}})}_{{\rm{sample}}}}}$} \!\mathord{\left/ {\vphantom {{{{({}_{}^{18}{\rm{O}}/{}_{}^{16}{\rm{O}})}_{{\rm{sample}}}}} {{{({}_{}^{18}{\rm{O}}/{}_{}^{16}{\rm{O}})}_{{\rm{standard}}}}}}}\right.\kern-0pt}\!\lower0.7ex\hbox{${{{({}_{}^{18}{\rm{O}}/{}_{}^{16}{\rm{O}})}_{{\rm{standard}}}}}$}} - 1} \right)*1000$$

The δ18O is expressed in per mil with respect to the reference (‰ PDB or ‰ SMOW). This technique can be performed using a few µg of pure carbonate (40 µg) obtained from machine-assisted sampling, or using laser or ionic ablation (Becker, Reference Becker2002). This ensures very easy data acquisition with minimal preparatory work, at very low cost and with relatively high precision (0.1−0.5 ‰ depending on the set-up used) (Swart et al. Reference Swart, Burns and Leder1991).

In the case of carbonates, studying oxygen isotopic ratio considers the exchange between the H2O content of the fluid and CaCO3. The formula that links the δ18O value of the carbonate to the δ18O of the water is as follows:

(2) $${{\rm{\delta }}^{18}}{{\rm{O}}_{{\rm{CaC}}{{\rm{O}}_3}}} = \;{{\rm{\delta }}^{18}}{{\rm{O}}_{{{\rm{H}}_2}{\rm{O}}}} + 1000\ln \alpha $$

The fractionation coefficient α is primarily controlled by the temperature at which the phase change occurs (precipitation or dissolution). Since the seminal work of Epstein et al. (Reference Epstein, Buchsbaum, Lowenstam and Urey1953) that used CaCO3 mineralized by gastropods to show a relationship between the δ18 CaCO3 and the temperature of the fluid, numerous experimental studies have established palaeothermometers valid for different temperatures (e.g. Craig, Reference Craig1957; Kim & O’Neil, Reference Kim and O’Neil1997; Zheng, Reference Zheng1999; Hu & Clayton, Reference Hu and Clayton2003; Chacko & Deines, Reference Chacko and Deines2008; Horita, Reference Horita2014). For instance, the most widely used equation might be that of Kim and O’Neil (Reference Kim and O’Neil1997):

(3) $$1000\ln {\alpha _{\left( {{\rm{CaC}}{{\rm{O}}_3} - {{\rm{H}}_2}{\rm{O}}} \right)}} = 18.03\left( {{{10}^{ - 3}}*{T^{ - 1}}} \right) - 32.42$$

where T is the phase transition temperature in kelvin. In other words, the δ18O value of the carbonate phase decreases as the temperature of precipitation increases. Even though the carbonate veins are mostly exclusively filled with calcite in orogenic forelands, it is worth noting that the equivalent relationships were also calibrated for the other carbonates (most used equations are gathered online by Beaudoin and Therrien, Reference Beaudoin and Therrien2004).

However, other physical and chemical processes affect the oxygen isotope fractionation in debated ways: the evaporation rate (Dreybrodt & Deininger, Reference Dreybrodt and Deininger2014) and the fluid composition and precipitation kinetics (Bottinga & Craig, Reference Bottinga and Craig1968; Rye & Bradbury, Reference Rye and Bradbury1988, Zeebe, Reference Zeebe2007). Most of these secondary factors seem negligible in a tectonic system where calcite precipitates in veins or in faults, yet the direct use of the δ18O value alone as a palaeothermometer is arguable. Indeed, both δ18OH2O and δ18OCaCO3 of the vein cement are required to access the temperature of precipitation of the cement at the time of deformation. The way around this is to assume the vein cement precipitated from a fluid that derives from the local fluids contained in the host rock. In that case, the value of δ18OH2O can be calculated using the measured δ18OCaCO3 in the host rock and a temperature of precipitation T assumed from the reconstruction of the surface temperature at the time the host rock precipitated. The value of T is then used in Eqn (3), the resulting value of 1000 ln α allowing for the calculation of the δ18OH2O using Eqn (2).

Assuming that the cement filling the fractures derives from the local fluid (seawater or freshwater) leads to the conjecture that a single fluid is involved in the system. That allowed many authors to explain an isotopic trend commonly observed in the fracture cements (Fig. 5). Indeed, the joint decrease of the δ18OCaCO3/δ13CCaCO3 values in vein cements related to burial and LPS is a recurring pattern (e.g. de Graaf et al. Reference de Graaf, Nooitgedacht, Goff, van der Lubbe, Vonhof and Reijmer2019), usually interpreted as a remobilization of local fluid during progressive burial, at increasing temperatures (Fig. 5a (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011), Fig. 5d (de Graaf et al. Reference de Graaf, Nooitgedacht, Goff, van der Lubbe, Vonhof and Reijmer2019)). Such interpretation needs to be supported by independent temperature constraints. In case a, homogenization temperatures obtained from fluid inclusion microthermometry support the idea that some of the cements characterized by δ18OCaCO3 >−15 ‰ PDB can be explained by cement precipitation from seawater during burial. Yet altogether the combination of T h and δ18OCaCO3 predicted that most of the cement had precipitated from a mixed fluid, highlighting extra-reservoir migrations. Cases b (Cruset et al. Reference Cruset, Cantarero, Travé, Vergés and John2016) and c (Curzi et al. Reference Curzi, Bernasconi, Billi, Boschi, Aldega, Franchini, Albert, Gerdes, Barberio and Carminati2021) in Fig. 5 considerably challenge this classical ‘burial’ interpretation. Indeed, independent temperatures of deformation-related cements obtained using clumped isotopes Δ47CO2 in both cases show that the cements exhibiting a more depleted δ18O precipitated at a significantly lower temperature (c. 30 °C) than the cements exhibiting a less depleted δ18O value. In that case, the isotopic pattern is interpreted as the precipitation from two different fluids (seawater and meteoric water) with no temperature effect relatable to burial or exhumation. These examples illustrate that an interpretation of a fluid system based solely on the classical δ18O–δ13C must be treated with care as it is easy to misinterpret the isotopic data. Following this line of argument, it is important not to use just the δ18O–δ13C of the vein and of the host rock to infer that the fluid is local or external to the reservoir. The isotopic equilibrium between the host rock and the vein cement is usually interpreted as due to rock buffering, i.e. a fluid/rock ratio strongly in favour of the rock, erasing the original fluid isotopic ratio. One needs to consider that with a fluid/rock ratio strongly in favour of the fluid, the isotopic ratio of the surrounding rock can be overprinted. Thus, it is important to couple with other techniques that will investigate the chemical equilibrium between the fluid and the host rock, such as elementary content.

Fig. 5. Simplified representation of δ18O vs δ13C plots obtained in syn-kinematic calcite cements from the fracture network (a, b, d) and in thrust fault zone (c), along with interpretation of the related fluid system. The published interpretations are reported on each graph. (a) The Sheep Mountain Anticline past fluid system (modified after Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011). (b) The Puig-Reig Anticline past fluid system (modified after Cruset et al. Reference Cruset, Cantarero, Travé, Vergés and John2016). (c) The Mount Tancia thrust fault zone past fluid system (modified after Curzi et al., Reference Curzi, Aldega, Bernasconi, Berra, Billi, Boschi, Franchini, van der Lelij, Viola and Carminati2020). (d) The past fluid system of the Albanide fold-and-thrust belt (modified after de Graaf et al. Reference de Graaf, Nooitgedacht, Goff, van der Lubbe, Vonhof and Reijmer2019). Note that the coloured frames in (a) locate the extension of plots (b–d). On all plots, the dashed grey frame represents the isotope values of the calcite fraction of the host rock when available. All values are given in ‰ PDB. T on plots refers to the measured temperatures of a given population, using an independent paleothermometer (T 47 for Δ47CO2, T h for fluid inclusion homogenization temperatures). LPS stands for layer-parallel shortening.

4.b.2. Carbon

Carbon stable isotope measurements in carbonate phases deal with the ratio between 13C and 12C, obtained with the same protocol as δ18O, and reported as ‰ PDB (see Eqn (1) for the formula, replacing O isotopes by C isotopes). In carbonates, the carbon is mainly provided by the dissolved carbonate species (HCO3, CO32−) and so the δ13C values of host rock carbonate are mostly similar to δ13C values of dissolved inorganic carbon, with other minor control parameters such as kinetics, pH and temperature of precipitation (Bottinga, Reference Bottinga1969). Similarly, the δ13C value of carbonate cement is bounded to the δ13C value of the parent fluid. The latter can vary significantly, in particular under the influence of organic matter degradation due to bacterial activity (e.g. sulphate reduction or methanogenesis) or thermal degradation (Irwin et al. Reference Arndt, Virgo, Cox and Urai1977; Curtis, Reference Curtis, Brooks and Welte1987). δ13C values of carbonates allow diagnosis of the bacterial activity related to oil and gas formation. δ13Cfluid is positive in the case methanogenesis occurred, while δ13Cfluid is negative in case there is an influence of a deeper methane source (δ13Cfluid shows values down to −60 ‰ PDB). In the case there is oil production, the δ13Cfluid values go down as far as −35 ‰ PDB (Peckmann & Thiel, Reference Peckmann and Thiel2004). δ13Cfluid of the inorganic carbon in seawater is c. 0 ‰ PDB while it is c. −10 ‰ PDB in meteoric water. However, interpreting a δ13C value in a calcite cement requires consideration of the mixing trends between end-members, δ13C values of which are often assumed because seldom accessible to measurements. As such, it is impossible to distinguish between (1) a partial mixing between a calcite precipitated from 70 % of pristine marine seawater (0 ‰ PDB) and 30 % of a seawater where organic carbon has been altered (−30 ‰ PDB) (considering the abundances of carbon are identical in both seawaters, the mix would have a value of −9 ‰ PDB) and (2) a calcite precipitated from 100 % pristine meteoric water (−10 ‰ PDB). As for δ18O values, the interpretation of δ13C remains ambiguous when the fluid source is unknown (Fig. 5). On top of this, recent experimental work has proven that δ13C values of carbonates precipitated from the same fluid source, at the same temperature, could present a shift of c. 3 ‰, due to the precipitation rate only (Yan et al. Reference Yan, Dreybrodt, Bao, Peng, Wei, Ma, Mo, Sun and Liu2021). This parameter being elusive when studying tectonic structures, such study supports that δ13C values must be used as a proxy with a great deal of care.

4.b.3. Δ47CO2 palaeothermometry

The clumping of isotopes, i.e. the association of isotopes into specific mass ion groups, so-called isotopologues, has been studied for decades, yet it is only recently that the concept has been developed for unravelling carbonate precipitation temperature conditions. After methodological locks were lifted by Eiler and Schauble (Reference Eiler and Schauble2004), showing it is possible to measure CO2 isotopologues with good accuracy, Ghosh et al. (Reference Ghosh, Adkins, Affek, Balta, Guo, Schauble, Schrag and Eiler2006) developed a thermometer based on the thermodynamical equilibrium of the doubly substituted isotopologue 13C18O16O. As defined by the authors, the enrichment of 13C18O16O in CO2 measured during acid digestion of CaCO3 relative to stochastic distribution of isotopes among all isotopologues is temperature-dependent. This enrichment, called Δ47CO2, relies on the fact the heavy isotopes 13C and 18O will be ordered in a different way (i.e. clumped or separated) according to the temperature. Since the clumping of heavy isotopes within a molecule is a purely stochastic process at high temperature but is systematically over-represented (relative to randomly distributing isotopes among molecules) at low temperature, the ‘absolute’ temperature of carbonate precipitation can be constrained using clumped-isotope abundances.

The coeval measure of Δ47CO2 and δ18O of a carbonate phase further allows reconstruction of both the absolute temperature of precipitation of the calcite and the δ18O value of the fluid it precipitated from, overcoming the classic limitation of δ18O measurements to reconstruct the past fluid temperature and origin. A growing community quickly applied this new tool that became a staple when it comes to characterization of diagenetic calcite and dolomite (Huntington et al. Reference Huntington, Budd, Wernicke and Eiler2011; MacDonald et al. Reference MacDonald, John and Girard2018; Mangenot et al. Reference Mangenot, Gasparrini, Rouchon and Bonifacie2018), fault cements (Swanson et al. Reference Swanson, Wernicke, Eiler and Losh2012; Bergman et al, Reference Bergman, Huntington and Crider2013) and vein cements (Pagel et al. Reference Pagel, Bonifacie, Schneider, Gautheron, Brigaud, Calmels, Cros, Saint-Bezar, Landrein, Sutcliffe, Davis and Chaduteau2018; MacDonald et al. Reference MacDonald, Faithfull, Roberts, Davies, Holdsworth, Newton, Williamson, Boyce and John2019; Beaudoin et al. Reference Beaudoin, Labeur, Lacombe, Koehn, Billi, Hoareau, Boyce, John, Marchegiano, Roberts, Millar, Claverie, Pecheyran and Callot2020 c; Hoareau et al.Reference Hoareau, Crognier, Lacroix, Aubourg, Roberts, Niemi, Branellec, Beaudoin and Suarez Ruiz2021 a; Labeur et al. Reference Labeur, Beaudoin, Lacombe, Emmanuel, Petracchini, Daëron, Klimowicz and Callot2021). Meanwhile, intensive calibration work has been conducted (Affek, Reference Affek2012; Zaarur et al. Reference Zaarur, Affek and Brandon2013; Hough et al. Reference Hough, Fan and Passey2014; Tang et al. Reference Tang, Dietzel, Fernandez, Tripati and Rosenheim2014; Wacker et al. Reference Wacker, Fiebig, Tödter, Schöne, Bahr, Friedrich, Tütken, Gischler and Joachimski2014; Bonifacie et al. Reference Bonifacie, Calmels, Eiler, Horita, Chaduteau, Vasconcelos, Agrinier, Katz, Passey, Ferry and Bourrand2017; Anderson et al. Reference Anderson, Kelson, Kele, Daëron, Bonifacie, Horita, Mackey, John, Kluge, Petschnig, Jost, Huntington, Bernasconi and Bergmann2021), with notable recent effort to improve the analytical accuracy (Daëron, Reference Daëron2021) and to reduce the persistent discrepancy between measurements performed in the various analytical set-ups (Bernasconi et al., Reference Bernasconi, Müller, Bergmann, Breitenbach, Fernandez, Hodell, Jaggi, Meckler, Millan and Ziegler2018, Reference Bernasconi, Daëron, Bergmann, Bonifacie, Meckler, Affek, Anderson, Bajnai, Barkan, Beverly, Blamart, Burgener, Calmels, Chaduteau, Clog, Davidheiser-Kroll, Davies, Dux, Eiler, Elliott, Fetrow, Fiebig, Goldberg, Hermoso, Huntington, Hyland, Ingalls, Jaggi, John, Jost, Katz, Kelson, Kluge, Kocken, Laskar, Leutert, Liang, Lucarelli, Mackey, Mangenot, Meinicke, Modestou, Müller, Murray, Neary, Packard, Passey, Pelletier, Petersen, Piasecki, Schauer, Snell, Swart, Tripati, Upadhyay, Vennemann, Winkelstern, Yarian, Yoshida, Zhang and Ziegler2021).

In spite of being widely used by the community, it is important to bear in mind the Δ47CO2 palaeothermometry is still a recent methodology, that suffers from an incomplete understanding of how diagenesis might affect the chemical record (Chen et al. Reference Chen, Ryb, Piasecki, Lloyd, Baker and Eiler2019; Hoareau et al. Reference Hoareau, Crognier, Lacroix, Aubourg, Roberts, Niemi, Branellec, Beaudoin and Suarez Ruiz2021 a; Nooitgedacht et al. Reference Nooitgedacht, van der Lubbe, de Graaf, Ziegler, Staudigel and Reijmer2021). The method is limited indeed to samples that have not been brought to temperature exceeding c. 180 °C, at which the solid-state reordering by isotopic diffusion of 13C and 18O becomes stochastic (Stolper & Eiler, Reference Stolper and Eiler2015). Solid-state reordering by isotopic diffusion is still active from lower temperature, and must systematically be corrected by thermodynamic modelling if the sample might have reached temperature >100 °C (e.g. Lloyd et al. Reference Lloyd, Ryb and Eiler2018). One efficient way of checking would be to systematically conduct fluid inclusion microthermometry on the same cements, which would have the additional benefit of quantifying the fluid pressure at the time of cement precipitation (e.g. Mangenot et al. Reference Mangenot, Bonifacie, Gasparrini, Götz, Chaduteau, Ader and Rouchon2017).

4.b.4. Strontium

Strontium has four natural isotopes (84Sr, 86Sr, 87Sr and 88Sr), including one related to radioactive decay (87Rb to 87Sr). However, as calcite cements are very poor in Rb, the radiogenic isotope ratio for strontium, 87Sr/86Sr, is very stable in calcite, making it a robust marker for fluid source and migration pathways (Graustein, Reference Graustein, Rundel, Ehleringer and Nagy1989). Indeed, this ratio is only related to the source fluid radiogenic signature, and to the potential mixing between fluid sources. The evolution of the 87Sr/86Sr value of seawater through geological time (McArthur et al. Reference McArthur, Howarth and Bailey2001) makes the 87Sr/86Sr value of mineralization a good way to distinguish whether the fluids are seawater or not, even though there are numerous overlaps in the radiogenic signature of seawater during the geological history. For instance, the typical radiogenic values for the Oligocene seawater (0.7078–0.7082) are significantly different from the values for the Pliocene seawater (0.7091–0.7092) but similar to the values of the early Triassic seawater (0.7078–0.7082). Typical ranges of 87Sr/86Sr values vary according to the epoch and age considered, being lower from the Dogger to the early Cretaceous (0.7068–0.7074) than since the Maastrichtian (0.7077–0.7092), for instance. In a stratigraphic sequence that is otherwise well constrained, it is then possible to distinguish a local seawater origin of the fluid from which cement precipitated in the fractures. In orogenic forelands, the mineralizing fluid could have been in frequent contact with K- or Rb-rich rocks such as clays, sandstones or granites, i.e. lithologies with a high radiogenic signature (87Sr/86Sr value > 0.710). Consequently, the 87Sr/86Sr value of a calcite cement has been used in numerous past fluid reconstructions, sometimes showing that fluids flowed in contact with radiogenic materials (Travé et al. Reference Travé, Calvet, Sans, Verges and Thirlwall2000; Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011, Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a), and sometimes preserving a pristine seawater signature in spite of being buried to metamorphic conditions (Dielforder et al. Reference Dielforder, Vollstaedt, Vennemann, Berger and Herwegh2015, Reference Dielforder, Villa, Berger and Herwegh2022). Even if the 87Sr/86Sr ratio does not vary during dissolution and precipitation, the recrystallization of calcite during diagenesis can release Sr in the fluid, altering its 87Sr/86Sr ratio (Dielforder et al. Reference Dielforder, Villa, Berger and Herwegh2022). Another example of misleading Sr radiogenic measurement has been documented in the vicinity of diapirs (Pichat et al. Reference Pichat, Hoareau, Callot, Legeay, Kavak, Révillon, Parat and Ringenbach2018), where the present-day seawater yields the radiogenic signature of the underlying evaporites. To reduce the risk of misinterpretation, a good understanding of the structural geology and a robust petrographic analysis to check that the calcite cement did not undergo recrystallization are mandatory. Also, the radiogenic value obtained can be due to mixing between different sources. Then, it is necessary to know the radiogenic value of each of the potential radiogenic rocks, i.e. the clays, evaporites, basement rocks, marine sediment, detrital-rich sandstone etc. This requires access to most of the sedimentary sequence, a condition that might not be fulfilled in a given study area.

4.b.5. Geochronology

The absolute dating of deformation by geochronology – a set of techniques that rely on the radioactive disintegration of parent isotopes in given parent–daughter isotope pairs – is long established, and there are a number of systems allowing the dating ofcarbonate cements related to tectonic activity. Some of these have restricted applicability, like Sm–Nd geochronology, which can be applied in contexts where mineralizing fluids are rich in REE. Such contexts include calcite veins associated to gold deposits (Xu et al. Reference Xu, Fan, Hu, Santosh, Yang, Lan and Wen2015), to basement-derived fluids (Barker et al. Reference Barker, Bennett, Cox, Norman and Gagan2009; Gao et al. Reference Gao, He, Zhao, He, Wu, Feng, Nguyen, Zhou and Yi2020) and to coal-bearing veins (Tonguç Uysal et al. Reference Tonguç Uysal, Zhao, Golding, Lawrence, Glikson and Collerson2007). U-series (also called U–Th) and U–Pb geochronology are increasingly popular in the field of tectonics thanks to their wider applicability to the calcite cements. Classically, the concentration in 230Th, 232Th, 234U and 238U for the U-series geochronology and in 235U, 238U, 207Pb, 206Pb and 204Pb for U–Pb geochronology is obtained by isotope dilution followed by thermal ionization mass spectrometry (TIMS) or inductively coupled plasma – mass spectrometry (ICP-MS), a technique that has a high precision and accuracy but is time-consuming and expensive (Brannon et al. Reference Brannon, Cole, Podosek, Ragan, Coveney, Wallace and Bradley1996; Rasbury & Cole, Reference Rasbury and Cole2009), even though recent development makes it faster (Engel et al. Reference Engel, Maas, Woodhead, Tympel and Greig2020). The recent reduction of the concentration threshold for the detection of trace elements using laser ablation coupled to ICP-MS (LA-ICP-MS) has enabled a democratization of LA-ICP-MS geochronology, that paved the way to obtaining the U–Pb absolute age of syn-kinematic calcite cements in veins (see Roberts et al., Reference Roberts, Drost, Horstwood, Condon, Chew, Drake, Milodowski, McLean, Smye, Walker, Haslam, Hodson, Imber, Beaudoin and Lee2020, for a review) and faults (see Roberts & Holdsworth, Reference Roberts and Holdsworth2022, for a review), with further developments of U–Pb dating of dolomite phases (Bar et al. Reference Bar, Nuriel, Kylander-Clark and Weinberger2021; Motte et al. Reference Motte, Hoareau, Callot, Revillon, Piccoli, Calassou and Gaucher2021; Su et al. Reference Su, Chen, Feng, Zhao, Wang, Hu, Jiang and Duc Nguyen2022). In contrast, the very low 230Th concentration in carbonates severely hampers the applicability of LA-ICP-MS to U-series geochronology (e.g. Lin et al. Reference Lin, Jochum, Scholz, Hoffmann, Stoll, Weis and Meinrat2017), which is moreover restricted to ages younger than 500 to 600 kyr (Villemant & Feuillet, Reference Villemant and Feuillet2003; Andersen et al. Reference Andersen, Stirling, Potter, Halliday, Blake, McCulloch, Ayling and O’Leary2008), thereby restricting the application to recent tectonic events. Alternatively, U–Pb does not suffer that limit and allows dating of cements as old as 108 years. The growing size of the community using U–Pb geochronology leads to a variety of data treatment that permits either treating the cement as a bulk or providing a mapping of U–Pb ratio in microstructures (Drost et al. Reference Drost, Chew, Petrus, Scholze, Woodhead, Schneider and Harper2018; Hoareau et al. Reference Hoareau, Claverie, Pecheyran, Paroissin, Grignard, Motte, Chailan and Girard2021 b), and also an interest in expanding the limits of LA-ICP-MS detection (Kylander-Clark, Reference Kylander-Clark2020). Beyond being fast and relatively cheap, the LA-ICP-MS U–Pb geochronology has been successfully compared to other geochronometers (Mottram et al. Reference Mottram, Kellett, Barresi, Zwingmann, Friend, Todd and Percival2020) like K–Ar, a system that has long been used in tectonics, especially through the dating of neoformed illite in fault gouges by K–Ar or Ar–Ar geochronology (Lyons & Snellenburg, Reference Lyons and Snellenburg1971; van der Pluijm et al. Reference van der Pluijm, Hall, Vrolijk, Pevear and Covey2001, Reference van der Pluijm, Vrolijk, Pevear, Hall and Solum2006; Zwingmann et al. Reference Zwingmann, Offler, Wilson and Cox2004; Haines & van der Pluijm, Reference Haines and van der Pluijm2008; Clauer, Reference Clauer2013; Viola et al. Reference Viola, Zwingmann, Mattila and Käpyaho2013; Hnat & van der Pluijm, Reference Hnat and van der Pluijm2014; Scheiber et al. Reference Scheiber, Viola, van der Lelij, Margreth and Schönenberger2019).

However, it is important to underline that calcite does commonly incorporate more Pb than U, making it difficult to date in numerous cases (Mottram et al. Reference Mottram, Kellett, Barresi, Zwingmann, Friend, Todd and Percival2020). Also, the role of diagenesis on the incorporation of U and Pb in carbonate minerals remains debated (Roberts et al. Reference Roberts, Drost, Horstwood, Condon, Chew, Drake, Milodowski, McLean, Smye, Walker, Haslam, Hodson, Imber, Beaudoin and Lee2020). The potential Pb redistribution by solid-state diffusion related to crystal-scale deformation has been observed in other minerals (rutile) and causes variation of the obtained age at the mineral scale (Moore et al. Reference Moore, Beinlich, Porter, Talavera, Berndt, Piazolo, Austrheim and Putnis2020). Thus, the diagenetic state of the calcite vein cement has to be systematically studied, along with its elementary content, in order to properly perform the analysis. Another important point is that the ages obtained from vein cements might not be straightforwardly interpreted as representative of the entire duration of the deformation event that developed the veins, so a certain number of data might be required to cover the entire timespan during which a deformation phase affected the rock (Beaudoin et al. Reference Beaudoin, Lacombe, Roberts and Koehn2018). In faults, the age of a cement can be younger than the fault activity because of a subsequent fluid flow (Roberts et al. Reference Roberts, Žák, Vacek and Sláma2021; Roberts & Holdsworth, Reference Roberts and Holdsworth2022). To check whether the U–Pb age legitimately represents the age of precipitation, a recent approach couples the Δ47CO2 palaeothermometry to U–Pb geochronology to check whether the vein cement precipitated at thermal equilibrium with the host or not (Pagel et al. Reference Pagel, Bonifacie, Schneider, Gautheron, Brigaud, Calmels, Cros, Saint-Bezar, Landrein, Sutcliffe, Davis and Chaduteau2018; Hoareau et al. Reference Hoareau, Crognier, Lacroix, Aubourg, Roberts, Niemi, Branellec, Beaudoin and Suarez Ruiz2021 a; Muñoz-López et al. Reference Muñoz-López, Cruset, Vergés, Cantarero, Benedicto, Mangenot, Albert, Gerdes, Beranoaguirre and Travé2022). In the Spanish Pyrenean foreland, Hoareau et al. (Reference Hoareau, Crognier, Lacroix, Aubourg, Roberts, Niemi, Branellec, Beaudoin and Suarez Ruiz2021 a) used this combination to validate that the cements precipitated during the last uplift phase and were not subject to any later burial, preserving their original Δ47CO2 values.

4.c. Developing isotope-based techniques

Being the main constitutive element of carbonates, calcium has five stable isotopes, two ratios being used to reconstruct carbonate diagenesis: the δ44/40Ca and δ44/42Ca, the former being the most used and reported as δ44Ca (Swart, Reference Swart2015). As the δ44Ca varies according to the fluid source, it is mainly used to characterize the source of calcium during events such as evaporation, weathering or dolomitization (e.g. Fantle & Higgins, Reference Fantle and Higgins2014; Tostevin et al. Reference Tostevin, Bradbury, Shields, Wood, Bowyer, Penny and Turchyn2019). The fractionation of δ44Ca between the fluid and a given mineral is related to temperature (Gussone et al. Reference Gussone, Ahm, Lau and Bradbury2020) and precipitation rate (Tang et al. Reference Tang, Dietzel, Böhm, Köhler and Eisenhauer2008). The few studies that rely upon application of δ44Ca to calcite cement in tectonic structures discussed the interaction between the fluid and the basement, or the origin of dolomitization (Husson et al. Reference Husson, Higgins, Maloof and Schoene2015).

Numerous secondary minerals can be found associated to carbonates, enabling the study of isotopes that can help unravel the past fluid system. Hydrogen isotopic ratio (δ2H or δD) is classically used along with δ18O values, both measured in the clay fractions in deformation features to access the fluid source. In fact, a restricted range of δ2H values and associated δ18O values of H2O corresponds to meteoric waters (Craig, Reference Craig1961; Sheppard, Reference Sheppard1986). This approach has been used in several case studies where δ2H–δ18O values of clay fractions in fault gouges determine the ingress within major faults to either meteoric or deep-sourced fluids (e.g. Haines et al. Reference Haines, Lynch, Mulch, Valley and van der Pluijm2016; Lynch & van der Pluijm, Reference Lynch and van der Pluijm2017; Lynch et al. Reference Lynch, Mulch, Yonkee and van der Pluijm2019, Reference Lynch, Pană and van der Pluijm2021). 7Li/6Li isotopic ratio (δ7Li) may prove valuable in the reconstruction of the past fluid system. While relatively recent, the fractionation of Li is related to the degree of weathering of the Li-rich igneous rocks. δ7Li is proven to be a good proxy for weathering intensity (Dellinger et al. Reference Dellinger, Gaillardet, Bouchez, Calmels, Louvat, Dosseto, Gorge, Alanoca and Maurice2015), and while it has been applied in modern rivers to date, future research could include a possible use on clay fractions to decipher the potential fluid migration pathway through the basement in orogenic forelands.

Ballentine and Burnard (Reference Ballentine and Burnard2002) proposed that noble gases transported in fluids can be used to assess fluid origin and transportation process. Among these, the radiogenic isotope ratio for helium, 3He/4He, is considered as the most efficient tracker of a deep source of fluid, either crustal or mantellic (e.g. Pik & Marty, Reference Pik and Marty2009) and has been applied to syntectonic calcite veins (Smeraglia et al. Reference Smeraglia, Bernasconi, Berra, Billi, Boschi, Caracausi, Carminati, Castorina, Doglioni, Italiano, Rizzo, Uysal and Zhao2018, Reference Smeraglia, Giuffrida, Grimaldi, Pullen, La Bruna, Billi and Agosta2021) to test whether the fault zone was a drain for deep-seated mantle-derived fluids. As gas will be accessed using crushing techniques (e.g. Blamey, Reference Blamey2012), a very rigorous petrographic study of fluid inclusions and their relation to diagenesis must be done beforehand. Recent efforts based on carbonate-bearing hydrocarbon seep-related deposits in the Cantabrian Basque Basin suggest that the use of neodynium isotopes in carbonates, ϵ Nd, is also a promising proxy to assess the volcanic origin of the fluids (Jakubowicz et al. Reference Jakubowicz, Agirrezabala, Belka, Siepak and Dopieralska2022).

4.d. Elemental geochemistry

Elemental geochemistry relies upon studying the elemental composition of the mineral phases (here, calcite) by means of either bulk dissolution or electron probe microanalysis on sections (e.g. Centrella et al. Reference Centrella, Putnis, Lanari and Austrheim2018). The concentrations of several trace elements in calcite (Mn, Fe, Sr …) can be interpreted in terms of fluid characteristics such as redox conditions at the time of precipitation (e.g. Curtis et al. 1986), or their origin (meteoric vs marine; e.g. Travé et al. Reference Travé, Labaume and Vergés2007). In the published studies that use elemental geochemistry, the concentration of an element of the fluid is commonly calculated using a partition coefficient (Kd) (Parekh et al. Reference Parekh, Möller, Dulski and Bausch1977) and the measured concentration of this element in the cement. The Kd is considered over a range of temperatures, and the concentration in the element of the fluid is compared to the expected concentration in each fluid source (meteoric, marine, deep-sourced, brine; e.g. Travé et al. Reference Travé, Labaume, Calvet and Soler1997; Lacroix et al., Reference Lacroix, Trave, Buatier, Labaume, Vennemann and Dubois2014). By doing so, researchers can derive the origin of the fluid, for example using one or all of Mg/Ca, Sr/Ca, Mn/Ca, Ca/Fe ratios. This approach remains arguable, because a number of assumptions have to be made. First, the precipitation of the crystal at chemical equilibrium with the fluid is of prime importance here, and such an assumption might be discussed (Paquette & Reeder, Reference Paquette and Reeder1995). Second, the range of Kd used to calculate the past fluid elemental concentration comes from a specific experimental dataset where the fluid chemistry is by definition different from the precipitation fluid in the geological system, yet the Kd depends strongly on the fluid chemistry. In other words, an implicit assumption is made that for a given fluid origin (meteoric, basinal, marine) the chemistry is the same regardless of the context and remains constant in space and time. Nonetheless, elemental geochemistry has been applied in combination with other proxies to characterize the nature of the fluid. For instance the study of Travé et al. (Reference Travé, Calvet, Sans, Verges and Thirlwall2000) documents an alternation between flows of formation water and meteoric water along the major thrust zone of the El Guix anticline in the southern Pyrenees.

Rare earth elements (REE) are also used to reconstruct the fluid origin by means of a spider diagram, a diagram reporting normalized (to chondrite) REE concentrations from the lightest (La) to the heaviest (Lu). The pattern of the diagram might be diagnostic of the fluid source (Motte et al. Reference Motte, Hoareau, Callot, Revillon, Piccoli, Calassou and Gaucher2021), as a negative anomaly in caesium is typical of seawater (Shields & Webb, Reference Shields and Webb2004), and a positive anomaly in europium supports deep-sourced fluids (Bau & Möller, Reference Bau and Möller1993). REE patterns have been successfully used to highlight (1) the migration of hydrothermal, potentially deep-sourced, fluids in the Abu Dhabi area (Morad et al. Reference Morad, Al-Aasm, Sirat and Sattar2010); (2) large-scale fluid flow through crack-seal veins in the Eastern Belt of the Lachlan orogen (Australia) (Barker et al. Reference Barker, Cox, Eggins and Gagan2006); (3) the link between travertine veins and hot fluid pulses in Turkey (Uysal et al. Reference Uysal, Feng, Zhao, Isik, P and SD2009); or (4) the seawater origin of the cement filling tectonic veins related to the Shizhu synclinorium (China) (Wang et al. Reference Wang, Gao, He, He, Zhou, Tao, Zhang and Wang2017). However, it is important to consider that this approach relies on the debated assumption that the precipitation occurs at equilibrium with the fluid, and with no specific different distribution effect from one REE to another.

Other elemental analyses of interest can be obtained from the crushing of fluid inclusions, which grants access to the molecular ratios of the analysed gas, such as N2/Ar, Ar/He or CO2–SH4–H2, and can help constrain if the fluid is derived from a meteoric source, a basinal source or its redox condition, respectively. Where conditions of concentration and preservation are met, a geothermometer can be derived from gas content, the most used one being the CO2/CH4 – H concentrations (D’Amore & Panichi, Reference D’Amore and Panichi1980; Henley et al. Reference Henley, Truesdell, Barton and Whitney1984; Giggenbach, Reference Giggenbach1996).

5. Reconstruction of the past fluid system at the local (fold/thrust) scale: case studies

5.a. Brief review of past fluid system studies over the past decade

At the scale of individual contractional structures like folds, numerous studies focused on the reconstruction of the past fluid system and its evolution during the deformation stages (LPS, fold growth, LSFT; see Section 2) of the so-called folding event (Lacombe et al. Reference Lacombe, Beaudoin, Hoareau, Labeur, Pecheyran and Callot2021). In these studies, effort was made to characterize the origin of the fluids and to picture the plumbing system. Mozafari et al. (Reference Mozafari, Swennen, Balsamo, Clemenzi, Storti, El Desouky, Vanhaecke, Tueckmantel, Solum and Taberner2015, Reference Mozafari, Swennen, Muchez, Vassilieva, Balsamo, Storti, Pironon and Taberner2017) used δ18O, δ13C, 87Sr/86Sr data along with fluid inclusion microthermometry and chemistry by Raman spectroscopy to show that in the Jabal Qusaybah anticline (Oman Mountains), fluids were rather local during LPS and evolved towards brines that interacted with siliciclastic strata during fold growth. Still in the Oman Mountains, Arndt et al. (Reference Arndt, Virgo, Cox and Urai2014) reconstructed a 50 m vertical fluid migration along the Jebel Shams fault zone using stable isotope geochemistry. In the southern Pyrenean foreland, fracture cements of the Puig Reig anticline, formed above a cover duplex, were studied by means of δ18O, δ13C, Δ47CO2, 87Sr/86Sr data, fluid inclusion microthermometry along with elemental geochemistry of carbonate and fluid inclusions. Results revealed a lateral, reservoir-bounded forelandward fluid migration of hydrothermal brines along the thrust during LPS (Cruset et al. Reference Cruset, Cantarero, Travé, Vergés and John2016). However, the cements within the curvature-related fractures precipitated from a mixture between downward-flowing meteoric fluids and upward-flowing brines (Cruset et al. Reference Cruset, Cantarero, Travé, Vergés and John2016), suggesting a vertical fluid flow, hence the opening of the fluid system during folding. Still in the southern Pyrenean foreland, the fluid system within the Bóixols – Sant Corneli anticline was studied by means of δ18O, δ13C, Δ47CO2, 87Sr/86Sr and elementary composition of the carbonate. Interpretations draw a fluid system first closed, then gradually opened to meteoric fluids flowing downwards, the décollement level being interpreted as an efficient barrier to deep-sourced fluids (Nardini et al. Reference Nardini, Muñoz-López, Cruset, Cantarero, Martín-Martín, Benedicto, Gomez-Rivas, John and Travé2019; Muñoz-López et al. Reference Muñoz-López, Cruset, Vergés, Cantarero, Benedicto, Mangenot, Albert, Gerdes, Beranoaguirre and Travé2022). In the Montagna dei Fiori, an anticline from the Umbria Marche Apennine Ridge, Mozafari et al. (Reference Mozafari, Swennen, Balsamo, El Desouky, Storti and Taberner2019) used δ18O, δ13C, 87Sr/86Sr data along with fluid inclusion microthermometry to reconstruct two pulses of hydrothermal fluids sourced by evaporitic brines, one related to the pre-contractional history, i.e. the Jurassic rifting, and a second one related to the onset of the growth of the anticline. In the Cingoli anticline (Apennines, Italy), a fault-bend fold linked to a thrust the rooting depth of which is debated (Triassic evaporites vs basement), Labeur et al. (Reference Labeur, Beaudoin, Lacombe, Emmanuel, Petracchini, Daëron, Klimowicz and Callot2021) showed, using δ18O, δ13C, Δ47CO2 data, that the reservoir remained a compartmentalized system where syn-kinematic vein cements precipitated from thermally equilibrated evolved formational fluids during the whole fold development. In the Jura Mountains, the geochemistry (δ18O, δ13C, Δ47CO2, 87Sr/86Sr) of cements in fault zones and veins within folded strata above the internal Jura frontal thrust depicted a progressive opening of the reservoir to surficial meteoric fluids cooler than the environment (Smeraglia et al. Reference Smeraglia, Fabbri, Choulet, Buatier, Boulvais, Bernasconi and Castorina2020 a). In the Bornes Massif (Subalpine belts), the reconstruction of the past fluid system associated to the development of the Parmelan anticline was based on δ18O, δ13C, Δ47CO2, 87Sr/86Sr data, fluid inclusion microthermometry along with elemental geochemistry of carbonate. The fluid system was stratigraphically compartmentalized with hot meteoric fluids flowing laterally in the sedimentary cover during LPS and folding, then opened to meteoric fluids during the late stages of folding and during post-folding deformation development (Berio et al., Reference Berio, Mittempergher, Storti, Bernasconi, Cipriani, Lugli and Balsamo2022). In the Vercors chain in southeastern France, Bilau et al. (Reference Bilau, Bienveignant, Rolland, Schwartz, Godeau, Guihou, Deschamps, Mangenot, Brigaud, Boschetti and Dumont2022) used U–Pb geochronology along with δ18O, δ13C and Δ47CO2 on syntectonic calcite veins and thrusts cements to picture a past fluid system dominated by heated brines.

5.b. General evolutionary trend of the past fluid system at the scale of the fold during the folding event

These recent studies support the conclusions presented in the review work conducted a decade ago by Evans and Fischer (Reference Evans and Fischer2012). In many cases, the hydraulic structure of a folded reservoir evolves in a similar way from LPS to folding regardless of the tectonic settings, yet the nature of the fluids involved varies according to the structure (Fig. 6). During LPS, cements generally precipitated from formational fluid, suggesting a limited inter-reservoir connectivity. The fluid system can be considered either as closed if the fluid is formational with limited migration or as open if the formational fluid flowed laterally in the reservoir which remains stratigraphically compartmentalized (Fig. 4). In contrast, during fold growth, a switch occurs in the chemical signature of the cements and of the related fluids, with a notable input of either meteoric-derived fluids, saline brines or basement-derived fluids flowing upwards or downwards, according to the occurrence of an efficient conduit underneath the folded reservoir, in most cases a fault, but also the outer arc-related joints (e.g. Lefticariu et al. Reference Lefticariu, Perry, Fischer and Banner2005; Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a). Such fluid input proceeded to homogenize the fluid system at the reservoir scale, with various degrees of homogenization going from a mixing to a total overprint. Considering that the time required for thermal equilibration between the fluid and the rock is instantaneous at the geological timescale (e.g. Heinze et al. Reference Heinze, Hamidi and Galvan2017), the fact a cement can precipitate at a temperature higher (or lower) than that of the surrounding host rock requires that the fluid did not thermally equilibrate before cement precipitated. That suggests (1) high flow and high precipitation rates together with limited exchanges with the host rock or (2) a very fast crystal growth, in line with the rate of few micrometres per second obtained in experimental set-ups (Lee & Morse, Reference Lee and Morse1999; Hilgers & Urai, Reference Hilgers and Urai2002 a, b; Hilgers et al. Reference Hilgers, Dilg-Gruschinski and Urai2004). Sometimes, however, the fluid system remains closed through the entire folding event (e.g. Labeur et al. Reference Labeur, Beaudoin, Lacombe, Emmanuel, Petracchini, Daëron, Klimowicz and Callot2021).

Fig. 6. Representation of the first-order evolution of the past fluid system at the scale of the individual fold or/and thrust. From an initial state related to layer-parallel shortening (left-hand side), the fluid system evolved according to the fault distribution/the décollement nature, and to the structural style of the structure. The resulting fluid from which cement precipitates is symbolized by a hexagon whose colour is related to the nature of the fluid, several colours representing a mixing with no implication on the ratio. Fluid temperature with respect to the local temperature is symbolized by C when cooler, H when hotter and E when at equilibrium.

6. How geochemistry helps reconstruct the origin and large-scale migrations of fluids in forelands: insights from selected case studies

Most studies of past fluid systems were actually carried out at the scale of the foreland. Indeed, a past fluid flow obviously is not limited to individual structures like a fold or a thrust, and larger-scale processes are usually involved when it comes to fluid migration. For instance, the large-scale pressure and topographic gradients required for the migration of extra-reservoir fluids that distribute hydrocarbons and ore deposits like Pb–Zn Mississippian Valley-type (MVT) cannot be explained at the scale of the individual fold or thrust (Cathles, Reference Cathles1981; Roure et al. Reference Roure, Swennen, Schneider, Faure, Ferket, Guilhaumou, Osadetz, Robion and Vandeginste2005). Hereinafter, we review how geochemical proxies helped to reconstruct the past fluid system which prevailed at the regional scale during the development of an orogenic foreland. To do that, we selected well-documented case studies: (1) the thin-skinned Canadian Rocky Mountains (Fig. 7) which extend southwards in the USA as the Rockies–Sevier belt (Fig. 8); (2) the thick-skinned Laramide province (USA; Fig. 8). Because the previous orogenic domains represent end-members in terms of tectonic style of deformation, hybrid cases where thin-skinned and thick-skinned deformations are superimposed are also considered; (3) the south Pyrenean FTB (Spain; Fig. 9), where the basement is also involved in shortening at depth under the internal Sierras; and (4) the Umbria–Marche Apennine Ridge in the central Apennines (Italy; Fig. 10), where the degree of basement involvement in shortening and its impact on the cover-scale deformation is debated.

Fig. 7. Simplified cross-section and corresponding location of the Canadian Rocky Mountains fold-and-thrust belt where the past fluid system was reconstructed. The vertical exaggeration is c. 7:1. FTB stands for fold-and-thrust belt. The legend is transferable to the other cross-sections of Figures 810. In the present case, the decoupling level comprises argillites (Proterozoic). See text for details and references.

Fig. 8. Simplified cross-section and corresponding location of the Sevier range and Laramide province (USA) where the past fluid system was reconstructed. On the location map, the yellow colour indicates the intracratonic basins related to the exhumation of basement arches. The vertical exaggeration is about 2:1. WRB: Wind-River Basin; BHB: Bighorn Basin; PRB: Powder River Basin. See text for details and references. Key is given in Figure 7.

Fig. 9. Simplified cross-section and corresponding location of the southern Pyrenean FTB (Spain), where the past fluid system was reconstructed along the thrusts. The vertical exaggeration is c. 3:1. See text for details and references. Key is given in Figure 7. The nature of the décollement is evaporites (Triassic).

Fig. 10. Simplified cross-section and corresponding location of Umbria Marche Apennine Ridge (Italy), where the past fluid system was reconstructed. The vertical exaggeration is c. 2:1. See text for details and references. Key is given in Figure 7. The nature of the décollement is evaporites (Triassic).

6.a. The Canadian Rocky Mountains

One of the most iconic reconstructions of the past fluid system in a FTB might well be the one conducted in the Canadian Rockies and the Western Canada Sedimentary Basin (WCSB). The Canadian Rockies developed from the Upper Jurassic (155 Ma; DeCelles, Reference DeCelles2004) as an effect of the subduction of the Farallon plate under the North America plate, leading to a thin-skinned belt referred to as the Sevier belt, going from Canada to Mexico (Fitz-Diaz et al. Reference Fitz-Diaz, Hudleston, Tolson, Poblet and Lisle2011 b). In the Canadian Rockies and the WCSB, the contractional deformation has been accommodated in the sedimentary cover mainly through detachment folding to the north and thrust-related folding to the south, forming a belt up to 200 km wide. It is a paragon of an orogenic wedge where thrusts developed in a cratonward sequence (Paña & van der Pluijm, Reference Pană and van der Pluijm2015) above a viscous décollement (Fig. 7). The large amounts of hydrothermal dolomite in Devonian carbonates, associated to important MVT ore deposits, are striking remnants of the fluid flows that have been characterized thanks to numerous studies (Mountjoy et al. Reference Mountjoy, Qing and McNutt1992; Ge & Garven, Reference Ge and Garven1994; Hairuo Qing & Mountjoy, Reference Mountjoy1994; Nesbitt & Muehlenbachs, Reference Nesbitt and Muehlenbachs1995; Machel & Cavell, Reference Machel and Cavell1999; Al-Aasm et al. Reference Al-Aasm, Lonnee and Clarke2002; Vandeginste et al. Reference Vandeginste, Swennen, Allaeys, Ellam, Osadetz and Roure2012). The sum of the work done – mainly based on O, C and Sr isotopic analyses, so quite ambiguous; see Section 4 – depicted a debated past fluid flow during the contraction. In a first view, large-scale flow of basinal brines, pushed forelandwards through a squeegee-type flow due to the belt development and propagation, would explain the extensive dolomitization together with ore deposition (Bradbury & Woodwell, Reference Bradbury and Woodwell1987; Qing & Mountjoy, Reference Qing and Mountjoy1992; Ge & Garven, Reference Ge and Garven1994; Al-Aasm et al. Reference Al-Aasm, Lonnee and Clarke2000). In an alternative view, the fluid migrations have been limited to inter-reservoir migrations, with local and transient hot-flash-type flow of basement-derived hydrothermal fluids during thrust activity (Nesbitt & Muehlenbachs, Reference Nesbitt and Muehlenbachs1995; Machel & Cavell, Reference Machel and Cavell1999; Kirschner & Kennedy, Reference Kirschner and Kennedy2001; Cooley et al. Reference Cooley, Price, Kyser and Dixon2011; Vandeginste et al. Reference Vandeginste, Swennen, Allaeys, Ellam, Osadetz and Roure2012). Building on the geometry of the belt, numerical simulations of fluid flow predict that this lateral fluid migration in the reservoir occurred in response to thrusting at a speed of c. 2 km Myr−1 ( Schneider, Reference Schneider2003). The difference between the two past fluid system models lies in the interpretation of high 87Sr/86Sr values (>0.7120): while Qing and Mountjoy (Reference Qing and Mountjoy1992) interpreted these in combination with elementary content as a marker of fluid migrations in contact with the basement rocks, Vandeginste et al. (Reference Vandeginste, Swennen, Allaeys, Ellam, Osadetz and Roure2012) instead interpreted them in combination with δ18O values as a witness of the contamination of the fluid by K-rich sedimentary strata. To support their revision of the former interpretation of the past fluid system, Vandeginste et al. (Reference Vandeginste, Swennen, Allaeys, Ellam, Osadetz and Roure2012) invoke the impact of a complex diagenetic history, that may have altered fracture cements, enriching the original 87Sr/86Sr content of the fluid. In their interpretation, Vandeginste et al. (Reference Vandeginste, Swennen, Allaeys, Ellam, Osadetz and Roure2012) proposed that the notable input of meteoric fluid in the system was related to the post-orogenic evolution. By analysing clays in fault zones, Lynch et al. (Reference Lynch, Pană and van der Pluijm2021) validated that the faults were efficient pathways to surficial, meteoric fluids, that were dominating (up to 90 %) the system compared to the existing but limited volume of metamorphic fluids.

6.b. The Laramide province (USA)

The Laramide province in the USA is an eastward continuation of the Sevier belt that extends from the southern part of Montana to the southern part of New Mexico (N–S) and from Arizona to South Dakota (W–E). From the end of Cretaceous times, a change in the mantle dynamic affected the subducting slab and led to a coupling between the slab and the lower crust (English et al. Reference English, Johnston and Wang2003; Yonkee & Weil, Reference Yonkee and Weil2015), changing the direction of the shortening (Craddock & van der Pluijm, Reference Craddock and van der Pluijm1999) and reactivating inherited Precambrian structures (Marshak et al. Reference Marshak, Karlstom and Timmons2000). The result of this history is a typical broken foreland, the Laramide province (Stone, Reference Stone1967; Erslev, Reference Erslev1986; Marshak et al. Reference Marshak, Karlstom and Timmons2000), where endorheic basins are separated by basement highs over a distance of 400 km (Fig. 8). Part of this broken foreland, the Bighorn Basin (Wyoming, USA), has been extensively studied for its fracture network and associated past fluid system. This Palaeocene–Eocene basin of c. 50 km width hosts basement-cored folds for which fracture patterns and related stress history have been extensively discussed (Craddock & van der Pluijm, Reference Craddock and van der Pluijm1989, Reference Craddock and van der Pluijm1999; Craddock & Relle, Reference Craddock and Relle2003; Bellahsen et al. Reference Bellahsen, Fiore and Pollard2006 a, b; Neely & Erslev, Reference Neely and Erslev2009; Amrouch et al. Reference Amrouch, Robion, Callot, Lacombe, Daniel, Bellahsen and Faure2010 a, b, Reference Amrouch, Beaudoin, Lacombe, Bellahsen and Daniel2011; Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011, Reference Beaudoin, Leprêtre, Bellahsen, Lacombe, Amrouch, Callot, Emmanuel and Daniel2012, Reference Beaudoin, Lacombe, Bellahsen, Amrouch and Daniel2014b, Reference Beaudoin and Lacombe2018, Reference Beaudoin, Lacombe, Roberts and Koehn2019 a, Reference Beaudoin, Lacombe, David and Koehn2020 a; Thacker & Karlstrom, Reference Thacker and Karlstrom2019). In the basin, the mesostructures developed during a deformation history that comprises three main stages: (1) pre-folding LPS fractures formed in response to the far-field stress transfer from the distant thin-skinned Sevier belt; (2) early folding LPS fractures formed as a result of Laramide contraction; (3) local extensional fractures striking parallel to fold axes and related to outer-arc extension due to strata curvature at fold hinges during Laramide fold growth. Studying the past fluid system associated to the fracture network enables one to reconstruct the evolution of the fluid sources, conditions of precipitation and migration pathways during the tectonic evolution, from being in the Sevier flexed foreland to becoming part of the Laramide broken foreland. Such reconstructions were based on O, C isotopic geochemistry, 87Sr/86Sr ratios and fluid inclusion microthermometry on calcite cements of veins (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011, Reference Beaudoin, Leprêtre, Bellahsen, Lacombe, Amrouch, Callot, Emmanuel and Daniel2012, Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a; Barbier et al. Reference Barbier, Leprêtre, Callot, Gasparrini, Daniel, Hamon, Lacombe and Floquet2012 a, Reference Barbier, Hamon, Callot, Floquet and Daniel2012 b, Reference Barbier, Floquet, Hamon and Callot2015) and more recently on U–Pb absolute dating (Beaudoin et al. Reference Beaudoin, Lacombe, Roberts and Koehn2018, Reference Beaudoin, Lacombe, Roberts and Koehn2019 a).

Combination of δ18O values with T h from fluid inclusions revealed that the fluids involved in the Ordovician–Permian reservoir during the Laramide deformation were a mixture of Palaeogene meteoric fluids and Palaeozoic seawater (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011; Barbier et al. Reference Barbier, Hamon, Callot, Floquet and Daniel2012 b). High 87Sr/86Sr signatures (>0.710) and high Th (>120 °C) suggested that the meteoric fluids first flowed into / in contact with the basement (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a). This constrains the potential migration pathways, with a recharge zone of meteoric fluids in the highly elevated Sevier range to the west (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a) and to the southwest (Barbier et al. Reference Barbier, Hamon, Callot, Floquet and Daniel2012 b). From there, a (north)eastward migration across the basin is suggested by the eastward decrease of the radiogenic values of the vein cements during the basin-scale LPS (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a). At the time of folding, the radiogenic and heated fluids were expelled as pulses along the inherited basement faults reactivated as high-angle thrusts and flowed upwards into the sedimentary cover through the vertically persistent fractures related to outer-arc extension at fold hinges, overprinting the former, stratigraphically compartmentalized local fluid system. Pulses of hydrothermal, meteoric-derived fluids flowing upwards from the basement–cover interface also occurred earlier and later in the tectonic history of the basin, during the forebulge development related to the Sevier orogeny and during the post-orogenic (post-Laramide) extension (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a). This evolution suggests a recharge mechanism for downward migration of meteoric fluids in the inner belt and an eastward squeegee-type flow (see Section 2.b) enhanced by an early basin-scale topographic gradient due to the early exhumation of the Bighorn Mountains easternmost basement arch (Fig. 8). This scenario of a large-scale eastward fluid migration was also supported by the reconstruction of the evolution of past fluid pressure at the basin scale using calcite twinning palaeopiezometry, fracture analysis and rock mechanics (Amrouch et al. Reference Amrouch, Beaudoin, Lacombe, Bellahsen and Daniel2011, Beaudoin et al. Reference Beaudoin, Lacombe, Bellahsen, Amrouch and Daniel2014 b). The asymmetrical evolution of the past fluid pressure during contraction, decreasing in the west as it increased in the east, suggested a lateral continuity of the Ordovician–Permian reservoir during the LPS, with a pressure gradient favouring an eastward fluid migration On the basis of the estimated age of the fracturing events (see Beaudoin et alReference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a) the rate of the past fluid flow from the recharge area to the deformation site during LPS was estimated to be c. 8 km Myr−1.

6.c. The south Pyrenean FTB (Spain)

The south Pyrenean FTB is a wide south-verging FTB formed at the expense of extensional basins developed during Aptian to Cenomanian times (Sibuet et al. Reference Sibuet, Srivastava and Spakman2004) and inverted during Late Cretaceous times during the Pyrenean orogeny, itself related to the convergence between the Iberian plate and the European plate (Choukroune, Reference Choukroune1992). In the south Pyrenean FTB, the deformation is mainly distributed in the sedimentary cover along south-verging thrusts developed in a piggyback sequence and soling in the Triassic and Eocene evaporite décollement levels (Fig. 9). In the southernmost part of the Axial Zone, thrusts such as the Bielsa thrust or the Gavarnie thrust involve the basement (Fig. 9). The wealth of studies aiming at understanding the local or regional past fluid system makes this belt one of the most studied in the world (McCaig, Reference McCaig1988; Henderson & McCaig, Reference Henderson and McCaig1996; Travé et al. Reference Travé, Labaume, Calvet and Soler1997, Reference Travé, Calvet, Sans, Verges and Thirlwall2000, Reference Travé, Labaume and Vergés2007; McCaig et al. Reference McCaig, Tritlla and Banks2000; Caja et al. Reference Caja, Permanyer, Marfil, Al-Aasm and Martín-Crespo2006; Lacroix et al. Reference Lacroix, Buatier, Labaume, Trave, Dubois, Charpentier, Ventalon and Convert-Gaubier2011, Reference Lacroix, Leclère, Buatier and Fabbri2013, Reference Lacroix, Trave, Buatier, Labaume, Vennemann and Dubois2014, Reference Lacroix, Baumgartner, Bouvier, Kempton and Vennemann2018; Beaudoin et al. Reference Beaudoin, Huyghe, Bellahsen, Lacombe, Emmanuel, Mouthereau and Ouanhnon2015; Cruset et al. Reference Cruset, Cantarero, Travé, Vergés and John2016, Reference Cruset, Cantarero, Vergés, John, Muñoz-López and Travé2018, Reference Cruset, Vergés, Benedicto, Gomez-Rivas, Cantarero, John and Travé2021; Crognier et al. Reference Crognier, Hoareau, Aubourg, Dubois, Lacroix, Branellec, Callot and Vennemann2018; Nardini et al. Reference Nardini, Muñoz-López, Cruset, Cantarero, Martín-Martín, Benedicto, Gomez-Rivas, John and Travé2019; Hoareau et al. Reference Hoareau, Crognier, Lacroix, Aubourg, Roberts, Niemi, Branellec, Beaudoin and Suarez Ruiz2021 a; Muñoz-López et al. Reference Muñoz-López, Cruset, Vergés, Cantarero, Benedicto, Mangenot, Albert, Gerdes, Beranoaguirre and Travé2022). Recent syntheses (Lacroix et al. Reference Lacroix, Buatier, Labaume, Trave, Dubois, Charpentier, Ventalon and Convert-Gaubier2011; Cruset et al. Reference Cruset, Cantarero, Vergés, John, Muñoz-López and Travé2018; Crognier et al. Reference Crognier, Hoareau, Aubourg, Dubois, Lacroix, Branellec, Callot and Vennemann2018) on the past fluid system during the evolution of the FTB agreed on the general scenario (Fig. 9) involving a fluid system dominated by formational seawater where metamorphic fluids flowed from the Axial Zone along the basement-rooted thrusts and mixed to trigger cement precipitation in fault zones. A gradual opening of the system to meteoric fluids is also characterized during the late stage of thrust activity (e.g. Nardini et al. Reference Nardini, Muñoz-López, Cruset, Cantarero, Martín-Martín, Benedicto, Gomez-Rivas, John and Travé2019; Cruset et al. Reference Cruset, Vergés, Benedicto, Gomez-Rivas, Cantarero, John and Travé2021; Hoareau et al. Reference Hoareau, Crognier, Lacroix, Aubourg, Roberts, Niemi, Branellec, Beaudoin and Suarez Ruiz2021 a; Muñoz-López et al. Reference Muñoz-López, Cruset, Vergés, Cantarero, Benedicto, Mangenot, Albert, Gerdes, Beranoaguirre and Travé2022) and, at fold scale, in fold-growth-related fractures (Beaudoin et al. Reference Beaudoin, Huyghe, Bellahsen, Lacombe, Emmanuel, Mouthereau and Ouanhnon2015). The implication of extra-reservoir fluids has been proposed on the basis of the difference between δ18O from cements in fault zones and δ18O from local host rocks (Lacroix et al. Reference Lacroix, Trave, Buatier, Labaume, Vennemann and Dubois2014). Indeed, the vein – host-rock difference in δ18O is much higher (10 ‰) in cements precipitated from metamorphic fluids than in cements precipitated from formational or meteoric fluids (1 ‰ and 4 ‰, respectively). Interestingly, the nature of the past fluids that migrated along the thrusts varies in space. For instance, the Gavarnie thrust enabled metamorphic fluids from the Axial Zone to flow into the foreland (Lacroix et al. Reference Lacroix, Buatier, Labaume, Trave, Dubois, Charpentier, Ventalon and Convert-Gaubier2011), while others such as the Bielsa thrust or the Boixols thrust do not host any geochemical imprint of deep-sourced fluids (e.g. Nardini et al. Reference Nardini, Muñoz-López, Cruset, Cantarero, Martín-Martín, Benedicto, Gomez-Rivas, John and Travé2019). This difference in the nature of the fluid that flowed along the thrusts has been interpreted by Lacroix et al. (Reference Lacroix, Trave, Buatier, Labaume, Vennemann and Dubois2014, Reference Lacroix, Baumgartner, Bouvier, Kempton and Vennemann2018) as resulting from the location and timing of thrust activation: either the metamorphic fluids could not be efficiently transmitted by the thrusts that progressively formed at increasing distance from the feeding Axial Zone, and/or the sequence of basement thrusting did not coincide with the metasomatism releasing the metamorphic fluids.

6.d. The Umbria Marche Apennine Ridge (Italy)

The Apennines developed from the end of the Cretaceous to the Early Pleistocene in response to the convergence between the African and European plates. The resulting contractional deformation migrated eastwards following the retreat of the subducting Adriatic Plate under the European Plate since the Neogene (Elter et al. Reference Elter, Grasso, Parotto and Vezzani2008). Post-orogenic extension affected the western part of the belt first, before propagating eastward, being active nowadays in the Umbria and Abruzzi areas. The recent/modern fluid system in relation to the active extensional tectonics in the westernmost part of the belt has been extensively studied by characterizing either fluid emanating from, or recent cementation within, active normal fault zones (Ghisetti & Vezzani, Reference Ghisetti and Vezzani2000; Minissale et al. Reference Minissale, Magro, Martinelli, Vaselli and Tassi2000; Cello et al. Reference Cello, Invernizzi, Mazzoli and Tondi2001; Capozzi & Picotti, Reference Capozzi and Picotti2002; Agosta & Kirschner, Reference Agosta and Kirschner2003; Accaino et al. Reference Accaino, Bratus, Conti, Fontana and Tinivella2007; Smeraglia et al. Reference Smeraglia, Bernasconi, Berra, Billi, Boschi, Caracausi, Carminati, Castorina, Doglioni, Italiano, Rizzo, Uysal and Zhao2018, Reference Smeraglia, Giuffrida, Grimaldi, Pullen, La Bruna, Billi and Agosta2021; Lucca et al. Reference Lucca, Storti, Balsamo, Clemenzi, Fondriest, Burgess and Di Toro2019; Curzi et al. Reference Curzi, Aldega, Bernasconi, Berra, Billi, Boschi, Franchini, van der Lelij, Viola and Carminati2020; Coppola et al. Reference Coppola, Correale, Barberio, Billi, Cavallo, Fondriest, Nazzari, Paonita, Romano, Stagno, Viti and Vona2021; Vignaroli et al. Reference Vignaroli, Rossetti, Petracchini, Argante, Bernasconi, Brilli, Giustini, Yu, Shen and Soligo2022). The past fluid system associated with the development of the Apennine FTB and foreland basins has also received attention (Ghisetti et al. Reference Ghisetti, Kirschner, Vezzani and Agosta2001; Vannucchi et al. Reference Vannucchi, Remitti, Bettelli, Boschi and Dallai2010; Gabellone et al. Reference Gabellone, Gasparrini, Iannace, Invernizzi, Mazzoli and D’Antonio2013; Beaudoin et al. Reference Beaudoin, Labeur, Lacombe, Koehn, Billi, Hoareau, Boyce, John, Marchegiano, Roberts, Millar, Claverie, Pecheyran and Callot2020 c;   Curzi et al. Reference Curzi, Aldega, Bernasconi, Berra, Billi, Boschi, Franchini, van der Lelij, Viola and Carminati2020; Smeraglia et al. Reference Smeraglia, Aldega, Bernasconi, Billi, Boschi, Caracausi, Carminati, Franchini, Rizzo, Rossetti and Vignaroli2020 b; Labeur et al. Reference Labeur, Beaudoin, Lacombe, Emmanuel, Petracchini, Daëron, Klimowicz and Callot2021). Namely, the Umbria Marche Apennine Ridge (central–northern Apennines; Fig. 10) was studied by Ghisetti et al. (Reference Ghisetti, Kirschner, Vezzani and Agosta2001) and Beaudoin et al. (Reference Beaudoin, Labeur, Lacombe, Koehn, Billi, Hoareau, Boyce, John, Marchegiano, Roberts, Millar, Claverie, Pecheyran and Callot2020 c) along W–E transects. Ghisetti et al. (Reference Ghisetti, Kirschner, Vezzani and Agosta2001) reported stable isotopic values of calcite cements associated to thrusts along a WSW–ENE transect in the Central Apennines. The range of isotopic values of the cements (δ13C = [−1 to 3] ‰ PDB and δ18O = [26 to 33] ‰ SMOW) was equal to the range of isotopic values of the sedimentary host rocks. This was interpreted as a hint to consider that these thrusts were not allowing deep or superficial fluids to migrate into the studied formations, picturing a closed fluid system limited by the faults. To the north, in another portion of the Central Apennines, the Umbria Marche Apennine Ridge (UMAR; Fig. 10), Beaudoin et al. (Reference Beaudoin, Labeur, Lacombe, Koehn, Billi, Hoareau, Boyce, John, Marchegiano, Roberts, Millar, Claverie, Pecheyran and Callot2020 c) carried out analyses of fracture networks in several anticlines along a transect running from the front of the Tuscan nappes to the Adriatic coast. The associated cements were analysed, and the values of the δ18O in the veins were found to exceed the values of the δ18O of the Jurassic–Palaeocene sedimentary host rocks. The Δ47CO2 analysis revealed that the δ18O of the fluid was positive (3 to 15 ‰ SMOW), suggesting either a deep-sourced fluid or a brine evolved from seawater. The former interpretation was discarded on account of the 87Sr/86Sr values, too low to relate to basement-derived fluids. Thus, Beaudoin et al. (Reference Beaudoin, Labeur, Lacombe, Koehn, Billi, Hoareau, Boyce, John, Marchegiano, Roberts, Millar, Claverie, Pecheyran and Callot2020 c) proposed a lateral fluid flow of evolved brines in a stratigraphically compartmentalized reservoir (Fig. 10). This interpretation is further supported by the fact that in the innermost part of the belt, the hydrothermal regime of the fluids flowing in faults and veins (>140 °C) can be explained by a burial depth resulting from the cumulated thicknesses of Tuscan and Ligurian nappes, thrusted over the westernmost part of the UMAR at that time, if considering the past geothermal gradient of 24°C km−1 as indicated by vitrinite reflectance (Caricchi et al. Reference Caricchi, Aldega and Corrado2014). Note that a few veins related to LSFT host cements that precipitated from meteoric fluids characterized by a negative δ18O value. To summarize, the UMAR is overall interpreted as a closed fluid system, with lateral migration of fluids related to the development of the FTB and nappes (Fig. 10).

7. Overview of other case studies

Section 6 reports detailed examples of how the fluid system related to the development of FTB/BF can be unravelled using geochemical proxies. It built on areas where an abundant literature was available. We hereinafter complete these examples by reporting in a more succinct way the other FTB where a less significant number of geochemical studies have been conducted to depict the past fluid system with respect to the tectonic evolution at large scale or at mesoscale.

7.a. Thin-skinned fold-and-thrust belts: the Mexican Cordillera, the Appalachians and the Oman Mountains

The Mexican FTB is a thin-skinned FTB developed above a siliciclastic frictional décollement level (e.g. Fig. 11b) and where the past fluid system has been studied by Fitz-Diaz et al. (Reference Fitz-Diaz, Hudleston, Siebenaller, Kirschner, Camprubí, Tolson and Puig2011 a). Although most vein cements from folds return δ18O and δ13C values buffered by the host rocks, δ2H values of clays and fluid inclusions in veins are much higher in the undeformed foreland (up to 0 ‰ SMOW) than in the FTB (c. −60 ‰ SMOW). This spectacular increase supports a strong influence of meteoric fluids flowing during deformation near the hinterland, which fades away toward the foreland. Authors proposed that this isotopic pattern can be explained by a recharge zone in the west, where the higher part of the range is located, with a downward migration triggered by seismic pumping (Henderson & McCaig, Reference Henderson and McCaig1996)

Fig. 11. Schematic cross-sections in the fold-and-thrust belts with respect to the structural style, inherited structures and rheology of the sediments. The expected order of structural development (thrust activation) is shown as numbers, along with the expected fluid migration at that scale (blue arrows).

In the Sierra Madre (Fitz-Diaz et al. Reference Fitz-Diaz, Hudleston, Tolson, Poblet and Lisle2011 b), the fluid system related to the detachment fold complex developed over thick Triassic evaporites (Nuncio fold complex) has been studied using isotope geochemistry and fluid inclusion microthermometry in calcite and quartz veins that developed in sequence from the burial until the syn-folding curvature (Fischer & Jackson, Reference Fischer and Jackson1999; Lefticariu et al. Reference Lefticariu, Perry, Fischer and Banner2005; Fischer et al. Reference Fischer, Higuera-Díaz, Evans, Perry and Lefticariu2009). The fluid system was stratified and closed until folding, when it became homogenized and dominated by meteoric fluids. Eastwards, in the Cordoba platform and Veracruz Basin, i.e. the undeformed foreland, stable isotopes and fluid inclusion microthermometry were applied to cemented veins formed during LPS (Ferket et al. Reference Ferket, Roure, Swennen and Ortuño2000, Reference Ferket, Swennen, Ortuño and Roure2003, Reference Ferket, Swennen, Ortuño Arzate and Roure2006, Reference Ferket, Guilhaumou, Roure and Swennen2011; Gonzalez et al. Reference Gonzalez, Ferket, Callot, Guillhaumou, Ortuno and Roure2012). The reconstructed fluid system involved a lateral migration of saline brines mixed with meteoric waters, with a flow rate of 50–100 m Myr−1 at the deformation front down to 0.1–2 m Myr−1 further from it (Gonzalez et al. Reference Gonzalez, Ferket, Callot, Guillhaumou, Ortuno and Roure2012).

The thin-skinned Appalachian FTB developed from the Ordovician to Permian times in the eastern part of the USA. Large-scale duplexes developed above Ordovician and Cambrian frictional clay levels (Fig 11b). Significant effort has been devoted to reconstructing the past fluid system at the scale of the entire FTB, especially in the Cambrian to Silurian Valley Ridge Province (Deloule & Turcotte, Reference Deloule and Turcotte1989; Evans & Battle, Reference Evans and Battle1999; Ramsey & Onasch, Reference Ramsey and Onasch1999; Kirkwood et al. Reference Kirkwood, Ayt-Ougougdal, Gayot, Beaudoin and Pironon2000; Evans & Hobbs, Reference Evans and Hobbs2003; Evans, Reference Evans and Tollo2010; Evans et al. Reference Evans, Bebout and Brown2012). Fluid inclusion microthermometry and C, O isotopic studies helped define the P-T-X characteristics of the fluid system during deformation (Evans & Battles, Reference Evans and Battle1999; Evans & Hobbs, Reference Evans and Hobbs2003; Evans, Reference Evans and Tollo2010; Evans et al. Reference Evans, Bebout and Brown2012). During LPS, the past fluid system appears to be compartmentalized, with distinct reservoirs overlying each other. During folding, an efficient vertical migration and thus a breakage of horizontal seals occurred (Evans, Reference Evans and Tollo2010; Evans et al. Reference Evans, Bebout and Brown2012), with local thermal and isotopic disequilibrium with the host rock (Evans & Battle, Reference Evans and Battle1999; Evans & Hobbs, Reference Evans and Hobbs2003). These geochemical signatures hint at the input of extra-reservoir fluids into the reservoir rocks. However, no distinction can be made concerning the nature of the fluids, possibly being either meteoric or metamorphic. Yet, authors proposed that the topographic high to the west served as a recharge zone for these fluids that would have flowed as a squeegee over 60 km across the belt (Evans & Battle, Reference Evans and Battle1999). Interestingly, not all the thrusts acted as drains to propel the fluids forelandwards: some of them acted as barriers restricting the fluid system according to their orientation and/or degree of reactivation (Ramsey & Onasch, Reference Ramsey and Onasch1999).

The Oman Mountains, in the United Arab Emirates, is a thin-skinned FTB in which the fluid system was characterized by studying hydrothermal dolomitization and fracture cements (Breesch et al. Reference Breesch, Swennen and Vincent2006, Reference Breesch, Swennen and Vincent2009, Reference Breesch, Swennen, Vincent, Ellison and Dewever2010, Reference Breesch, Swennen, Dewever, Roure and Vincent2011; Holland et al. Reference Holland, Saxena and Urai2009a, b; Holland & Urai, Reference Holland and Urai2010; Arndt et al. Reference Arndt, Virgo, Cox and Urai2014). Authors discriminated a burial phase, a pre-LPS phase, a syn-folding phase and a post-deformation phase. During the burial phase, crack-seal cements precipitated from rock-buffered fluids, possibly formational, an interpretation also proposed by Hilgers et al. (Reference Hilgers, Kirschner, Breton and Urai2006). The fracture cements in the LPS phase were interpreted as reflecting temperature increase during burial diagenesis. During contractional deformation including LPS, the fluid system was dominated by hydrothermal brines channelized from the basal décollement along a network of connected thrusts. Fluid flow simulations suggest hot flashes along the thrusts, the duration of which depends on the thrust activity (Callot et al. Reference Callot, Breesch, Guilhaumou, Roure, Swennen and Vilasi2010 b).

7.b. Thin-skinned fold-and-thrust belts with salt tectonics: the Albanides and the Sivas Basin

The past fluid system in the Albanides Foothills, mainly the Ionian Basin, located in the southwest of Albania, was reconstructed by van Geet et al. (Reference van Geet, Swennen, Durmishi, Roure and Muchez2002), Vilasi et al. (Reference Vilasi, Swennen and Roure2006, Reference Vilasi, Malandain, Barrier, Callot, Amrouch, Guilhaumou, Lacombe, Muska, Roure and Swennen2009), Roure et al. (Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010) and de Graaf et al. (Reference de Graaf, Nooitgedacht, Goff, van der Lubbe, Vonhof and Reijmer2019). This former Jurassic–Oligocene intracontinental rift basin was inverted during the Alpine orogeny, strongly mobilizing the Triassic evaporites as the main décollement with some inherited (and rejuvenated as well) diapirism, making this mountain range a thin-skinned FTB with salt tectonics influence (e.g. Fig. 11e). For van Geet et al. (Reference van Geet, Swennen, Durmishi, Roure and Muchez2002), isotopic geochemistry on the fracture cements across the belt depicts a rather buffered fluid system during burial and LPS, with an enrichment in δ18O during thrust activation that authors relate to interaction with Triassic evaporites. A similar enrichment is observed by Vilasi et al. (Reference Vilasi, Swennen and Roure2006, Reference Vilasi, Malandain, Barrier, Callot, Amrouch, Guilhaumou, Lacombe, Muska, Roure and Swennen2009) coevally with the deposition of barite, strengthening the idea of interactions with the Triassic evaporites. Vilasi et al. (Reference Vilasi, Swennen and Roure2006, Reference Vilasi, Malandain, Barrier, Callot, Amrouch, Guilhaumou, Lacombe, Muska, Roure and Swennen2009) also characterized a complex pre-deformational fluid system with input of meteoric fluids flowing through soil. The input of either meteoric or ‘hydrothermal’ fluids suggests an opening of the fluid system during contraction. De Graaf et al. (Reference de Graaf, Nooitgedacht, Goff, van der Lubbe, Vonhof and Reijmer2019) proposed a slightly different scenario, with fluid interacting with evaporites since LPS in a reservoir that opened to meteoric fluid flow during the folding phase.

The Sivas Basin, located in the Anatolian plateau in Turkey, is an Oligocene to Pliocene foreland basin developed between the Taurides–Anatolides block and the Pontides block. Recent structural analyses have established the Sivas Basin as a striking example of a synorogenic salt tectonics-controlled basin (Fig. 11e), where the geometry at the time of deformation was directly controlled by the distribution and deformation of the Late Eocene evaporites of the Tuzhisar Fm (Kergaravat et al. Reference Kergaravat, Ribes, Legeay, Callot, Kavak and Ringenbach2016; Legeay et al. Reference Legeay, Ringenbach, Kergaravat, Pichat, Mohn, Vergés, Sevki Kavak and Callot2020). The influence of synorogenic salt tectonics can be illustrated from the geochemical characterization of the surrounding sandstones (Pichat et al. Reference Pichat, Hoareau, Callot and Ringenbach2016) and of the evaporites collected around diapiric structures (Pichat et al. Reference Pichat, Hoareau, Callot, Legeay, Kavak, Révillon, Parat and Ringenbach2018). Salt tectonics enabled the complete clogging of sandstone porosity by precipitation of calcite and analcime in the vicinity of the diapiric structures early in its diagenetic evolution. Also, the Sr isotopic study of the evaporites shows that the saline waters currently at the surface yield the same radiogenic signature as the evaporites from the Tuzhisar Fm (Pichat et al. Reference Pichat, Hoareau, Callot, Legeay, Kavak, Révillon, Parat and Ringenbach2018), which suggests that current saline fluids are polluted by the Sr isotopic ratio of the salt through continuous diapiric recycling. This is a point to consider in basins where salt tectonics is present: as diapirs are dynamic structures, they may carry with them the radiogenic signature of their formation through mechanical advection or through secondary evaporite deposition. Consequently, any fluids flowing in contact with evaporites will presumably acquire their radiogenic signature, limiting the use of Sr-based tracers (87Sr/86Sr) or geochronometry (Rb–Sr) to infer the source and pathway of the fluids that interacted with the evaporites.

7.c. Fold-and-thrust belts with superimposed thin-skinned and thick-skinned styles of deformation: the Zagros FTB

Resulting from the ongoing collision of the Arabian plate with the Eurasian plate, the Zagros FTB in Kurdistan and Iran shows superimposed thin-skinned and thick-skinned tectonic styles of deformation (Fig. 11d; Ahmadhadi et al. Reference Ahmadhadi, Lacombe, Daniel, Lacombe, Lavé, Vergés and Roure2007; Mouthereau et al. Reference Mouthereau, Tensi, Bellahsen, Lacombe, de Boisgrollier and Kargar2007, Reference Mouthereau, Lacombe and Vergés2012; Lacombe & Bellahsen, Reference Lacombe and Bellahsen2016). The regional fluid system deduced from the study of the folds of the Zagros FTB was assessed using isotopic geochemistry and fluid inclusion microthermometry from mesoscale fracture networks from several major folds by Kareem et al. (Reference Kareem, Al-Aasm and Mansurbeg2019). Authors depicted an early pervasive forelandward flow of hydrothermal brines, related to extensive dolomitization during LPS, followed by high flux of hydrothermal brines leading to another dolomitization event focused around the thrusts during their activation.

7.d. General evolutionary trend of the past fluid system at the scale of the orogenic foreland

The evolution of the fluid system as reported in the above case studies seemingly exhibits a common first-order pattern, independent of the structural style (Fig. 12). Indeed, as the pre-contraction strata now incorporated into the FTB likely were formerly parts of the undeformed foreland, they underwent (1) burial, (2) potentially extension related to the flexure/forebulge development, and (3) LPS. During burial, a closed fluid system is pictured, with the implication of local fluids (among others, van Geet et al. Reference van Geet, Swennen, Durmishi, Roure and Muchez2002; Hilgers et al. Reference Hilgers, Kirschner, Breton and Urai2006; Vandeginste et al. Reference Vandeginste, Swennen, Allaeys, Ellam, Osadetz and Roure2012; Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a). However, the fluid flowing in fractures formed during forebulge development is usually not well documented. Indeed, unambiguously relating veins to outer-arc extension related to large-scale forebulge development requires careful observation of the chronological relationship between fractures and of their spatial distribution. Complementary microstructural study with techniques such as palaeostress reconstructions from calcite twins can support the timing of development of these fractures (e.g. Beaudoin et al. Reference Beaudoin, Leprêtre, Bellahsen, Lacombe, Amrouch, Callot, Emmanuel and Daniel2012). Consequently the paucity of published data limits our ability to propose a common trend for the fluid system related to the forebulge development, yet some authors suggest a transient vertical connection between reservoirs at that time, similar to that observed during folding (Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a). The systematic occurrence of such connection is still to be proven in other case studies, implying the need to properly decipher the fractures related to the forebulge development when studying the fracture network (Tavani et al. Reference Tavani, Storti, Lacombe, Corradetti, Muñoz and Mazzoli2015). In contrast, the past fluid system bound to the LPS has been widely documented in all case studies, especially because LPS propagates deformation over hundreds of kilometres far into the stable craton (Craddock & Relle, Reference Craddock and Relle2003; van der Pluijm et al. Reference van der Pluijm, Craddock, Graham and Harris1997; Lacombe & Mouthereau, Reference Lacombe and Mouthereau1999; Lacombe, Reference Lacombe2010; Beaudoin & Lacombe, Reference Beaudoin and Lacombe2018). During LPS, the mesostructures allow for a compartmentalized fluid system with lateral migration over kilometric distances, maintained by topographically driven or squeegee fluid flow. In contrast, almost all studies document a transient and strong enhancement of the vertical fracture permeability during thrusting and folding, leading to local opening to extra-reservoir fluids, either meteoric or basement-derived, that flow in fold hinges and along fault zones.

Fig. 12. Sketch illustrating the expected first-order evolution of the fluid system in FTB, without considering specificities related to structural style. (a) During layer-parallel shortening; (b) during folding/thrusting.

In spite of displaying a common evolution for the past fluid system, the timing and nature of fluids involved during thrusting and folding are strongly controlled by the structural style (Fig. 11). The impact of structural style at that time is twofold. On the one hand, the structural style imposes the nature of the fluids that can fill in the reservoir in the sedimentary rocks: for instance, basement fluids are only encountered in forelands where the basement is involved in shortening. On the other hand, the structural style impacts the fluid flow by imposing a sequence of thrusting (Fig. 11). Thin-skinned tectonic wedges are usually associated with a sequence thrust activation and migration of the deformation front. In that case, the migration of fluids at the scale of the FTB occurs forelandwards (Fig. 11a), as documented in the Canadian Rocky Mountains (Vandeginste et al. Reference Vandeginste, Swennen, Allaeys, Ellam, Osadetz and Roure2012). In some cases, howewer, the development of thrusts can be out-of-sequence, with some inner thrusts developing later than the outermost ones. This occurred in the southern Pyrenees (Henderson & McCaig, Reference Henderson and McCaig1996; McCaig et al. Reference McCaig, Tritlla and Banks2000; Lacroix et al. Reference Lacroix, Trave, Buatier, Labaume, Vennemann and Dubois2014), where the fluids flowing along the thrusts are different in spite of the latter being of similar spatial extension (Fig. 9). Thick-skinned systems often display a more irregular and erratic sequence of deformation, especially in broken forelands. The sequence affects the timing of the fluid flowing from the basement into the sedimentary cover (Beaudoin et al. Reference Beaudoin, Lacombe, Roberts and Koehn2018). Finally, in FTB where the tectonics is linked to salt dynamics, the chemical signature of the fluids is strongly controlled by where and when diapiric structures develop and are dynamically involved in the shortening history (Pichat et al. Reference Pichat, Hoareau, Callot, Legeay, Kavak, Révillon, Parat and Ringenbach2018).

8. Discussion: towards geochemistry-assisted structural geology

8.a. Main feedbacks of past fluid system studies on structural geology

The very concept of geochemistry-assisted structural geology is to use the geochemical signatures of the past fluid to constrain subsurface geology via the understanding of fluid sources and conduits, with quantitative information about the temperature and age of the fluid from which the calcite cement precipitated. The overview of the case studies reveals a relationship between the fluid system evolution and tectonics that can be established by linking the opening of the fluid system in a reservoir to the structural style of deformation and/or the local or regional tectonic evolution. In the presented case studies, the interpretation of the past fluid system in terms of sources and pathways directly confronts the previous understanding of (1) the structure with an extended knowledge about the subsurface geometry thanks to wells or seismic imagery, and (2) deformation timing constrained by sedimentological, geochronological and structural observations and analyses. In turn, when the past fluid system is studied in order to unravel the spatial and temporal distribution of conduits and, by extension, the associated tectonic history, it appears possible to further constrain the structural geology whether at the scale of the individual structure or at the scale of the FTB. To do so and to be as unambiguous as possible in the interpretation of the geochemistry of calcite cements (e.g. Muñoz-López et al. Reference Muñoz-López, Cruset, Vergés, Cantarero, Benedicto, Mangenot, Albert, Gerdes, Beranoaguirre and Travé2022), we propose that the geochemical characteristics of the past fluid system should be built on (1) the interpretation of the difference in δ18O and δ13C between the vein cements and the local host rocks in terms of the degree of opening of the fluid system; and (2) a combination of δ18OCaCO3 values of the cements with independent information on the temperature of the fluid at the time the cements precipitated as provided by another palaeothermometer (fluid inclusion microthermometry, Δ47CO2). Such a combination enables the characterization of the nature of the fluid (meteoric, seawater or sedimentary). In some cases, other isotopic (4He or δ2H) analyses further help discriminate the fluid nature (metamorphic, mantellic, meteoric); (3) the use of tracers like 87Sr/86Sr or trace elementary composition, that help infer the specific lithologies through which the fluid has migrated (crystalline basement, evaporites, clays), constraining the migration pathways and the extension of the fracture network; and (4) age dating of the cements by means of geochronology, more specifically the uranium-based geochronology.

A direct instance comes from the Sheep Mountain Anticline (Laramide province), a fold developed above a blind basement-fault whose geometry at depth is unknown and still debated (see Bellahsen et al. Reference Bellahsen, Fiore and Pollard2006 a). By reconstructing the past fluid system using isotopes and fluid inclusion microthermometry, Beaudoin et al. (Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011) unravelled a two-step fluid system, compartmentalized during LPS and homogenized during fold growth. The latter comprises a strong overprint of the seawater and evolved seawater by meteoric-derived fluids flowing from the basement, likely channelized along the underlying high-angle thrust and precipitating in a hydrothermal regime. The distribution in map-view of the geochemical signature of the cements precipitated from these hydrothermal fluids, which display a typical and homogeneous range of δ18O values (from −19 ‰ PDB to −23 ‰ PDB; Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011, Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a; Fig. 13a), shows that the hydrothermal pulse was focused in a zone in the backlimb striking parallel to the current position of the hinge. The authors interpreted the distribution and orientation of this zone as a proxy to locate the fault tip at the time of folding, and proposed a dynamic model of hinge migration during fold evolution (Fig. 13).

Fig. 13. Example of direct application of geochemistry-assisted structural geology, based on the case study of the Sheep Mountain Anticline (SMA; Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011). (a) Geological map of the SMA, where the values of the δ18O of tectonic veins are reported as coloured points, and where iso-δ18O value curves are plotted. (b) Simplified δ18O vs δ13C plots of the cements in the fracture network. The colours of the dots and the corresponding frames relate to the deformation phase during which the vein developed. Red and orange vertical lines correspond to the past fluid system interpretation, in accordance with the curves plotted in (a). (c) Cross-section along the line located on (a), with the corresponding location of the iso-δ18O value curves, with a proposed subsurface geometry. (d) Interpretation of the deformation sequence accounting for a potential migration of the hinge during fold growth. (a) and (b) are modified after Beaudoin et al. (Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011).

On the basis of the past fluid system reconstruction from numerous structures in the southern Pyrenees, Cruset et al. (2017) proposed to use the evolution of the chemical signatures of the successive calcite cements precipitated in a thrust zone to infer whether the thrust (and the related fold) is detached through evaporites or clays. They observe two geochemical trends, in which the fluids are always a mixture of deep-sourced fluids and meteoric fluids. Where evaporites are involved in the décollement, the sequence of mineralization in the thrust zone shows successive calcite cements precipitated from a fluid with a gradually increasing δ18O value with respect to the fluid from which the previous cement precipitated. Where the thrust is not soling into evaporites, the sequence of mineralization in the thrust zone shows successive calcite cements precipitated from a fluid with a decreasing δ18O value and decreasing Fe and Sr contents with respect to the fluid from which the previous cement precipitated. If this rule looks very promising and could pave the way to a geochemical imagery of the nature of décollement based on fault/fracture calcite cements, it needs more applications to validate it further and to clarify the mechanisms that can explain this geochemical pattern.

Another striking example of geochemistry-assisted structural geology comes from past fluid system reconstruction in the southern Pyrenees (Lacroix et al. Reference Lacroix, Buatier, Labaume, Trave, Dubois, Charpentier, Ventalon and Convert-Gaubier2011). Coupling δ18O δ13C of cements in veins and fault zones to fluid inclusion microthermometry allowed those authors to show that the fluid system was closed to external fluids. Doing so, it is possible to estimate the depth at which a thrust was active, assuming a given geothermal gradient. That resulted in the picture of how the thrust evolved in the southern Pyrenees (Lacroix et al. Reference Lacroix, Buatier, Labaume, Trave, Dubois, Charpentier, Ventalon and Convert-Gaubier2011) or, when coupled to U–Pb absolute dating, in the reconstruction of the last stage of activity of the southernmost Pyrenean Thrust (Hoareau et al. 2021). By coupling the Δ47CO2 temperature of cement precipitation in a closed system to a 1D burial-time model, a similar approach was also used in the Cingoli Anticline (Apennines, Italy) by Labeur et al. (Reference Labeur, Beaudoin, Lacombe, Emmanuel, Petracchini, Daëron, Klimowicz and Callot2021) to estimate the absolute timing of the different stages of contractional deformation (LPS, folding, LSFT), in good agreement with the age of folding as constrained by growth strata.

Geochemistry-assisted structural geology is also a relevant approach to refine the large-scale structural geology in FTB where the structural style is debated, or has evolved over time. For example, in the Umbria Marches Apennine Ridge, Labeur et al. (Reference Labeur, Beaudoin, Lacombe, Emmanuel, Petracchini, Daëron, Klimowicz and Callot2021) and Beaudoin et al. (Reference Beaudoin, Labeur, Lacombe, Koehn, Billi, Hoareau, Boyce, John, Marchegiano, Roberts, Millar, Claverie, Pecheyran and Callot2020 c) reported that the main carbonate reservoir of the FTB hosted seawater fluids, i.e. with trace of neither evaporitic saline brines nor basement-derived fluids. In this FTB, where the degree of mechanical coupling between the basement and the sedimentary cover is strongly debated (Calamita et al. Reference Calamita, Satolli, Scisciani, Esestime and Pace2011; Scisciani et al. Reference Scisciani, Agostini, Calamita, Pace, Cilli, Giori and Paltrinieri2014, Reference Scisciani, Patruno, Tavarnelli, Calamita, Pace and Iacopini2019), especially as thick evaporites lie in between, such an absence of extra-reservoir fluids is insightful. It can mean either that (1) thrusts in relation to which the folds were built are not rooting deep enough to reach either evaporites or basement, a hypothesis discarded by seismic imagery, or (2) the thrusts more likely root in the basement rather than in the evaporites. Authors favoured this interpretation, relating the absence of any basement-derived fluid signature to the fact that the impervious Triassic evaporites acted as a major barrier and prevented any upward fluid flow from the basement even along the (presumably limited) damage zone of the basement-rooted thrusts.

8.b. Timescales of deformation and fluid systems in orogenic forelands

An obvious side effect of understanding how fractures are related to the tectonic evolution of an area is that it grants a direct way of dating the tectonic event (see Roberts & Holdsworth, Reference Roberts and Holdsworth2022, for a recent review). Of course, it is mandatory to have the best understanding of the diagenesis history of a fracture-filling cement to be sure of what is actually dated, especially in fault zones where numerous dissolution–precipitation, recrystallization and/or successive fluid flow events can be recorded (e.g. Aubert et al. Reference Aubert, Léonide, Lamarche and Salardon2020). That being done, it becomes possible to date not only the timing of deformation, but also the duration of some events. For example, in the broken foreland of the Laramide province, U–Pb ages obtained in cements coeval with vein opening reveal that the Laramide LPS-related mesostructures developed over a duration of c. 30 Myr (Beaudoin et al. Reference Beaudoin, Lacombe, Roberts and Koehn2018, Reference Beaudoin, Lacombe, Roberts and Koehn2019a). This means that a tectonic deformation stage can last tens of millions of years, and in the case of the Laramide province, fractures from the LPS time are filled with fluids that flowed eastward at the scale of the basin (Barbier et al. Reference Barbier, Hamon, Callot, Floquet and Daniel2012 b; Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a). That means that during 30 Myr, the mesostructures developed in a way that maintained the required permeability of the sedimentary cover to enable the fluid migration at the basin scale in spite of cement precipitation in some of the fractures. Even the fractures related to strata curvature during fold growth have been estimated to develop over 1.5 to 15 Myr in various folds (Lacombe et al. Reference Lacombe, Beaudoin, Hoareau, Labeur, Pecheyran and Callot2021). Over these rather long timespans, the past fluid system remains isotopically homogeneous (see Section 5), with most of the curvature-related veins filled with cements precipitated either from basement-derived fluids (e.g. Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe and Emmanuel2011; Lacroix et al. Reference Lacroix, Trave, Buatier, Labaume, Vennemann and Dubois2014) or from meteoric fluids (e.g. Lynch et al. Reference Laubach, Eichhubl, Hilgers and Lander2021). An overlooked but major implication of involving external fluids of the same source and chemistry over Myrs is that there is a need for long-lasting (> few Myrs) active recharge and migration mechanisms, e.g. convection cells at the scale of the FTB and/or the foreland basin (Gomez-Rivas et al. Reference Gomez-Rivas, Corbella, Martín-Martín, Stafford, Teixell, Bons, Griera and Cardellach2014. That brings into question the dynamics of the fluids and its timescale in FTB, which have been extensively investigated by means of numerical simulations of the past fluid system, especially when it comes to hydrocarbon exploration (see Roure et al. Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010, for a review). It must be kept in mind that measuring a migration speed for a given episode of deformation is out of reach when studying the past fluid system. Indeed, the absolute dating of a cement yields not the age of the fluid, but the age of the precipitation of the material dissolved within this fluid, and the fluid from which the cement precipitated in a fracture is therefore no longer available to mineralize another fracture. Nevertheless, in cases where the deformation event and related fracture development occurred over a short and well-constrained timespan, it becomes possible to estimate an average migration speed (e.g. Beaudoin et al. Reference Beaudoin, Bellahsen, Lacombe, Emmanuel and Pironon2014 a). Beyond that limitation, being able to perform the direct dating of a cement that is related to a specific fluid flow (e.g. basement-derived fluids) is a promising way to build a calendar of when fluid conduits developed, that can be related, for example, to the timing of basement thrust activation/propagation into the cover rocks.

8.c. Future developments

We have seen in this review that reconstructing the past fluid system from a simple mineralogical system (i.e. calcite) enables a better understanding of the evolution of the fluid migration, of the connectivity of the conduits and ultimately of the structural geology and tectonics of an area. Considering the case studies presented, the fracture network is depicted as an important drain in the palaeohydrology at the fold and the orogenic foreland scales. As the fracture network records much more information about the deformation history than thrusts and other seismic-scale faults, that means that the syn-kinematic cements precipitated in the fracture network are good targets to reconstruct the fluid system and its evolution in relation to large-scale tectonics. In particular, the role of stylolites on the fluid system remains vastly overlooked in that context, except to constrain the diagenesis related to the burial phase or the early LPS (Swennen et al. Reference Swennen, Muskha and Roure2000; Roure et al. Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010). As recent studies support that stylolites can be efficient conduits for fluid flow beyond the local scale (Koehn et al. Reference Koehn, Rood, Beaudoin, Chung, Bons and Gomez-Rivas2016; Martín-Martín et al. Reference Martín-Martín, Gomez-Rivas, Gómez-Gras, Travé, Ameneiro, Koehn and Bons2018), and that they can develop during most of the folding stage (Beaudoin et al. Reference Beaudoin, Koehn, Lacombe, Lecouty, Billi, Aharonov and Parlangeau2016), it appears necessary to further investigate stylolites to reconstruct the past fluid system. This can be done simply by considering the tectonic and sedimentary stylolites, and by analysing cements precipitated in tension gashes or at the tip of stylolites (Aharonov & Karcz, Reference Aharonov and Karcz2019). On top of this, it is noteworthy that sedimentary stylolites yield quantitative estimates of the depth at which they developed, helping refine the expected temperature for a fluid flow event (Beaudoin et al. Reference Beaudoin, Gasparrini, David, Lacombe and Koehn2019 b, Reference Beaudoin, Lacombe, Koehn, David, Farrell and Healy2020 b). Another feature that deserves more attention to better appraise structural evolution and conduit distribution when reconstructing the past fluid system is the mineralogical transformation processes. As carbonates are extremely sensitive to fluid-mediated mineralogical transformation like dolomitization or apatitization of calcite, reconstructing the type of fluid that leads to the transformation can be insightful for the conduits. In the Maestrat Basin (Spain), such an approach revealed a deep-source fluid migrating over long timespans (Martin-Martin et al. Reference Martín-Martín, Gomez-Rivas, Gómez-Gras, Travé, Ameneiro, Koehn and Bons2018), which implies potential convection cells in the subsurface (Gomez-Rivas et al. Reference Gomez-Rivas, Corbella, Martín-Martín, Stafford, Teixell, Bons, Griera and Cardellach2014). Of course, such an approach is limited by our understanding of the fluid-mediated transformation process itself and the associated volume of fluids required. These are much debated to date (Davies & Smith, Reference Davies and Smith2006; Merino & Canals, Reference Merino and Canals2011; Jonas et al. Reference Jonas, John, King, Geisler and Putnis2014; Koeshidayatullah et al. Reference Koeshidayatullah, Corlett, Stacey, Swart, Boyce, Robertson, Whitaker and Hollis2020 a; Centrella et al. Reference Centrella, Beaudoin, Derluyn, Motte, Hoareau, Lanari, Piccoli, Pecheyran and Callot2021; Weber et al. Reference Weber, Cheshire, Bleuel, Mildner, Chang, Ievlev, Littrell, Ilavsky, Stack and Anovitz2021) and merit further investigation.

9. Conclusions

In this contribution, we provide a methodological and applicative review of published work focusing on the characterization of the past fluid system associated to orogenic forelands using geochemistry on fault and vein calcite cements. We also highlight/discuss whether and how past fluid system studies can shed light on tectonics. This review points out that some of the most commonly used geochemical tools like δ18O–δ13C of carbonates can be very misleading if used alone, and that complementary methods are required in order to properly define a past fluid system, i.e. reconstructing the source fluid chemistry to infer the potential pathways. We can separate the geochemical proxies into three categories: (1) the palaeothermo(baro)meters, such as Δ47CO2 and fluid inclusion microthermometry, which each have application domains limited by temperature (180 °C max for Δ47CO2, 50 °C min for FIM in large inclusions); (2) the tracers, like δ2H, 87Sr/86Sr, REE patterns, noble gas such as 4He, fluid inclusions and to a certain extent the trace element contents; and (3) the geochronometers, mainly U–Pb in carbonates. Of course, the fast-growing development of analytical capability, in both spatial and concentration resolution, has already made interpretation of the fluid chemistry less ambiguous and so provides a more reliable picture of the fluid system. One can be confident that ongoing analytical developments will democratize further the ability of the community to conduct convincing fluid system reconstructions in the future. For instance, extracting noble gases from fluid inclusions related to a given phase of deformation will enable more accessible ways to assess the source and chemistry of the fluids; lowering the detection threshold and improving the data treatment will allow dating of smaller volume by means of U–Pb geochronology; coupling the clumped isotopy to LA-ICP-MS will allow characterization of the fluid nature and temperature at a higher spatial resolution. Nevertheless, the currently existing range of proxies, when used together, seem enough to comprehend most of the past fluid systems.

In published studies, there is a common trend of past fluid system evolution at the scale of the fold, the thrust and of the orogenic foreland (Fig. 12): (1) the system is closed or compartmentalized before the onset of folding/thrusting; (2) during folding/thrusting, the system opens to external fluids, either meteoric fluids, deep-sourced brines or metamorphic fluids according to what the structural geology allows for. However, there are variations according to the structural style of deformation in the belt and to the sedimentary succession of the area, with possibly no deep-sourced fluid in case the thrusts cross a layer of non-permeable material, like evaporites. The latter, however, always affect the fluid system of any thrust rooted in the evaporite layer. It is interesting to note that deciphering a past fluid system at the fold scale alone is always complicated by the fact that contractional structures in FTBs are mostly hydrologically connected with a strong influence of lateral migration during the tectonic history. Another important point to highlight is that even though large-scale fault zones remain an object of prime interest for reconstructing the past fluid system, numerous studies have established that the fracture network is of the utmost relevance to reconstruct the long-term past fluid system evolution. That supports that fractures below the seismic resolution are efficient drains and reliable recorders of the fluid system variation controlled by large-sale migration engines. The ambiguous impact of pressure solution on the fluid flow at the fracture-network scale remains to be better investigated in future studies.

The feedback that the characterization of the past fluid system grants to understand tectonics is twofold. (1) In cases where the subsurface is well known (i.e. seismic lines, wells), understanding the nature, temperature and preferential pathways of the fluid has been proven to be valuable for refining the structural-based basin models. This is mainly based on the fact the past fluid system helps better constrain the location of the drains and seals by reconstructing past pressure levels in reservoir, or to locate the extension of a thrust (refer to the numerous examples in the review of Roure et al. Reference Roure, Andriessen, Callot, Faure, Ferket, Gonzales, Guilhaumou, Lacombe, Malandain, Sassi, Schneider, Swennen, Vilasi, Goffey, Craig, Needham and Scott2010). In that case, the recent democratization of geochronology in carbonates will improve the constraints on the time frame of deformation, refining further the 4D basin modelling. (2) In cases where little information is available about the subsurface, or if the tectonics is debated, it seems that past fluid flow holds some keys to comprehension. The isotopic and elemental signature of the cements precipitated in fractures is likely to reflect the source of the mineralizing fluid, especially in cases where the crystalline basement or evaporite levels are involved in shortening. That can help discussion of the tectonic style of deformation of an orogenic foreland, even though one needs to keep in mind that the absence of, for instance, basement fluids does not mean the basement is not involved at all in the deformation (e.g. Jura Mountains). At the fold scale, the geochemical signature of cements can be used to locate the main drains (e.g. underlying fault zone, hinge), which paves the way to alternative ways to discuss the structural geology. Finally, the past fluid system related to major faults and fractures allows the reconstruction of the absolute age of the deformation. To date, constraining the fluid dynamics in an orogenic foreland using that approach is out of reach, as only the age of the cement precipitation, and not of the source fluid, can be obtained. Yet, absolute ages grant another constraint to discuss how deformation affects strata in forelands, and to put invaluable time constraints (timing, duration, rate) on dynamic events such as layer-parallel shortening, fold growth and thrust development.

We believe that geochemistry-based structural geology is a growing domain that will help refine predictive fluid flow models, beyond, obviously, the sole case of forelands, but the potential of which will also be developed, helping better understand the multiscale deformation history in the upper crust.

Acknowledgements

N.B. is funded through the isite-E2S, supported by the ANR PIA and the Région Nouvelle-Aquitaine. N.B. thanks S. Centrella and A. Battani for insightful discussions. We thank L Smeraglia, A Dielforder, S Mittempergher and an anonymous reviewer for their insightful comments and suggestions that led to significant improvement of the manuscript, together with Editor-in Chief P. Clift for his editorial work.

References

Accaino, F, Bratus, A, Conti, S, Fontana, D and Tinivella, U (2007) Fluid seepage in mud volcanoes of the northern Apennines: an integrated geophysical and geological study. Journal of Applied Geophysics 63, 90101.CrossRefGoogle Scholar
Affek, HP (2012) Clumped isotope paleothermometry: principles, applications, and challenges. The Paleontological Society Papers 18, 101–14.CrossRefGoogle Scholar
Affolter, S, Fleitmann, D and Leuenberger, M (2014) New online method for water isotope analysis of speleothem fluid inclusions using laser absorption spectroscopy (WS-CRDS). Climate of the Past 10, 1291–304.CrossRefGoogle Scholar
Agosta, F (2008) Fluid flow properties of basin-bounding normal faults in platform carbonates, Fucino Basin, central Italy. In The Geodynamics of the Aegean and Anatolia (eds T Taymaz, Y Yilmaz and Y Dilek), pp. 277–91. Geological Society of London, Special Publication no. 299.CrossRefGoogle Scholar
Agosta, F and Kirschner, DL (2003) Fluid conduits in carbonate-hosted seismogenic normal faults of central Italy. Journal of Geophysical Research: Solid Earth 108, 2221.CrossRefGoogle Scholar
Agosta, F, Luetkemeyer, PB, Lamarche, J, Crider, JG and Lacombe, O (2016) An introduction to the Special Issue on “The role of fluids in faulting and fracturing in carbonates and other upper crustal rocks”. Tectonophysics 690, 13.CrossRefGoogle Scholar
Aharonov, E and Karcz, Z (2019) How stylolite tips crack rocks. Journal of Structural Geology 118, 299307.CrossRefGoogle Scholar
Ahmadhadi, F, Lacombe, O and Daniel, JM (2007) Early reactivation of basement faults in Central Zagros (SW Iran): evidence from pre-folding fracture populations in the Asmari Formation and Lower Tertiary paleogeography. In Thrust Belts and Foreland Basins: From Fold Kinematics to Hydrocarbon Systems (eds Lacombe, O, Lavé, J, Vergés, J and Roure, F), pp. 205–28. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Al-Aasm, I, Lonnee, J and Clarke, J (2000) Multiple fluid flow events and the formation of saddle dolomite: examples from Middle Devonian carbonates of the Western Canada Sedimentary Basin. Journal of Geochemical Exploration 69–70, 11–5.CrossRefGoogle Scholar
Al-Aasm, IS, Lonnee, J and Clarke, J (2002) Multiple fluid flow events and the formation of saddle dolomite: case studies from the Middle Devonian of the Western Canada Sedimentary Basin. Marine and Petroleum Geology 19, 209–17.CrossRefGoogle Scholar
Amrouch, K, Beaudoin, N, Lacombe, O, Bellahsen, N and Daniel, J-M (2011) Paleostress magnitudes in folded sedimentary rocks. Geophysical Research Letters 38, L17301.CrossRefGoogle Scholar
Amrouch, K, Lacombe, O, Bellahsen, N, Daniel, JM and Callot, JP (2010b) Stress and strain patterns, kinematics and deformation mechanisms in a basement-cored anticline: Sheep Mountain Anticline, Wyoming. Tectonics 29, TC1005.CrossRefGoogle Scholar
Amrouch, K, Robion, P, Callot, JP, Lacombe, O, Daniel, JM, Bellahsen, N and Faure, JL (2010a) Constraints on deformation mechanisms during folding provided by rock physical properties: a case study at Sheep Mountain anticline (Wyoming, USA). Geophysical Journal International 182, 1105–23.CrossRefGoogle Scholar
Andersen, MB, Stirling, CH, Potter, E-K, Halliday, AN, Blake, SG, McCulloch, MT, Ayling, BF and O’Leary, M (2008) High-precision U-series measurements of more than 500,000 year old fossil corals. Earth and Planetary Science Letters 265, 229–45.CrossRefGoogle Scholar
Anderson, NT, Kelson, JR, Kele, S, Daëron, M, Bonifacie, M, Horita, J, Mackey, TJ, John, CM, Kluge, T, Petschnig, P, Jost, AB, Huntington, KW, Bernasconi, SM and Bergmann, KD (2021) A unified clumped isotope thermometer calibration (0.5–1,100°C) using carbonate-based standardization. Geophysical Research Letters 48, e2020GL092069.CrossRefGoogle Scholar
André, AS, Sausse, J and Lespinasse, M (2001) New approach for the quantification of paleostress magnitudes: application to the Soultz vein system (Rhine graben, France). Tectonophysics 336, 215–31.CrossRefGoogle Scholar
Andresen, KJ (2012) Fluid flow features in hydrocarbon plumbing systems: what do they tell us about the basin evolution? Marine Geology 332, 89–108.CrossRefGoogle Scholar
Archie, GE (1952) Classification of carbonate reservoir rocks and petrophysical considerations. AAPG Bulletin 36, 278–98.Google Scholar
Arndt, M, Virgo, S, Cox, SF and Urai, JL (2014) Changes in fluid pathways in a calcite vein mesh (Natih Formation, Oman Mountains): insights from stable isotopes. Geofluids 14, 391418.CrossRefGoogle Scholar
Aubert, I, Léonide, P, Lamarche, J and Salardon, R (2020) Diagenetic evolution of fault zones in Urgonian microporous carbonates, impact on reservoir properties (Provence – Southeast France). Solid Earth 11, 1163–86.CrossRefGoogle Scholar
Baietto, A, Cadoppi, P, Martinotti, G, Perello, P, Perrochet, P and Vuataz, FD (2008) Assessment of thermal circulations in strike-slip fault systems: the Terme di Valdieri case (Italian western Alps). In The Geodynamics of the Aegean and Anatolia (eds T Taymaz, Y Yilmaz and Y Dilek), pp. 317–39. Geological Society of London, Special Publication no. 299.CrossRefGoogle Scholar
Bakker, RJ (2003) Package FLUIDS 1. Computer programs for analysis of fluid inclusion data and for modelling bulk fluid properties. Chemical Geology 194, 323.CrossRefGoogle Scholar
Ballentine, CJ and Burnard, PG (2002). Production, release and transport of noble gases in the continental crust. Reviews in Mineralogy and Geochemistry 47, 481–538.CrossRefGoogle Scholar
Bar, E, Nuriel, P, Kylander-Clark, A and Weinberger, R (2021) Towards in situ U-Pb dating of dolomite. Geochronology 3, 337–49.Google Scholar
Barbier, M, Floquet, M, Hamon, Y and Callot, JP (2015) Nature and distribution of diagenetic phases and petrophysical properties of carbonates: the Mississippian Madison Formation (Bighorn Basin, Wyoming, USA). Marine and Petroleum Geology 67, 230–48.CrossRefGoogle Scholar
Barbier, M, Hamon, Y, Callot, JP, Floquet, M and Daniel, JM (2012b) Sedimentary and diagenetic controls on the multiscale fracturing pattern of a carbonate reservoir: the Madison Formation (Sheep Mountain, Wyoming, USA). Marine and Petroleum Geology 29, 5067.CrossRefGoogle Scholar
Barbier, M, Leprêtre, R, Callot, JP, Gasparrini, M, Daniel, JM, Hamon, Y, Lacombe, O and Floquet, M (2012a) Impact of fracture stratigraphy on the paleo-hydrogeology of the Madison Limestone in two basement-involved folds in the Bighorn basin, (Wyoming, USA). Tectonophysics 576–577, 116–32.CrossRefGoogle Scholar
Barker, SL, Cox, SF, Eggins, SM and Gagan, MK (2006) Microchemical evidence for episodic growth of antitaxial veins during fracture-controlled fluid flow. Earth and Planetary Science Letters 250, 331–44.CrossRefGoogle Scholar
Barker, SLL, Bennett, VC, Cox, SF, Norman, MD and Gagan, MK (2009) Sm–Nd, Sr, C and O isotope systematics in hydrothermal calcite–fluorite veins: implications for fluid–rock reaction and geochronology. Chemical Geology 268, 5866.CrossRefGoogle Scholar
Bau, M and Möller, P (1993) Rare earth element systematics of the chemically precipitated component in early Precambrian iron formations and the evolution of the terrestrial atmosphere-hydrosphere-lithosphere system. Geochimica et Cosmochimica Acta 57, 2239–49.CrossRefGoogle Scholar
Beaudoin, N, Bellahsen, N, Lacombe, O and Emmanuel, L (2011) Fracture-controlled paleohydrogeology in a basement-cored, fault-related fold: Sheep Mountain Anticline, Wyoming, United States. Geochemistry, Geophysics, Geosystems 12, Q06011.CrossRefGoogle Scholar
Beaudoin, N, Bellahsen, N, Lacombe, O, Emmanuel, L and Pironon, J (2014a) Crustal-scale fluid flow during the tectonic evolution of the Bighorn Basin (Wyoming, USA). Basin Research 26, 403–35.CrossRefGoogle Scholar
Beaudoin, N and Lacombe, O (2018) Recent and future trends in paleopiezometry in the diagenetic domain: insights into the tectonic paleostress and burial depth history of fold-and-thrust belts and sedimentary basins. Journal of Structural Geology 114, 357–65.CrossRefGoogle Scholar
Beaudoin, N, Lacombe, O, Bellahsen, N, Amrouch, K and Daniel, J-M (2014b) Evolution of pore-fluid pressure during folding and basin contraction in overpressured reservoirs: insights from the Madison-Phosphoria carbonate formations in the Bighorn Basin (Wyoming, USA). Marine and Petroleum Geology 55, 214–29.CrossRefGoogle Scholar
Beaudoin, N, Lacombe, O, Bellahsen, N and Emmanuel, L (2013) Contribution of studies of sub-seismic fracture populations to paleo-hydrological reconstructions (Bighorn Basin, USA). Procedia Earth and Planetary Science 7, 5760.CrossRefGoogle Scholar
Beaudoin, N, Leprêtre, R, Bellahsen, N, Lacombe, O, Amrouch, K, Callot, J-P, Emmanuel, L and Daniel, J-M (2012) Structural and microstructural evolution of the Rattlesnake Mountain Anticline (Wyoming, USA): new insights into the Sevier and Laramide orogenic stress build-up in the Bighorn Basin. Tectonophysics 576–577, 2045.CrossRefGoogle Scholar
Beaudoin, G and Therrien, P (2004) The web stable isotope fractionation calculator. In Handbook of Stable Isotope Analytical Techniques, (ed P. de Groot), Elsevier pp. 1045–48.CrossRefGoogle Scholar
Beaudoin, N, Gasparrini, M, David, ME, Lacombe, O and Koehn, D (2019b) Bedding-parallel stylolites as a tool to unravel maximum burial depth in sedimentary basins: application to Middle Jurassic carbonate reservoirs in the Paris basin. Geological Society of America Bulletin 131, 1239–54.CrossRefGoogle Scholar
Beaudoin, N, Huyghe, D, Bellahsen, N, Lacombe, O, Emmanuel, L, Mouthereau, F and Ouanhnon, L (2015) Fluid systems and fracture development during syn-depositional fold growth: an example from the Pico del Aguila anticline, Sierras Exteriores, southern Pyrenees, Spain. Journal of Structural Geology 70, 2338.CrossRefGoogle Scholar
Beaudoin, N, Koehn, D, Lacombe, O, Lecouty, A, Billi, A, Aharonov, E and Parlangeau, C (2016) Fingerprinting stress: stylolite and calcite twinning paleopiezometry revealing the complexity of progressive stress patterns during folding: the case of the Monte Nero anticline in the Apennines, Italy. Tectonics 35, 1687–712.CrossRefGoogle Scholar
Beaudoin, NE, Labeur, A, Lacombe, O, Koehn, D, Billi, A, Hoareau, G, Boyce, A, John, CM, Marchegiano, M, Roberts, NM, Millar, IL, Claverie, F, Pecheyran, C and Callot, JP (2020c) Regional-scale paleofluid system across the Tuscan Nappe-Umbria-Marche Apennine Ridge (northern Apennines) as revealed by mesostructural and isotopic analyses of stylolite-vein networks. Solid Earth 11, 1617–41.CrossRefGoogle Scholar
Beaudoin, N, Lacombe, O, David, ME and Koehn, D (2020a) Does stress transmission in forelands depend on structural style? Distinctive stress magnitudes during Sevier thin-skinned and Laramide thick-skinned layer-parallel shortening in the Bighorn Basin (USA) revealed by stylolite and calcite twinning paleopiezometry. Terra Nova 32, 225–33.CrossRefGoogle Scholar
Beaudoin, N, Lacombe, O, Koehn, D, David, ME, Farrell, N and Healy, D (2020b) Vertical stress history and paleoburial in foreland basins unravelled by stylolite roughness paleopiezometry: insights from bedding-parallel stylolites in the Bighorn Basin, Wyoming, USA. Journal of Structural Geology 136, 104061.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, 1015–8.CrossRefGoogle Scholar
Beaudoin, N, Lacombe, O, Roberts, NMW and Koehn, D (2019a) U-Pb dating of calcite veins reveals complex stress evolution and thrust sequence in the Bighorn Basin, Wyoming, USA: REPLY. Geology 47, e481.CrossRefGoogle Scholar
Becker, JS (2002) State-of-the-art and progress in precise and accurate isotope ratio measurements by ICP-MS and LA-ICP-MS. Journal of Analytical Atomic Spectrometry 17, 1172–85.CrossRefGoogle Scholar
Bellahsen, N, Fiore, PE and Pollard, DD (2006a) From spatial variation of fracture patterns to fold kinematics: a geomechanical approach. Geophysical Research Letters 33, 14.CrossRefGoogle Scholar
Bellahsen, N, Fiore, P and Pollard, DD (2006b) The role of fractures in the structural interpretation of Sheep Mountain Anticline, Wyoming. Journal of Structural Geology 28, 850–67.CrossRefGoogle Scholar
Bense, VF, Gleeson, T, Loveless, SE, Bour, O and Scibek, J (2013) Fault zone hydrogeology. Earth-Science Reviews 127, 171–92.CrossRefGoogle Scholar
Bergbauer, S and Pollard, DD (2004) A new conceptual fold-fracture model including prefolding joints, based on the Emigrant Gap anticline, Wyoming. Bulletin of the Geological Society of America 116, 294307.CrossRefGoogle Scholar
Bergman, SC, Huntington, KW and Crider, JG (2013) Tracing paleofluid sources using clumped isotope thermometry of diagenetic cements along the Moab Fault, Utah. American Journal of Science 313, 490515.CrossRefGoogle Scholar
Berio, LR, Mittempergher, S, Storti, F, Bernasconi, SM, Cipriani, A, Lugli, F and Balsamo, F (2022) Open-closed-open paleofluid system conditions recorded in the tectonic vein networks of the Parmelan anticline (Bornes Massif, France). Journal of the Geological Society 179 (5), 2021–117. https://doi.org/10.1144/jgs2021-117.CrossRefGoogle Scholar
Bernasconi, SM, Daëron, M, Bergmann, KD, Bonifacie, M, Meckler, AN, Affek, HP, Anderson, N, Bajnai, D, Barkan, E, Beverly, E, Blamart, D, Burgener, L, Calmels, D, Chaduteau, C, Clog, M, Davidheiser-Kroll, B, Davies, A, Dux, F, Eiler, J, Elliott, B, Fetrow, AC, Fiebig, J, Goldberg, S, Hermoso, M, Huntington, KW, Hyland, E, Ingalls, M, Jaggi, M, John, CM, Jost, AB, Katz, S, Kelson, J, Kluge, T, Kocken, IJ, Laskar, A, Leutert, TJ, Liang, D, Lucarelli, J, Mackey, TJ, Mangenot, X, Meinicke, N, Modestou, SE, Müller, IA, Murray, S, Neary, A, Packard, N, Passey, BH, Pelletier, E, Petersen, S, Piasecki, A, Schauer, A, Snell, KE, Swart, PK, Tripati, A, Upadhyay, D, Vennemann, T, Winkelstern, I, Yarian, D, Yoshida, N, Zhang, N and Ziegler, M (2021) InterCarb: a community effort to improve interlaboratory standardization of the carbonate clumped isotope thermometer using carbonate standards. Geochemistry, Geophysics, Geosystems 22, e2020GC009588.CrossRefGoogle ScholarPubMed
Bernasconi, SM, Müller, IA, Bergmann, KD, Breitenbach, SFM, Fernandez, A, Hodell, DA, Jaggi, M, Meckler, AN, Millan, I and Ziegler, M (2018) Reducing uncertainties in carbonate clumped isotope analysis through consistent carbonate-based standardization. Geochemistry, Geophysics, Geosystems 19, 2895–914.CrossRefGoogle ScholarPubMed
Bilau, A, Bienveignant, D, Rolland, Y, Schwartz, S, Godeau, N, Guihou, A, Deschamps, P, Mangenot, X, Brigaud, B, Boschetti, L and Dumont, T (2022) The Tertiary structuration of the Western Subalpine foreland deciphered by calcite-filled faults and veins. Earth-Science Reviews 236, 104270.CrossRefGoogle Scholar
Bjørlykke, K (1994) Fluid-flow processes and diagenesis in sedimentary basins. In Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins (ed. Farrell, J.), pp. 127–40. Geological Society of London, Special Publication no. 78.Google Scholar
Bjørlykke, K (2015) Subsurface water and fluid flow in sedimentary basins. In Petroleum Geoscience: From Sedimentary Environments to Rock Physics, 2nd edn., pp. 279300. Berlin: Springer.CrossRefGoogle Scholar
Blamey, NJF (2012) Composition and evolution of crustal, geothermal and hydrothermal fluids interpreted using quantitative fluid inclusion gas analysis. Journal of Geochemical Exploration 116–117, 1727.CrossRefGoogle Scholar
Bodnar, RJ (2003) Reequilibration of fluid inclusions. In Fluid Inclusions: Analysis and Interpretation (eds Samson, I, Anderson, A and Marshall, D), pp. 213–30. Vancouver, Canada: Mineralogical Association of Canada Short Course 32.Google Scholar
Boggs, S and Krinsley, D (2006) Application of Cathodoluminescence Imaging to the Study of Sedimentary Rocks. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Böhlke, JK and Irwin, JJ (1992) Laser microprobe analyses of noble gas isotopes and halogens in fluid inclusions: analyses of microstandards and synthetic inclusions in quartz. Geochimica et Cosmochimica Acta 56, 187201.CrossRefGoogle Scholar
Bonifacie, M, Calmels, D, Eiler, JM, Horita, J, Chaduteau, C, Vasconcelos, C, Agrinier, P, Katz, A, Passey, BH, Ferry, JM and Bourrand, JJ (2017) Calibration of the dolomite clumped isotope thermometer from 25 to 350 °C, and implications for a universal calibration for all (Ca, Mg, Fe) CO3 carbonates. Geochimica et Cosmochimica Acta 200, 255–79.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
Bottinga, Y (1969) Calculated fractionation factors for carbon and hydrogen isotope exchange in the system calcite-carbon dioxide-graphite-methane-hydrogen-water vapor. Geochimica et Cosmochimica Acta 33, 4964.CrossRefGoogle Scholar
Bottinga, Y and Craig, H (1968) Oxygen isotope fractionation between CO2 and water, and the isotopic composition of marine atmospheric CO2 . Earth and Planetary Science Letters 5, 285–95.CrossRefGoogle Scholar
Bourdet, J, Pironon, J, Levresse, G and Tritlla, J (2008) Petroleum type determination through homogenization temperature and vapour volume fraction measurements in fluid inclusions. Geofluids 8, 4659.CrossRefGoogle Scholar
Bradbury, HJ and Woodwell, GR (1987) Ancient fluid flow within foreland terrains. In Petroleum Geology of the Continental Shelf of Northwest Europe: Oil and Source Rocks of the North Sea (eds LV Illing and GD Hobson), pp. 87102. Geological Society of London, Special Publication no. 34.CrossRefGoogle Scholar
Braithwaite, CJR (1989) Stylolites as open fluid conduits. Marine and Petroleum Geology 6, 9396.Google Scholar
Brannon, JC, Cole, SC, Podosek, FA, Ragan, VM, Coveney, RM, Wallace, MW and Bradley, AJ (1996) Th-Pb and U-Pb dating of ore-stage calcite and Paleozoic fluid flow. Science 271, 491–3.CrossRefGoogle Scholar
Breesch, L, Swennen, R, Dewever, B, Roure, F and Vincent, B (2011) Diagenesis and fluid system evolution in the northern Oman Mountains, United Arab Emirates: implications for petroleum exploration. GeoArabia 16, 111–48.CrossRefGoogle Scholar
Breesch, L, Swennen, R and Vincent, B (2006) Dolomite formation in breccias at the Musandam Platform border, Northern Oman Mountains, United Arab Emirates. Journal of Geochemical Exploration 89, 1922.CrossRefGoogle Scholar
Breesch, L, Swennen, R and Vincent, B (2009) Fluid flow reconstruction in hanging and footwall carbonates: compartmentalization by Cenozoic reverse faulting in the Northern Oman Mountains (UAE). Marine and Petroleum Geology 26, 113–28.CrossRefGoogle Scholar
Breesch, L, Swennen, R, Vincent, B, Ellison, R and Dewever, B (2010) Dolomite cementation and recrystallisation of sedimentary breccias along the Musandam Platform margin (United Arab Emirates). Journal of Geochemical Exploration 106, 3443.CrossRefGoogle Scholar
Bruna, PO, Lavenu, APC, Matonti, C and Bertotti, G (2019) Are stylolites fluid-flow efficient features? Journal of Structural Geology 125, 270–7.CrossRefGoogle Scholar
Caine, JS, Evans, JP and Forster, CB (1996) Fault zone architecture and permeability structure. Geology 24, 1025–8.2.3.CO;2>CrossRefGoogle Scholar
Caja, MA, Permanyer, A, Marfil, R, Al-Aasm, IS and Martín-Crespo, T (2006) Fluid flow record from fracture-fill calcite in the Eocene limestones from the South-Pyrenean Basin (NE Spain) and its relationship to oil shows. Journal of Geochemical Exploration 89, 2732.CrossRefGoogle Scholar
Calamita, F, Satolli, S, Scisciani, V, Esestime, P and Pace, P (2011) Contrasting styles of fault reactivation in curved orogenic belts: examples from the central Apennines (Italy). Bulletin of the Geological Society of America 123, 1097–111.CrossRefGoogle Scholar
Callot, JP, Breesch, L, Guilhaumou, N, Roure, F, Swennen, R and Vilasi, N (2010b) Paleo-fluids characterisation and fluid flow modelling along a regional transect in Northern United Arab Emirates (UAE). Arabian Journal of Geosciences 3, 413–37.CrossRefGoogle Scholar
Callot, JP, Robion, P, Sassi, W, Guiton, MLE, Faure, JL, Daniel, JM, Mengus, JM and Schmitz, J (2010a) Magnetic characterisation of folded aeolian sandstones: interpretation of magnetic fabrics in diamagnetic rocks. Tectonophysics 495, 230–45.CrossRefGoogle Scholar
Callot, JP, Sassi, W, Roure, F, Hill, K, Wilson, N and Divies, R (2017) Pressure and Basin modeling in Foothill belts: a study of the Kutubu area, Papua New Guinea Fold and Thrust Belt. AAPG Memoir 114, 165–89.Google Scholar
Callot, JP, Trocmé, V, Letouzey, J, Albouy, E, Jahani, S and Sherkati, S (2012) Pre-existing salt structures and the folding of the Zagros Mountains. In Salt Tectonics, Sediments and Prospectivity (ed. SG Archer), pp. 545–61. Geological Society of London, Special Publication no. 363.CrossRefGoogle Scholar
Capozzi, R and Picotti, V (2002) Fluid migration and origin of a mud volcano in the Northern Apennines (Italy): the role of deeply rooted normal faults. Terra Nova 14, 363–70.CrossRefGoogle Scholar
Caricchi, C, Aldega, L and Corrado, S (2014) Reconstruction of maximum burial along the Northern Apennines thrust wedge (Italy) by indicators of thermal exposure and modeling. Bulletin of the Geological Society of America 127, 428–42.CrossRefGoogle Scholar
Cathles, LM (1981) Fluid flow and genesis of hydrothermal ore deposits. In Economic Geology, Seventy-fifth Anniversary Volume (1905–1980) (eds. Brian J Skinner), pp. 424–57. El Paso, USA: The Economic Geology Publishing Company.Google Scholar
Caumon, M-C, Dubessy, J, Robert, P and Tarantola, A (2014) Fused-silica capillary capsules (FSCCs) as reference synthetic aqueous fluid inclusions to determine chlorinity by Raman spectroscopy. European Journal of Mineralogy 25, 755–63.CrossRefGoogle Scholar
Cello, G, Invernizzi, C, Mazzoli, S and Tondi, E (2001) Fault properties and fluid flow patterns from Quaternary faults in the Apennines, Italy. Tectonophysics 336, 6378.CrossRefGoogle Scholar
Centrella, S, Beaudoin, NE, Derluyn, H, Motte, G, Hoareau, G, Lanari, P, Piccoli, F, Pecheyran, C and Callot, JP (2021) Micro-scale chemical and physical patterns in an interface of hydrothermal dolomitization reveals the governing transport mechanisms in nature: case of the Layens anticline, Pyrenees, France. Sedimentology 68, 834–54.CrossRefGoogle Scholar
Centrella, S, Beaudoin, NE, Koehn, D, Motte, G, Hoareau, G and Callot, JP (2022) How fluid-mediated rock transformations can mimic hydro-fracturing patterns in hydrothermal dolomite. Marine and Petroleum Geology 140, 105657.CrossRefGoogle Scholar
Centrella, S, Putnis, A, Lanari, P and Austrheim, H (2018) Textural and chemical evolution of pyroxene during hydration and deformation: a consequence of retrograde metamorphism. Lithos 296–299, 245–64.CrossRefGoogle Scholar
Chacko, T and Deines, P (2008) Theoretical calculation of oxygen isotope fractionation factors in carbonate systems. Geochimica et Cosmochimica Acta 72, 3642–60.CrossRefGoogle Scholar
Chen, S, Ryb, U, Piasecki, AM, Lloyd, MK, Baker, MB and Eiler, JM (2019) Mechanism of solid-state clumped isotope reordering in carbonate minerals from aragonite heating experiments. Geochimica et Cosmochimica Acta 258, 156–73.CrossRefGoogle Scholar
Chi, G, Diamond, LW, Lu, H, Lai, J and Chu, H (2020) Common problems and pitfalls in fluid inclusion study: a review and discussion. Minerals 2021, 711. https://doi.org/10.3390/MIN11010007.CrossRefGoogle Scholar
Choukroune, P (1992) Tectonic evolution of the Pyrenees. Annual Review of Earth and Planetary Sciences 20, 143.CrossRefGoogle Scholar
Clauer, N (2013) The K-Ar and 40Ar/39Ar methods revisited for dating fine-grained K-bearing clay minerals. Chemical Geology 354, 163–85.CrossRefGoogle Scholar
Cobbold, PR, Zanella, A, Rodrigues, N and Løseth, H (2013) Bedding-parallel fibrous veins (beef and cone-in-cone): worldwide occurrence and possible significance in terms of fluid overpressure, hydrocarbon generation and mineralization. Marine and Petroleum Geology 43, 120.CrossRefGoogle Scholar
Cooley, MA, Price, RA, Kyser, TK and Dixon, JM (2011) Stable-isotope geochemistry of syntectonic veins in Paleozoic carbonate rocks in the Livingstone Range anticlinorium and their significance to the thermal and fluid evolution of the southern Canadian foreland thrust and fold belt. AAPG Bulletin 95, 1851–82.CrossRefGoogle Scholar
Coppola, M, Correale, A, Barberio, MD, Billi, A, Cavallo, A, Fondriest, M, Nazzari, M, Paonita, A, Romano, C, Stagno, V, Viti, C and Vona, JM (2021) Meso- to nano-scale evidence of fluid-assisted co-seismic slip along the normal Mt. Morrone Fault, Italy: implications for earthquake hydrogeochemical precursors. Earth and Planetary Science Letters 568, 117010.CrossRefGoogle Scholar
Craddock, JP and Relle, M (2003) Fold axis-parallel rotation within the Laramide Derby Dome Fold, Wind River Basin, Wyoming, USA. Journal of Structural Geology 25, 1959–72.CrossRefGoogle Scholar
Craddock, JP and van der Pluijm, BA (1989) Late Paleozoic deformation of the cratonic carbonate cover of eastern North America. Geology 17, 416–9.2.3.CO;2>CrossRefGoogle Scholar
Craddock, JP and van der Pluijm, BA (1999) Sevier-Laramide deformation of the continental interior from calcite twinning analysis, west-central North America. Tectonophysics 305, 275–86.CrossRefGoogle Scholar
Craig, H (1957) Isotopic standards for carbon and oxygen and correction factors for mass-spectrometric analysis of carbon dioxide. Geochimica et Cosmochimica Acta 12, 133–49.CrossRefGoogle Scholar
Craig, H (1961) Standard for reporting concentrations of deuterium and oxygen-18 in natural waters. Science 133, 1833–4.CrossRefGoogle ScholarPubMed
Crognier, N, Hoareau, G, Aubourg, C, Dubois, M, Lacroix, B, Branellec, M, Callot, JP and Vennemann, T (2018) Syn-orogenic fluid flow in the Jaca basin (south Pyrenean fold and thrust belt) from fracture and vein analyses. Basin Research 30, 187216.CrossRefGoogle Scholar
Cruset, D, Cantarero, I, Travé, A, Vergés, J and John, CM (2016) Crestal graben fluid evolution during growth of the Puig-Reig anticline (South Pyrenean fold and thrust belt). Journal of Geodynamics 101, 3050.CrossRefGoogle Scholar
Cruset, D, Cantarero, I, Vergés, J, John, CM, Muñoz-López, D and Travé, A (2018) Changes in fluid regime in syn-orogenic sediments during the growth of the south Pyrenean fold and thrust belt. Global and Planetary Change 171, 207–24.CrossRefGoogle Scholar
Cruset, D, Vergés, J, Benedicto, A, Gomez-Rivas, E, Cantarero, I, John, CM and Travé, A (2021) Multiple fluid flow events from salt-related rifting to basin inversion (Upper Pedraforca thrust sheet, SE Pyrenees). Basin Research 33, 3102–36.CrossRefGoogle Scholar
Curtis, CD (1987) Inorganic geochemistry and petroleum exploration. In Advances in Petroleum Geochemistry (eds Brooks, J and Welte, D), pp. 91141. London: Academic Press.Google Scholar
Curzi, M, Aldega, L, Bernasconi, SM, Berra, F, Billi, A, Boschi, C, Franchini, S, van der Lelij, R, Viola, G and Carminati, E (2020) Architecture and evolution of an extensionally-inverted thrust (Mt. Tancia Thrust, Central Apennines): geological, structural, geochemical, and K–Ar geochronological constraints. Journal of Structural Geology 136, 104059.CrossRefGoogle Scholar
Curzi, M, Bernasconi, SM, Billi, A, Boschi, C, Aldega, L, Franchini, S, Albert, R, Gerdes, A, Barberio, MD and Carminati, E (2021) U–Pb age of the 2016 Amatrice earthquake causative fault (Mt. Gorzano, Italy) and paleo-fluid circulation during seismic cycles inferred from inter-and co-seismic calcite. Tectonophysics 819, 229076.CrossRefGoogle Scholar
Daëron, M (2021) Full propagation of analytical uncertainties in Δ47 measurements. Geochemistry, Geophysics, Geosystems 22, e2020GC009592.CrossRefGoogle Scholar
D’Amore, F and Panichi, C (1980) Evaluation of deep temperatures of hydrothermal systems by a new gas geothermometer. Geochimica et Cosmochimica Acta 44, 549–56.CrossRefGoogle Scholar
Dassié, EP, Genty, D, Noret, A, Mangenot, X, Massault, M, Lebas, N, Duhamel, M, Bonifacie, M, Gasparrini, M, Minster, B and Michelot, JL (2018) A newly designed analytical line to examine fluid inclusion isotopic compositions in a variety of carbonate samples. Geochemistry, Geophysics, Geosystems 19, 1107–22.CrossRefGoogle Scholar
Davies, GR and Smith, LB (2006) Structurally controlled hydrothermal dolomite reservoir facies: an overview. American Association of Petroleum Geologists Bulletin 90, 1641–90.CrossRefGoogle Scholar
Davis, DW, Lowenstein, TK and Spencer, RJ (1990) Melting behavior of fluid inclusions in laboratory-grown halite crystals in the systems NaCl-H2O, NaCl-KCl-H2O, NaCl-MgCl2-H2O, and NaCl-CaCl2-H2O. Geochimica et Cosmochimica Acta 54, 591601 CrossRefGoogle Scholar
de Graaf, S, Nooitgedacht, CW, Goff, JLE, van der Lubbe, JHJL, Vonhof, HB and Reijmer, JJG (2019) Fluid-flow evolution in the Albanide fold-thrust belt: insights from hydrogen and oxygen isotope ratios of fluid inclusions. AAPG Bulletin 103, 2421–45.CrossRefGoogle Scholar
del Sole, L, Antonellini, M, Soliva, R, Ballas, G, Balsamo, F and Viola, G (2020) Structural control on fluid flow and shallow diagenesis: insights from calcite cementation along deformation bands in porous sandstones. Solid Earth 11, 2169–195.CrossRefGoogle Scholar
DeCelles, PG (2004). Late Jurassic to Eocene evolution of the Cordilleran thrust belt and foreland basin system, western USA. American Journal of Science 304, 105–68.CrossRefGoogle Scholar
Dellinger, M, Gaillardet, J, Bouchez, J, Calmels, D, Louvat, P, Dosseto, A, Gorge, C, Alanoca, L and Maurice, L (2015) Riverine Li isotope fractionation in the Amazon River basin controlled by the weathering regimes. Geochimica et Cosmochimica Acta 164, 7193.CrossRefGoogle Scholar
Deloule, E and Turcotte, DL (1989) The flow of hot brines in cracks and the formation of ore deposits. Economic Geology 84, 2217–25.CrossRefGoogle Scholar
Dielforder, A, Villa, IM, Berger, A and Herwegh, M (2022) Tracing wedge-internal deformation by means of strontium isotope systematics of vein carbonates. Geological Magazine. https://doi.org/10.1017/ S0016756821001357 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.CrossRefGoogle Scholar
Dreybrodt, W and Deininger, M (2014) The impact of evaporation to the isotope composition of DIC in calcite precipitating water films in equilibrium and kinetic fractionation models. Geochimica et Cosmochimica Acta 125, 433–9.CrossRefGoogle Scholar
Dromgoole, EL.and Walter, LM (1990) Iron and manganese incorporation into calcite: effects of growth kinetics, temperature and solution chemistry. Chemical Geology 81, 311–36.CrossRefGoogle Scholar
Drost, K, Chew, D, Petrus, JA, Scholze, F, Woodhead, J D, Schneider, JW and Harper, DA (2018) An image mapping approach to U-Pb LA-ICP-MS carbonate dating and applications to direct dating of carbonate sedimentation. Geochemistry, Geophysics, Geosystems 19, 4631–48.CrossRefGoogle Scholar
Eichhubl, P, Taylor, WL, Pollard, DD and Aydin, A (2004) Paleo-fluid flow and deformation in the Aztec Sandstone at the Valley of Fire, Nevada: evidence for the coupling of hydrogeologic, diagenetic, and tectonic processes. GSA Bulletin 116, 1120–36.CrossRefGoogle Scholar
Eiler, JM and Schauble, E (2004) 18O13C16O in Earth’s atmosphere. Geochimica et Cosmochimica Acta 68, 4767–77.CrossRefGoogle Scholar
Elter, P, Grasso, M, Parotto, M and Vezzani, L (2008) Structural setting of the Apennine-Maghrebian thrust belt. Episodes 26, 205–11.CrossRefGoogle Scholar
Emery, D and Robinson, A (1993) Inorganic Geochemistry: Applications to Petroleum Geology, 255 pp. Oxford, UK: Blackwell Scientific Publications.CrossRefGoogle Scholar
Engel, J, Maas, R, Woodhead, J, Tympel, J and Greig, A (2020) A single-column extraction chemistry for isotope dilution U-Pb dating of carbonate. Chemical Geology 531, 119311.CrossRefGoogle Scholar
English, JM, Johnston, ST and Wang, K (2003). Thermal modelling of the Laramide orogeny: testing the flat-slab subduction hypothesis. Earth and Planetary Science Letters 214, 619–32.CrossRefGoogle Scholar
Epstein, S, Buchsbaum, R, Lowenstam, HA and Urey, HC (1953) Revised carbonate-water isotopic temperature scale. GSA Bulletin 64, 1315–26.CrossRefGoogle Scholar
Erslev, EA (1986) Basement balancing of Rocky Mountain foreland uplifts (USA). Geology 14, 259–62.2.0.CO;2>CrossRefGoogle Scholar
Evans, MA (2010) Temporal and spatial changes in deformation conditions during the formation of the Central Appalachian fold-and-thrust belt: evidence from joints, vein mineral paragenesis, and fluid inclusions. In From Rodinia to Pangea: The Lithotectonic Record of the Appalachian Region (ed. Tollo, RP), pp. 417518: Boulder, Colorado: Geological Society of America. GSA Memoir.Google Scholar
Evans, MA and Battle, DA (1999) Fluid inclusion and stable isotope analyses of veins from the central Appalachian Valley and Ridge province: implications for regional synorogenic hydrologic structure and fluid migration. GSA Bulletin 111, 1841–60.2.3.CO;2>CrossRefGoogle Scholar
Evans, MA, Bebout, GE and Brown, CH (2012) Changing fluid conditions during folding: an example from the central Appalachians. Tectonophysics 576–577, 99115.CrossRefGoogle Scholar
Evans, MA and Fischer, MP (2012) On the distribution of fluids in folds: a review of controlling factors and processes. Journal of Structural Geology 44, 224.CrossRefGoogle Scholar
Evans, MA and Hobbs, GC (2003) Fate of ‘warm’ migrating fluids in the central Appalachians during the Late Paleozoic Alleghanian orogeny. Journal of Geochemical Exploration 78–79, 327–31.CrossRefGoogle Scholar
Fantle, MS and Higgins, J (2014) The effects of diagenesis and dolomitization on Ca and Mg isotopes in marine platform carbonates: implications for the geochemical cycles of Ca and Mg. Geochimica et Cosmochimica Acta 142, 458–81.CrossRefGoogle Scholar
Faulkner, DR, Jackson, CAL, Lunn, RJ, Schlische, RW, Shipton, ZK, Wibberley, CAJ and Withjack, MO (2010) A review of recent developments concerning the structure, mechanics and fluid flow properties of fault zones. Journal of Structural Geology 32, 1557–75.CrossRefGoogle Scholar
Ferket, H, Guilhaumou, N, Roure, F and Swennen, R (2011) Insights from fluid inclusions, thermal and PVT modeling for paleo-burial and thermal reconstruction of the Córdoba petroleum system (NE Mexico). Marine and Petroleum Geology 28, 936–58.CrossRefGoogle Scholar
Ferket, H, Roure, F, Swennen, R and Ortuño, S (2000) Fluid migration placed into the deformation history of fold-and-thrust belts: an example from the Veracruz basin (Mexico). Journal of Geochemical Exploration 69–70, 275–9.CrossRefGoogle Scholar
Ferket, H, Swennen, R, Ortuño, S and Roure, F (2003) Reconstruction of the fluid flow history during Laramide forelandfold and thrust belt development in eastern Mexico: cathodoluminescence and δ18O-δ13C isotope trends of calcite-cemented fractures. Journal of Geochemical Exploration 78–79, 163–7.CrossRefGoogle Scholar
Ferket, H, Swennen, R, Ortuño Arzate, S and Roure, F (2006) Fluid flow evolution in petroleum reservoirs with a complex diagenetic history: an example from Veracruz, Mexico. Journal of Geochemical Exploration 89, 108–11.CrossRefGoogle Scholar
Fischer, MP, Higuera-Díaz, IC, Evans, MA, Perry, EC and Lefticariu, L (2009) Fracture-controlled paleohydrology in a map-scale detachment fold: insights from the analysis of fluid inclusions in calcite and quartz veins. Journal of Structural Geology 31, 1490–510.CrossRefGoogle Scholar
Fischer, MP and Jackson, PB (1999) Stratigraphic controls on deformation patterns in fault-related folds: a detachment fold example from the Sierra Madre Oriental, northeast Mexico. Journal of Structural Geology 21, 613–33.CrossRefGoogle Scholar
Fitz-Diaz, E, Hudleston, P, Siebenaller, L, Kirschner, D, Camprubí, A, Tolson, G and Puig, TP (2011a) Insights into fluid flow and water-rock interaction during deformation of carbonate sequences in the Mexican fold-thrust belt. Journal of Structural Geology 33, 1237–53.CrossRefGoogle Scholar
Fitz-Diaz, E, Hudleston, P and Tolson, G (2011b) Comparison of tectonic styles in the Mexican and Canadian Rocky Mountain Fold-Thrust Belt. In Kinematic Evolution and Structural Styles of Fold-and-Thrust Belts (eds Poblet, J and Lisle, RJ), pp.149–67. Geological Society of London, Special Publication no. 349.Google Scholar
Fossen, H, Schultz, RA. Shipton, ZK and Mair, K (2007) Deformation bands in sandstone: a review. Journal of the Geological Society 164, 755–69.CrossRefGoogle Scholar
Gabellone, T, Gasparrini, M, Iannace, A, Invernizzi, C, Mazzoli, S and D’Antonio, M (2013) Fluid channeling along thrust zones: the Lagonegro case history, southern Apennines, Italy. Geofluids 13, 140–58.CrossRefGoogle Scholar
Gao, J, He, S, Zhao, JX, He, Z, Wu, C, Feng, Y, Nguyen, AD, Zhou, J and Yi, Z (2020) Sm-Nd isochron dating and geochemical (rare earth elements, 87Sr/86Sr, δ18O, δ13C) characterization of calcite veins in the Jiaoshiba shale gas field, China: implications for the mechanisms of vein formation in shale gas systems. GSA Bulletin 132, 1722–40.CrossRefGoogle Scholar
Ge, S and Garven, G (1994) A theoretical model for thrust-induced deep groundwater expulsion with application to the Canadian Rocky Mountains. Journal of Geophysical Research 99, 13851–868.CrossRefGoogle Scholar
Ghisetti, F, Kirschner, DL, Vezzani, L and Agosta, F (2001) Stable isotope evidence for contrasting paleofluid circulation in thrust faults and normal faults of the central Apennines, Italy. Journal of Geophysical Research: Solid Earth 106, 8811–25.CrossRefGoogle Scholar
Ghisetti, F and Vezzani, L (2000) Detachments and normal faulting in the Marche fold-and-thrust belt (central Apennines, Italy): inferences on fluid migration paths. Journal of Geodynamics 29, 345–69.CrossRefGoogle Scholar
Ghosh, P, Adkins, J, Affek, H, Balta, B, Guo, W, Schauble, EA, Schrag, D and Eiler, JM (2006) 13C-18O bonds in carbonate minerals: a new kind of paleothermometer. Geochimica et Cosmochimica Acta 70, 1439–56.CrossRefGoogle Scholar
Giggenbach, WF (1996) Chemical composition cf volcanic gases. In Monitoring and Mitigation of Volcano Hazards, (eds Scarpa/Tilling), pp. 221–56. Germany: Springer-Verlag Berlin.Google Scholar
Goldstein, RH (1986) Reequilibration of fluid inclusions in low-temperature calcium-carbonate cement. Geology 14, 792–5.2.0.CO;2>CrossRefGoogle Scholar
Goldstein, RH and Reynolds, TJ (1994) Systematics of Fluid Inclusions in Diagenetic Minerals. SEPM Short Course 31, 213 pp.CrossRefGoogle Scholar
Gomez-Rivas, E, Corbella, M, Martín-Martín, JD, Stafford, SL, Teixell, A, Bons, PD, Griera, A and Cardellach, E (2014) Reactivity of dolomitizing fluids and Mg source evaluation of fault-controlled dolomitization at the Benicàssim outcrop analogue (Maestrat basin, E Spain). Marine and Petroleum Geology 55, 2642.CrossRefGoogle Scholar
Gonzalez, E, Ferket, H, Callot, J-P, Guillhaumou, N, Ortuno, S and Roure, F (2012) Paleoburial, hydrocarbon generation, and migration in the Córdoba Platform and Veracruz Basin: insights from fluid inclusion studies and two-dimensional (2D) basin modeling. In Analyzing the Thermal History of Sedimentary Basins: Methods and Case Studies (ND Naeser and TH McCulloh), pp. 167–86. Tulsa, Oklahoma: SEPM Special Publication.Google Scholar
Gratier, JP, Frery, E, Deschamps, P, Røyne, A, Renard, F, Dysthe, D, Ellouz-Zimmerman, N and Hamelin, B (2012) How travertine veins grow from top to bottom and lift the rocks above them: the effect of crystallization force. Geology 40, 1015–18.CrossRefGoogle Scholar
Graustein, WC (1989) 87Sr/86Sr ratios measure the sources and flow of strontium in terrestrial ecosystems. In Stable Isotopes in Ecological Research. Ecological Studies (Analysis and Synthesis) (eds Rundel, PW, Ehleringer, JR and Nagy, KA), pp. 491512. New York: Springer.CrossRefGoogle Scholar
Groshong, RH, Kronenberg, A, Couzens-Schultz, BA and Newman, J (2014) Fluids and structures in fold and thrust belts with recognition of the work of David Wiltschko. Journal of Structural Geology 69, 281–3.CrossRefGoogle Scholar
Guillaume, D, Teinturier, S, Dubessy, J and Pironon, J (2003) Calibration of methane analysis by Raman spectroscopy in H2O–NaCl–CH4 fluid inclusions. Chemical Geology 194, 41–9.CrossRefGoogle Scholar
Guiton, MLE, Leroy, YM and Sassi, W (2003) Activation of diffuse discontinuities and folding of sedimentary layers. Journal of Geophysical Research: Solid Earth 108, 120.CrossRefGoogle Scholar
Gussone, N, Ahm, ASC, Lau, KV and Bradbury, HJ (2020) Calcium isotopes in deep time: potential and limitations. Chemical Geology 544, 119601.CrossRefGoogle Scholar
Haines, S, Lynch, E, Mulch, A, Valley, JW and van der Pluijm, B (2016) Meteoric fluid infiltration in crustal-scale normal fault systems as indicated by δ18O and δ2H geochemistry and 40Ar/39Ar dating of neoformed clays in brittle fault rocks. Lithosphere 8, 587–600.CrossRefGoogle Scholar
Haines, SH and van der Pluijm, BA (2008) Clay quantification and Ar–Ar dating of synthetic and natural gouge: application to the Miocene Sierra Mazatán detachment fault, Sonora, Mexico. Journal of Structural Geology 30, 525–38.CrossRefGoogle Scholar
Hairuo Qing and Mountjoy, EW (1994) Formation of coarsely crystalline, hydrothermal dolomite reservoirs in the Presqu’ile Barrier, Western Canada Sedimentary Basin. AAPG Bulletin 78, 5577.Google Scholar
Hanks, CL, Parris, TM and Wallace, WK (2006) Fracture paragenesis and microthermometry in Lisburne Group detachment folds: implications for the thermal and structural evolution of the northeastern Brooks Range, Alaska. AAPG Bulletin 90, 120.CrossRefGoogle Scholar
Heap, M, Reuschlé, T, Baud, P, Renard, F and Iezzi, G (2018) The permeability of stylolite-bearing limestone. Journal of Structural Geology 116, 8193.CrossRefGoogle Scholar
Heinze, T, Hamidi, S and Galvan, B. (2017). A dynamic heat transfer coefficient between fractured rock and flowing fluid. Geothermics 65, 10–16.CrossRefGoogle Scholar
Henderson, IHC and McCaig, AM (1996) Fluid pressure and salinity variations in shear zone-related veins, central Pyrenees, France: implications for the fault-valve model. Tectonophysics 262, 321–48.CrossRefGoogle Scholar
Henley, R, Truesdell, A, Barton, P and Whitney, J (1984) Fluid-mineral equilibria in hydrothermal systems. In Reviews in Economic Geology 1 (eds Henley, Truesdell and Barton, Jr). Littleton, USA: Society of Economic Geologists.CrossRefGoogle Scholar
Hilgers, C, Dilg-Gruschinski, K and Urai, JL (2004) Microstructural evolution of syntaxial veins formed by advective flow. Geology 32, 261–4.CrossRefGoogle Scholar
Hilgers, C, Kirschner, DL, Breton, JP and Urai, JL (2006) Fracture sealing and fluid overpressures in limestones of the Jabal Akhdar dome, Oman Mountains. Geofluids 6, 168–84.CrossRefGoogle Scholar
Hilgers, C and Urai, JL (2002a) Experimental study of syntaxial vein growth during lateral fluid flow in transmitted light: first results. Journal of Structural Geology 24, 1029–43.CrossRefGoogle Scholar
Hilgers, C and Urai, JL (2002b) Microstructural observations on natural syntectonic fibrous veins: implications for the growth process. Tectonophysics 352, 257–74.CrossRefGoogle Scholar
Hnat, JS and van der Pluijm, BA (2014) Fault gouge dating in the Southern Appalachians, USA. GSA Bulletin 126, 639–51.CrossRefGoogle Scholar
Hoareau, G, Claverie, F, Pecheyran, C, Paroissin, C, Grignard, P-A, Motte, G, Chailan, O and Girard, J-P (2021b) Direct U–Pb dating of carbonates from micron-scale femtosecond laser ablation inductively coupled plasma mass spectrometry images using robust regression. Geochronology 3, 6787.CrossRefGoogle Scholar
Hoareau, G, Crognier, N, Lacroix, B, Aubourg, C, Roberts, NMW, Niemi, N, Branellec, M, Beaudoin, N and Suarez Ruiz, I (2021a) Combination of Δ47 and U-Pb dating in tectonic calcite veins unravel the last pulses related to the Pyrenean Shortening (Spain). Earth and Planetary Science Letters 553, 116636.CrossRefGoogle Scholar
Holland, M, Saxena, N and Urai, JL (2009a) Evolution of fractures in a highly dynamic thermal, hydraulic, and mechanical system-(II) remote sensing fracture analysis, Jabal Shams, Oman Mountains. GeoArabia 14, 163–94.CrossRefGoogle Scholar
Holland, M and Urai, JL (2010) Evolution of anastomosing crack–seal vein networks in limestones: insight from an exhumed high-pressure cell, Jabal Shams, Oman Mountains. Journal of Structural Geology 32, 1279–90.CrossRefGoogle Scholar
Holland, M, Urai, JL, Muchez, P and Willemse, EJM (2009b) Evolution of fractures in a highly dynamic thermal, hydraulic, and mechanical system – (I) Field observations in Mesozoic Carbonates, Jabal Shams, Oman Mountains. GeoArabia 14, 57110.CrossRefGoogle Scholar
Hollocher, K (1991) Prograde amphibole dehydration reactions during high-grade regional metamorphism, central Massachusetts, USA. American Mineralogist 76, 956–70.Google Scholar
Horita, J (2014) Oxygen and carbon isotope fractionation in the system dolomite–water–CO2 to elevated temperatures. Geochimica et Cosmochimica Acta 129, 111–24.CrossRefGoogle Scholar
Horton, BK, Capaldi, TN, Mackaman-Lofland, C, Perez, ND, Bush, MA, Fuentes, F and Constenius, KN (2022) Broken foreland basins and the influence of subduction dynamics, tectonic inheritance, and mechanical triggers. Earth-Science Reviews, 104193.CrossRefGoogle Scholar
Hough, BG, Fan, M and Passey, BH (2014) Calibration of the clumped isotope geothermometer in soil carbonate in Wyoming and Nebraska, USA: implications for paleoelevation and paleoclimate reconstruction. Earth and Planetary Science Letters 391, 110–20.CrossRefGoogle Scholar
Hu, G and Clayton, RN (2003) Oxygen isotope salt effects at high pressure and high temperature and the calibration of oxygen isotope geothermometers. Geochimica et Cosmochimica Acta 67, 3227–46.CrossRefGoogle Scholar
Hudec, MR and Jackson, MPA (2007) Terra infirma: understanding salt tectonics. Earth-Science Reviews 82, 128.CrossRefGoogle Scholar
Humphrey, E, Gomez-Rivas, E, Koehn, D, Bons, PD, Neilson, J, Martín-Martín, JD and Schoenherr, J (2019) Stylolite-controlled diagenesis of a mudstone carbonate reservoir: a case study from the Zechstein_2_Carbonate (Central European Basin, NW Germany). Marine and Petroleum Geology 109, 88107.CrossRefGoogle Scholar
Huntington, KW, Budd, DA, Wernicke, BP and Eiler, JM (2011) Use of clumped-isotope thermometry to constrain the crystallization temperature of diagenetic calcite. Journal of Sedimentary Research 81, 656–69.CrossRefGoogle Scholar
Husson, JM, Higgins, JA, Maloof, AC and Schoene, B (2015) Ca and Mg isotope constraints on the origin of Earth’s deepest δ13C excursion. Geochimica et Cosmochimica Acta 160, 243–66.CrossRefGoogle Scholar
Irwin, H, Curtis, C and Coleman, M (1977) Isotopic evidence for source of diagenetic carbonates formed during burial of organic-rich sediments. Nature 269, 209–13.CrossRefGoogle Scholar
Jakubowicz, M, Agirrezabala, LM, Belka, Z, Siepak, M and Dopieralska, J (2022) Sr–Nd isotope decoupling at Cretaceous hydrocarbon seeps of the Basque-Cantabrian Basin (Spain): implications for tracing volcanic-influenced fluids in sedimented rifts. Marine and Petroleum Geology 135, 105430.CrossRefGoogle Scholar
Jin, XY, Zhao, JX, Feng, YX, Hofstra, AH, Deng, XD, Zhao, XF and Li, JW (2021) Calcite U-Pb dating unravels the age and hydrothermal history of the giant Shuiyindong carlin-type gold deposit in the Golden Triangle, South China. Economic Geology 116, 1253–65.CrossRefGoogle Scholar
Jonas, L, John, T, King, HE, Geisler, T and Putnis, A (2014) The role of grain boundaries and transient porosity in rocks as fluid pathways for reaction front propagation. Earth and Planetary Science Letters 386, 6474.CrossRefGoogle Scholar
Jordan, TE and Allmendinger, RW (1986). The Sierras Pampeanas of Argentina: a modern analogue of Rocky Mountain foreland deformation. American Journal of Science 286, 737–64.CrossRefGoogle Scholar
Kareem, KH, Al-Aasm, IS and Mansurbeg, H (2019) Structurally-controlled hydrothermal fluid flow in an extensional tectonic regime: a case study of Cretaceous Qamchuqa Formation, Zagros Basin, Kurdistan Iraq. Sedimentary Geology 386, 5278.CrossRefGoogle Scholar
Kergaravat, C, Ribes, C, Legeay, E, Callot, JP, Kavak, KS and Ringenbach, JC (2016) Minibasins and salt canopy in foreland fold-and-thrust belts: the central Sivas Basin, Turkey. Tectonics 35, 1342–66.CrossRefGoogle Scholar
Kim, ST and O’Neil, JR (1997) Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates. Geochimica et Cosmochimica Acta 61, 3461–75.CrossRefGoogle Scholar
Kirkwood, D, Ayt-Ougougdal, M, Gayot, T, Beaudoin, G and Pironon, J (2000) Paleofluid-flow in a foreland basin, Northern Appalachians: from syntectonic flexural extension to Taconian overthrusting. Journal of Geochemical Exploration 69–70, 269–73.CrossRefGoogle Scholar
Kirschner, DL and Kennedy, LA (2001) Limited syntectonic fluid flow in carbonate-hosted thrust faults of the Front Ranges, Canadian Rockies, inferred from stable isotope data and structures. Journal of Geophysical Research: Solid Earth 106, 8827–40.CrossRefGoogle Scholar
Koehn, D, Rood, MP, Beaudoin, N, Chung, P, Bons, PD and Gomez-Rivas, E (2016) A new stylolite classification scheme to estimate compaction and local permeability variations. Sedimentary Geology 346, 60–71.CrossRefGoogle Scholar
Koeshidayatullah, A, Corlett, H, Stacey, J, Swart, PK, Boyce, and Hollis, C (2020b) Origin and evolution of fault-controlled hydrothermal dolomitization fronts: a new insight. Earth and Planetary Science Letters 541, 116291.CrossRefGoogle Scholar
Koeshidayatullah, A, Corlett, H, Stacey, J, Swart, PK, Boyce, A, Robertson, H, Whitaker, F and Hollis, C (2020a) Evaluating new fault-controlled hydrothermal dolomitization models: insights from the Cambrian Dolomite, Western Canadian Sedimentary Basin. Sedimentology 67, 2945–73.Google Scholar
Kylander-Clark, ARC (2020) Expanding the limits of laser-ablation U–Pb calcite geochronology. Geochronology 2, 343–54.CrossRefGoogle Scholar
Labeur, A, Beaudoin, NE, Lacombe, O, Emmanuel, L, Petracchini, L, Daëron, M, Klimowicz, S and Callot, JP (2021) Burial-deformation history of folded rocks unraveled by fracture analysis, stylolite paleopiezometry and vein cement geochemistry: a case study in the Cingoli Anticline (Umbria-Marche, Northern Apennines). Geosciences 11, 135.CrossRefGoogle Scholar
Lacombe, O (2010) Calcite twins, a tool for tectonic studies in thrust belts and stable orogenic forelands. Oil & Gas Science and Technology – Revue d’IFP Energies nouvelles 65, 809–38.CrossRefGoogle Scholar
Lacombe, O, Beaudoin, NE, Hoareau, G, Labeur, A, Pecheyran, C and Callot, JP (2021) Dating folding beyond folding, from layer-parallel shortening to fold tightening, using mesostructures: lessons from the Apennines, Pyrenees, and Rocky Mountains. Solid Earth 12, 2145–57.CrossRefGoogle Scholar
Lacombe, O and Bellahsen, N (2016) Thick-skinned tectonics and basement-involved fold-thrust belts: insights from selected Cenozoic orogens. Geological Magazine 153, 763810.CrossRefGoogle Scholar
Lacombe, O and Mouthereau, F (1999) What is the real front of orogens? The Pyrenean orogen as a case study. Compte rendu de l’académie des sciences 329, 889–96.Google Scholar
Lacombe, O and Rolland, Y (2016) An introduction to the Special Issue on “Fluids in crustal deformation: fluid flow, fluid-rock interactions, rheology, melting and resources”. Journal of Geodynamics 101, 14.CrossRefGoogle Scholar
Lacombe, O, Swennen, R and Caracausi, A (2014) An introduction to the Special Issue on “Fluid-rock-tectonics interactions in basins and orogens”. Marine and Petroleum Geology 55, 15.CrossRefGoogle Scholar
Lacroix, B, Baumgartner, LP, Bouvier, AS, Kempton, PD and Vennemann, T (2018) Multi fluid-flow record during episodic mode I opening: a microstructural and SIMS study (Cotiella Thrust Fault, Pyrenees). Earth and Planetary Science Letters 503, 3746.CrossRefGoogle Scholar
Lacroix, B, Buatier, M, Labaume, P, Trave, 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
Lacroix, B, Leclère, H, Buatier, M and Fabbri, O (2013) Weakening processes in thrust faults: insights from the Monte Perdido thrust fault (southern Pyrenees, Spain). Geofluids 13, 5665.CrossRefGoogle Scholar
Lacroix, B, Trave, A, Buatier, M, Labaume, P, Vennemann, T and Dubois, M (2014) Syntectonic fluid-flow along thrust faults: example of the South-Pyrenean fold-and-thrust belt. Marine and Petroleum Geology 49, 8498.CrossRefGoogle Scholar
Laubach, SE, Eichhubl, P, Hilgers, C and Lander, RH (2010) Structural diagenesis. Journal of Structural Geology 32, 1866–72.CrossRefGoogle Scholar
Laubach, SE, Lander, RH, Criscenti, LJ, Anovitz, LM, Urai, JL, Pollyea, RM, Hooker, JN, Narr, W, Evans, MA, Kerisit, SN, Olson, JE, Dewers, T, Fisher, D, Bodnar, R, Evans, B, Dove, P, Bonnell, LM, Marder, MP and Pyrak-Nolte, L (2019) The role of chemistry in fracture pattern development and opportunities to advance interpretations of geological materials. Reviews of Geophysics 57, 1065–111.CrossRefGoogle Scholar
Lee, YJ and Morse, JW (1999) Calcite precipitation in synthetic veins: implications for the time and fluid volume necessary for vein filling. Chemical Geology 156, 151–70.CrossRefGoogle Scholar
Lefticariu, L, Perry, EC, Fischer, MP and Banner, JL (2005) Evolution of fluid compartmentalization in a detachment fold complex. Geology 33, 6972.CrossRefGoogle Scholar
Legeay, E, Ringenbach, JC, Kergaravat, C, Pichat, A, Mohn, G, Vergés, J, Sevki Kavak, K and Callot, JP (2020) Structure and kinematics of the Central Sivas Basin (Turkey): salt deposition and tectonics in an evolving fold-and-thrust belt. In Fold and Thrust Belts: Structural Style, Evolution and Exploration (ed. JA Hammerstein), pp. 361–96. Geological Society of London, Special Publication no. 490CrossRefGoogle Scholar
Lin, Y, Jochum, KP, Scholz, D, Hoffmann, DL, Stoll, B, Weis, U and Meinrat, OA (2017) In-situ high spatial resolution LA-MC-ICPMS 230Th/U dating enables detection of small-scale age inversions in speleothems. Solid Earth Sciences 2, 19.CrossRefGoogle Scholar
Lloyd, MK, Ryb, U and Eiler, JM (2018) Experimental calibration of clumped isotope reordering in dolomite. Geochimica et Cosmochimica Acta 242, 120.CrossRefGoogle Scholar
Lucca, A, Storti, F, Balsamo, F, Clemenzi, L, Fondriest, M, Burgess, R and Di Toro, G (2019) From submarine to subaerial out-of-sequence thrusting and gravity-driven extensional faulting: Gran Sasso Massif, central Apennines, Italy. Tectonics 38, 4155–84.CrossRefGoogle Scholar
Lynch, EA, Mulch, A, Yonkee, A and van der Pluijm, B (2019) Surface fluids in the evolving Sevier fold–thrust belt of ID–WY indicated by hydrogen isotopes in dated, authigenic clay minerals. Earth and Planetary Science Letters 513, e2021GC009868.CrossRefGoogle Scholar
Lynch, EA, Pană, D and van der Pluijm, BA (2021) Focusing fluids in faults: evidence from stable isotopic studies of dated clay-rich fault gouge of the Alberta Rockies. Geochemistry, Geophysics, Geosystems 22, e2021GC009868.CrossRefGoogle Scholar
Lynch, EA and van der Pluijm, B (2017) Meteoric fluid infiltration in the Argentine Precordillera fold-and-thrust belt: evidence from H isotopic studies of neoformed clay minerals. Lithosphere 9, 134–45.CrossRefGoogle Scholar
Lyons, JB and Snellenburg, J (1971) Dating faults. GSA Bulletin 82, 1749–52.CrossRefGoogle Scholar
MacDonald, JM, Faithfull, JW, Roberts, NMW, Davies, AJ, Holdsworth, CM. Newton, M, Williamson, S, Boyce, A and John, CM (2019) Clumped-isotope palaeothermometry and LA-ICP-MS U–Pb dating of lava-pile hydrothermal calcite veins. Contributions to Mineralogy and Petrology 174, 115.CrossRefGoogle Scholar
MacDonald, JM, John, CM and Girard, JP (2018) Testing clumped isotopes as a reservoir characterization tool: a comparison with fluid inclusions in a dolomitized sedimentary carbonate reservoir buried to 2–4 km. In From Source to Seep: Geochemical Applications of Hydrocarbon Systems (ed. M Lawson), pp. 189202. Geological Society of London, Special Publication no. 468.Google Scholar
Machel, HG (1985a) Cathodoluminescence in calcite and dolomite and its chemical interpretation. Geoscience Canada 12, 139–47.Google Scholar
Machel, HG (1985b) Fibrous gypsum and fibrous anhydrite in veins. Sedimentology 32, 443–54.CrossRefGoogle Scholar
Machel, HG (1997) Recrystallization versus neomorphism, and the concept of “significant recrystallization” in dolomite research. Sedimentary Geology 113, 161–8.CrossRefGoogle Scholar
Machel, HG and Cavell, PA (1999) Low-flux, tectonically-induced squeegee fluid flow (“hot flash”) into the Rocky Mountain Foreland Basin. Bulletin of Canadian Petroleum Geology 47, 510–33.Google Scholar
Machel, HG and Lonnee, J (2002) Hydrothermal dolomite – a product of poor definition and imagination. Sedimentary Geology 152, 163–71.CrossRefGoogle Scholar
Mangenot, X, Bonifacie, M, Gasparrini, M, Götz, A, Chaduteau, C, Ader, M and Rouchon, V (2017) Coupling Δ47 and fluid inclusion thermometry on carbonate cements to precisely reconstruct the temperature, salinity and δ18O of paleo-groundwater in sedimentary basins. Chemical Geology 472, 4457.CrossRefGoogle Scholar
Mangenot, X, Gasparrini, M, Rouchon, V and Bonifacie, M (2018) Basin-scale thermal and fluid flow histories revealed by carbonate clumped isotopes (Δ47) – Middle Jurassic carbonates of the Paris Basin depocentre. Sedimentology 65, 123–50.CrossRefGoogle Scholar
Marshak, S, Karlstom, K and Timmons, JM (2000) Inversion of Proterozoic extensional faults: an explanation for the pattern of Laramide and Ancestral Rockies intracratonic deformation, United States. Geology 28, 735–8.2.0.CO;2>CrossRefGoogle Scholar
Martín-Martín, JD, Gomez-Rivas, E, Bover-Arnal, T, Travé, A, Salas, R, Moreno-Bedmar, JA, Tomás, S, Corbella, M, Teixell, A, Vergés, J and Stafford, SL (2013) The Upper Aptian to Lower Albian syn-rift carbonate succession of the southern Maestrat Basin (Spain): facies architecture and fault-controlled stratabound dolostones. Cretaceous Research 41, 217–36.CrossRefGoogle Scholar
Martín-Martín, JD, Gomez-Rivas, E, Gómez-Gras, D, Travé, A, Ameneiro, R, Koehn, D and Bons, PD (2018) Activation of stylolites as conduits for overpressured fluid flow in dolomitized platform carbonates. Geological Society of London, Special Publications 459, 15776.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 (1988) Deep fluid circulation in fault zones. Geology 16, 867–70.2.3.CO;2>CrossRefGoogle Scholar
McCaig, AM, Tritlla, J and Banks, DA (2000) Fluid flow patterns during Pyrenean thrusting. Journal of Geochemical Exploration 69–70, 539–43.CrossRefGoogle Scholar
Merino, E and Canals, À (2011) Self-accelerating dolomite-for-calcite replacement: self-organized dynamics of burial dolomitization and associated mineralization. American Journal of Science 311, 573607.CrossRefGoogle Scholar
Micarelli, L, Benedicto, A and Wibberley, CAJ (2006) Structural evolution and permeability of normal fault zones in highly porous carbonate rocks. Journal of Structural Geology 28, 1214–27.CrossRefGoogle Scholar
Michael, K and Bachu, S (2001) Fluids and pressure distributions in the foreland-basin succession in the west-central part of the Alberta Basin, Canada: evidence for permeability barriers and hydrocarbon generation and migration. AAPG Bulletin 85, 1231–52.Google Scholar
Minissale, A, Magro, G, Martinelli, G, Vaselli, O and Tassi, GF (2000) Fluid geochemical transect in the Northern Apennines (central-northern Italy): fluid genesis and migration and tectonic implications. Tectonophysics 319, 199222.CrossRefGoogle Scholar
Missenard, Y, Bertrand, A, Vergely, P, Benedicto, A, Cushing, ME and Rocher, M (2014) Fracture-fluid relationships: implications for the sealing capacity of clay layers – insights from field study of the Blue Clay formation, Maltese islands. Bulletin de la Société Géologique de France 185, 5163.CrossRefGoogle Scholar
Moore, J, Beinlich, A, Porter, JK, Talavera, C, Berndt, J, Piazolo, S, Austrheim, H and Putnis, A (2020) Microstructurally controlled trace element (Zr, U–Pb) concentrations in metamorphic rutile: an example from the amphibolites of the Bergen Arcs. Journal of Metamorphic Geology 38, 103–27.CrossRefGoogle Scholar
Morad, S, Al-Aasm, IS, Sirat, M and Sattar, MM (2010) Vein calcite in Cretaceous carbonate reservoirs of Abu Dhabi: record of origin of fluids and diagenetic conditions. Journal of Geochemical Exploration 106, 156–70.CrossRefGoogle Scholar
Moragas, M, Martínez, C, Baqués, V, Playà, E, Travé, A, Alías, G and Cantarero, I (2013) Diagenetic evolution of a fractured evaporite deposit (Vilobí Gypsum Unit, Miocene, NE Spain). Geofluids 13, 180–93.CrossRefGoogle Scholar
Motte, G, Hoareau, G, Callot, JP, Revillon, S, Piccoli, F, Calassou, S and Gaucher, EC (2021) Rift and salt-related multi-phase dolomitization: example from the northwestern Pyrenees. Marine and Petroleum Geology 126, 104932.CrossRefGoogle Scholar
Mottram, CM, Kellett, DA, Barresi, T, Zwingmann, H, Friend, M, Todd, A and Percival, JB (2020) Syncing fault rock clocks: direct comparison of U-Pb carbonate and K-Ar illite fault dating methods. Geology 48, 1179–83.CrossRefGoogle Scholar
Mountjoy, EW, Qing, H and McNutt, RH (1992) Strontium isotopic composition of Devonian dolomites, Western Canada Sedimentary Basin: significance of sources of dolomitizing fluids. Applied Geochemistry 7, 5975.CrossRefGoogle Scholar
Mourgues, R and Cobbold, PR (2003) Some tectonic consequences of fluid overpressures and seepage forces as demonstrated by sandbox modelling. Tectonophysics 376, 7597.CrossRefGoogle Scholar
Mouthereau, F, Lacombe, O and Vergés, J (2012) Building the Zagros collisional orogen: timing, strain distribution and the dynamics of Arabia/Eurasia plate convergence. Tectonophysics 532–535, 2760.CrossRefGoogle Scholar
Mouthereau, F, Tensi, J, Bellahsen, N, Lacombe, O, de Boisgrollier, T and Kargar, S (2007) Tertiary sequence of deformation in a thin-skinned/thick-skinned collision belt: the Zagros Folded Belt (Fars, Iran). Tectonics 26, TC5006.CrossRefGoogle Scholar
Mozafari, M, Swennen, R, Balsamo, F, Clemenzi, L, Storti, F, El Desouky, H, Vanhaecke, F, Tueckmantel, C, Solum, J and Taberner, C (2015) Palofluid evolution in fault-damage zones: evidence from fault-fold interaction events in the Jabal Qusaybah Anticline (Adam Foothills, North Oman). Journal of Sedimentary Research 85, 1525–51.CrossRefGoogle Scholar
Mozafari, M, Swennen, R, Balsamo, F, El Desouky, H, Storti, F and Taberner, C (2019) Fault-controlled dolomitization in the Montagna dei Fiori Anticline (Central Apennines, Italy): record of a dominantly pre-orogenic fluid migration. Solid Earth 10, 1355–83.CrossRefGoogle Scholar
Mozafari, M, Swennen, R, Muchez, P, Vassilieva, E, Balsamo, F, Storti, F, Pironon, J and Taberner, C (2017) Origin of the saline paleofluids in fault-damage zones of the Jabal Qusaybah Anticline (Adam Foothills, Oman): constraints from fluid inclusions geochemistry. Marine and Petroleum Geology 86, 537–46.CrossRefGoogle Scholar
Muñoz-López, D, Cruset, D, Vergés, J, Cantarero, I, Benedicto, A, Mangenot, X, Albert, R, Gerdes, A, Beranoaguirre, A and Travé, A (2022) Spatio-temporal variation of fluid flow behavior along a fold: the Bóixols-Sant Corneli anticline (Southern Pyrenees) from U–Pb dating and structural, petrographic and geochemical constraints. Marine and Petroleum Geology 143, 105788.CrossRefGoogle Scholar
Nardini, N, Muñoz-López, D, Cruset, D, Cantarero, I, Martín-Martín, JD, Benedicto, A, Gomez-Rivas, E, John, CM and Travé, A (2019) From early contraction to post-folding fluid evolution in the frontal part of the Bóixols Thrust Sheet (Southern Pyrenees) as revealed by the texture and geochemistry of calcite cements. Minerals 9, 117.CrossRefGoogle Scholar
Neely, TG and Erslev, EA (2009) The interplay of fold mechanisms and basement weaknesses at the transition between Laramide basement-involved arches, north-central Wyoming, USA. Journal of Structural Geology 31, 1012–27.CrossRefGoogle Scholar
Nelson, RA (1981) Significance of fracture sets associated with stylolite zones. AAPG Bulletin 65, 2417–25.Google Scholar
Nesbitt, BE and Muehlenbachs, K (1995) Geochemical studies of the origins and effects of synorogenic crustal fluids in the southern Omineca Belt of British Columbia, Canada. GSA Bulletin 107, 1033–50.2.3.CO;2>CrossRefGoogle Scholar
Nooitgedacht, CW, van der Lubbe, HJL, de Graaf, S, Ziegler, M, Staudigel, PT and Reijmer, JJG (2021) Restricted internal oxygen isotope exchange in calcite veins: constraints from fluid inclusion and clumped isotope-derived temperatures. Geochimica et Cosmochimica Acta 297, 2439.CrossRefGoogle Scholar
Oliver, J (1986) Fluids expelled tectonically from orogenic belts: their role in hydrocarbon migration and other geologic phenomena. Geology 14, 99102.2.0.CO;2>CrossRefGoogle Scholar
Oliver, NHS and Bons, PD (2001) Mechanisms of fluid flow and fluid-rock interaction in fossil metamorphic hydrothermal systems inferred from vein-wall rock patterns, geometry and microstructure. Geofluids 1, 137–62.CrossRefGoogle Scholar
Ortega, OJ, Gale, JFW and Marrett, R (2010) Quantifying diagenetic and stratigraphic controls on fracture intensity in platform carbonates: an example from the Sierra Madre Oriental, northeast Mexico. Journal of Structural Geology 32, 1943–59.CrossRefGoogle Scholar
Ortiz, JP, Person, MA, Mozley, PS, Evans, JP and Bilek, SL (2019) The role of fault-zone architectural elements on pore pressure propagation and induced seismicity. Groundwater 57, 465–78.CrossRefGoogle ScholarPubMed
Pagel, M, Bonifacie, M, Schneider, DA, Gautheron, C, Brigaud, B, Calmels, D, Cros, A, Saint-Bezar, B, Landrein, P, Sutcliffe, C, Davis, D and Chaduteau, C (2018) Improving paleohydrological and diagenetic reconstructions in calcite veins and breccia of a sedimentary basin by combining Δ47 temperature, δ18O water and U-Pb age. Chemical Geology 481, 117.CrossRefGoogle Scholar
Pană, DI and van der Pluijm, BA (2015) Orogenic pulses in the Alberta Rocky Mountains: radiometric dating of major faults and comparison with the regional tectono-stratigraphic record. GSA Bulletin 127, 480502.Google Scholar
Paquette, J and Reeder, RJ (1995) Relationship between surface structure, growth mechanism, and trace element incorporation in calcite. Geochimica et Cosmochimica Acta 59, 735–49.CrossRefGoogle Scholar
Parekh, PP, Möller, P, Dulski, P and Bausch, WM (1977) Distribution of trace elements between carbonate and non-carbonate phases of limestone. Earth and Planetary Science Letters 34, 3950.CrossRefGoogle Scholar
Parry, WT, Chan, MA and Beitler, B (2004) Chemical bleaching indicates episodes of fluid flow in deformation bands in sandstone. AAPG Bulletin 88, 175–91.CrossRefGoogle Scholar
Peckmann, J and Thiel, V (2004) Carbon cycling at ancient methane-seeps. Chemical Geology 205, 443–67.CrossRefGoogle Scholar
Pedrosa, ET, Boeck, L, Putnis, CV and Putnis, A (2017) The replacement of a carbonate rock by fluorite: kinetics and microstructure. American Mineralogist 102, 126–34.CrossRefGoogle Scholar
Petrus, K and Szymczak, P (2016) Influence of layering on the formation and growth of solution pipes. Frontiers in Physics 3, 113.CrossRefGoogle Scholar
Pfiffner, OA. (2017) Thick-skinned and thin-skinned tectonics: a global perspective. Geosciences 7, 71.CrossRefGoogle Scholar
Pichat, A, Hoareau, G, Callot, J-P, Legeay, E, Kavak, KS, Révillon, S, Parat, C and Ringenbach, J-C (2018) Evidence of multiple evaporite recycling processes in a salt-tectonic context, Sivas Basin, Turkey. Terra Nova 30, 40–9.CrossRefGoogle Scholar
Pichat, A, Hoareau, G, Callot, J-P and Ringenbach, J-C (2016) Diagenesis of Oligocene continental sandstones in salt-walled mini-basins – Sivas Basin, Turkey. Sedimentary Geology 339, 1331.CrossRefGoogle Scholar
Pik, R and Marty, B (2009) Helium isotopic signature of modern and fossil fluids associated with the Corinth rift fault zone (Greece): implication for fault connectivity in the lower crust. Chemical Geology 266, 6775.CrossRefGoogle Scholar
Pironon, J and Bourdet, J (2008) Petroleum and aqueous inclusions from deeply buried reservoirs: experimental simulations and consequences for overpressure estimates. Geochimica et Cosmochimica Acta 72, 4916–28.CrossRefGoogle Scholar
Qing, H and Mountjoy, E (1992) Large-scale fluid flow in the Middle Devonian Presqu’ile barrier, western Canada sedimentary basin. Geology 20, 903–6.Google Scholar
Ramsey, DW and Onasch, CM (1999) Fluid migration in a cratonic setting: the fluid histories of two fault zones in the eastern midcontinent. Tectonophysics 305, 307–23.CrossRefGoogle Scholar
Rasbury, ET and Cole, JM (2009) Directly dating geologic events: U-Pb dating of carbonates. Reviews of Geophysics 47, RG3001.CrossRefGoogle Scholar
Reeder, RJ and Grams, JC (1987) Sector zoning in calcite cement crystals: implications for trace element distributions in carbonates. Geochimica et Cosmochimica Acta 51, 187–94.CrossRefGoogle Scholar
Regnet, JB, David, C, Robion, P and Menéndez, B (2019) Microstructures and physical properties in carbonate rocks: a comprehensive review. Marine and Petroleum Geology 103, 366–76.CrossRefGoogle Scholar
Roberts, NMW, Drost, K, Horstwood, MSA, Condon, DJ, Chew, D, Drake, H, Milodowski, AE, McLean, NM, Smye, AJ, Walker, RJ, Haslam, R, Hodson, K, Imber, J, Beaudoin, N and Lee, JK (2020) Laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) U–Pb carbonate geochronology: strategies, progress, and limitations. Geochronology 2, 3361.CrossRefGoogle Scholar
Roberts, NMW and Holdsworth, RE (2022) Timescales of faulting through calcite geochronology: a review. Journal of Structural Geology 158, 104578.CrossRefGoogle Scholar
Roberts, NMW, Žák, J, Vacek, F and Sláma, J (2021) No more blind dates with calcite: fluid-flow vs. fault-slip along the Očkov thrust, Prague Basin. Geoscience Frontiers 12, 101143.CrossRefGoogle Scholar
Roedder, E (1984) Fluid Inclusions: An Introduction to Studies of All Types of Fluid Inclusions, Gas, Liquid, or Melt, Trapped in Materials from Earth and Space, and Their Application to the Understanding of Geologic Processes, Volume 12, 644 pp. Washington, DC: Mineralogical Society of America.Google Scholar
Rosasco, G, Roedder, E and Simmons, JH (1975) Laser-excited Raman spectroscopy for nondestructive partial analysis of individual phases in fluid inclusions in minerals. Science 190, 557–60.CrossRefGoogle Scholar
Roure, F, Andriessen, P, Callot, JP, Faure, JL, Ferket, H, Gonzales, E, Guilhaumou, N, Lacombe, O, Malandain, J, Sassi, W, Schneider, F, Swennen, R and Vilasi, N (2010) The use of palaeo-thermo-barometers and coupled thermal, fluid flow and pore-fluid pressure modelling for hydrocarbon and reservoir prediction in fold and thrust belts. In Hydrocarbons in Contractional Belts (eds Goffey, G, Craig, J, Needham, T and Scott, R), pp. 87114. Geological Society of London, Special Publication no. 348.Google Scholar
Roure, F, Swennen, R, Schneider, F, Faure, JL, Ferket, H, Guilhaumou, N, Osadetz, K, Robion, P and Vandeginste, V (2005) Incidence and importance of tectonics and natural fluid migration on reservoir evolution in foreland fold-and-thrust belts. Oil & Gas Science and Technology 60, 67106.CrossRefGoogle Scholar
Rye, DM and Bradbury, HJ (1988) Fluid flow in the crust: an example from a Pyrenean thrust ramp. American Journal of Science 288, 197235.CrossRefGoogle Scholar
Sachau, T, Bons, PD and Gomez-Rivas, E (2015) Transport efficiency and dynamics of hydraulic deformation networks. Frontiers in Physics 3, 113.CrossRefGoogle Scholar
Sassi, W, Guiton, MLE, Leroy, YM, Daniel, JM and Callot, JP (2012) Constraints on bed scale fracture chronology with a FEM mechanical model of folding: the case of Split Mountain (Utah, USA). Tectonophysics 576–577, 197215.CrossRefGoogle Scholar
Scheiber, T, Viola, G, van der Lelij, R, Margreth, A and Schönenberger, J (2019) Microstructurally-constrained versus bulk fault gouge K-Ar dating. Journal of Structural Geology 127, 103868.CrossRefGoogle Scholar
Schneider, F (2003) Basin modeling in complex area: examples from eastern Venezuelan and Canadian Foothills. Oil & Gas Science and Technology 58, 313–24.CrossRefGoogle Scholar
Scisciani, V, Agostini, S, Calamita, F, Pace, P, Cilli, A, Giori, I and Paltrinieri, W (2014) Positive inversion tectonics in foreland fold-and-thrust belts: a reappraisal of the Umbria-Marche Northern Apennines (Central Italy) by integrating geological and geophysical data. Tectonophysics 637, 218–37.CrossRefGoogle Scholar
Scisciani, V, Patruno, S, Tavarnelli, E, Calamita, F, Pace, P and Iacopini, D (2019) Multi-phase reactivations and inversions of Paleozoic–Mesozoic extensional basins during the Wilson cycle: case studies from the North Sea (UK) and the Northern Apennines (Italy). In Fifty Years of the Wilson Cycle Concept in Plate Tectonics (eds RW Wilson, GA Houseman, KJW McCaffrey, AG Dore and SJH Butler), pp. 205–43. Geological Society of London, Special Publication no. 470.CrossRefGoogle Scholar
Segnit, ER, Holland, HD and Biscardi, CJ (1962). The solubility of calcite in aqueous solutions – I. The solubility of calcite in water between 75° and 200° at CO2 pressures up to 60 atm. Geochimica et Cosmochimica Acta 26, 1301–31.CrossRefGoogle Scholar
Sheppard, SMF (1986) Characterization and isotopic variations in natural waters. Reviews in Mineralogy, 16, 165–83.Google Scholar
Shields, GA and Webb, GE (2004) Has the REE composition of seawater changed over geological time? Chemical Geology 204, 103–7.CrossRefGoogle Scholar
Sibson, RH (1990) Conditions for fault-valve behaviour. In Deformation Mechanisms, Rheology and Tectonics (ed. Knipe, RJ), pp. 1528. Geological Society of London, Special Publication no. 54.Google Scholar
Sibson, RH (2000) Fluid involvement in normal faulting. Journal of Geodynamics 29, 469–99.CrossRefGoogle Scholar
Sibson, RH (2005) Hinge-parallel fluid flow in fold-thrust belts: how widespread? Proceedings of the Geologists’ Association 116, 301–9.CrossRefGoogle Scholar
Sibuet, JC, Srivastava, SP and Spakman, W (2004) Pyrenean orogeny and plate kinematics. Journal of Geophysical Research: Solid Earth 109, 118.CrossRefGoogle Scholar
Smeraglia, L, Aldega, L, Bernasconi, SM, Billi, A, Boschi, C, Caracausi, A, Carminati, E, Franchini, S, Rizzo, AL, Rossetti, F and Vignaroli, G (2020b) The role of trapped fluids during the development and deformation of a carbonate/shale intra-wedge tectonic mélange (Mt. Massico, Southern Apennines, Italy). Journal of Structural Geology 138, 104086.CrossRefGoogle Scholar
Smeraglia, L, Bernasconi, SM, Berra, F, Billi, A, Boschi, C, Caracausi, A, Carminati, E, Castorina, F, Doglioni, C, Italiano, F, Rizzo, AL, Uysal, IT and Zhao, JX (2018) Crustal-scale fluid circulation and co-seismic shallow comb-veining along the longest normal fault of the central Apennines, Italy. Earth and Planetary Science Letters 498, 152–68.CrossRefGoogle Scholar
Smeraglia, L, Fabbri, O, Choulet, F, Buatier, M, Boulvais, P, Bernasconi, SM and Castorina, F (2019) Syntectonic fluid flow and deformation mechanisms within the frontal thrust of foreland fold-and-thrust belt: example from the Internal Jura, Eastern France. Tectonophysics 778, 228178.Google Scholar
Smeraglia, L, Fabbri, O, Choulet, F, Buatier, M, Boulvais, P, Bernasconi, SM and Castorina, F (2020a) Syntectonic fluid flow and deformation mechanisms within the frontal thrust of a foreland fold-and-thrust belt: example from the Internal Jura, Eastern France. Tectonophysics 778, 228178.CrossRefGoogle Scholar
Smeraglia, L, Giuffrida, A, Grimaldi, S, Pullen, A, La Bruna, V, Billi, A and Agosta, F (2021) Fault-controlled upwelling of low-T hydrothermal fluids tracked by travertines in a fold-and-thrust belt, Monte Alpi, Southern Apennines, Italy. Journal of Structural Geology 144, 104276.CrossRefGoogle Scholar
Stel, H (2009) Diagenetic crystallization and oxidation of siderite in red bed (Buntsandstein) sediments from the Central Iberian Chain, Spain. Sedimentary Geology 213, 8996.CrossRefGoogle Scholar
Stolper, DA and Eiler, JM (2015) The kinetics of solid-state isotope-exchange reactions for clumped isotopes: a study of inorganic calcites and apatites from natural and experimental samples. American Journal of Science 315, 363411.CrossRefGoogle Scholar
Stone, DS (1967) Theory of Paleozoic oil and gas accumulation in Bighorn Basin, Wyoming. AAPG Bulletin 51, 2056–114.Google Scholar
Su, A, Chen, H, Feng, Y-X, Zhao, JX, Wang, Z, Hu, M, Jiang, H and Duc Nguyen, A (2022) In situ U-Pb dating and geochemical characterization of multi-stage dolomite cementation in the Ediacaran Dengying Formation, Central Sichuan Basin, China: constraints on diagenetic, hydrothermal and paleo-oil filling events. Precambrian Research 368, 106481.CrossRefGoogle Scholar
Sutherland, R, Toy, VG, Townend, J, Cox, SC, Eccles, JD, Faulkner, DR, Prior, DJ, Norris, RJ, Mariani, E, Boulton, C, Carpenter, BM, Menzies, CD, Little, TA, Hasting, M, De Pascale, GP, Langridge, RM, Scott, HR, Reid Lindroos, Z, Fleming, B and Kopf, J (2012) Drilling reveals fluid control on architecture and rupture of the Alpine Fault, New Zealand. Geology 40, 1143–6.CrossRefGoogle Scholar
Swanson, EM, Wernicke, BP, Eiler, JM and Losh, S (2012) Temperatures and fluids on faults based on carbonate clumped-isotope thermometry. American Journal of Science 312, 121.CrossRefGoogle Scholar
Swart, PK (2015) The geochemistry of carbonate diagenesis: the past, present and future. Sedimentology 62, 1233–304.CrossRefGoogle Scholar
Swart, PK, Burns, SJ and Leder, JJ (1991) Fractionation of the stable isotopes of oxygen and carbon in carbon dioxide during the reaction of calcite with phosphoric acid as a function of temperature and technique. Chemical Geology: Isotope Geoscience section 86, 8996.Google Scholar
Swennen, R, Muskha, K and Roure, F (2000) Fluid circulation in the Ionian fold and thrust belt (Albania): implications for hydrocarbon prospectivity. 70, 629–34.CrossRefGoogle Scholar
Szymczak, P and Ladd, AJC (2009) Wormhole formation in dissolving fractures. Journal of Geophysical Research: Solid Earth 114, 122.CrossRefGoogle Scholar
Szymczak, P and Ladd, AJC (2014) Reactive-infiltration instabilities in rocks. Part 2. Dissolution of a porous matrix. Journal of Fluid Mechanics 738, 591630.CrossRefGoogle Scholar
Tang, J, Dietzel, M, Böhm, F, Köhler, SJ and Eisenhauer, A (2008) Sr2+/Ca2+ and 44Ca/40Ca fractionation during inorganic calcite formation: II. Ca isotopes. Geochimica et Cosmochimica Acta 72, 3733–45.CrossRefGoogle Scholar
Tang, J, Dietzel, M, Fernandez, A, Tripati, AK and Rosenheim, BE (2014) Evaluation of kinetic effects on clumped isotope fractionation (Δ47) during inorganic calcite precipitation. Geochimica et Cosmochimica Acta 134, 120–36.CrossRefGoogle Scholar
Tavani, S, Storti, F, Lacombe, O, Corradetti, A, Muñoz, JA and Mazzoli, S (2015) A review of deformation pattern templates in foreland basin systems and fold-and-thrust belts: implications for the state of stress in the frontal regions of thrust wedges. Earth-Science Reviews 141, 82104.CrossRefGoogle Scholar
Thacker, JO and Karlstrom, KE (2019) U-Pb dating of calcite veins reveals complex stress evolution and thrust sequence in the Bighorn Basin, Wyoming, USA. Geology 47, e480.CrossRefGoogle Scholar
Tondi, E, Antonellini, M, Aydin, A, Marchegiani, L and Cello, G (2006) The role of deformation bands, stylolites and sheared stylolites in fault development in carbonate grainstones of Majella Mountain, Italy. Journal of Structural Geology 28, 376–91.CrossRefGoogle Scholar
Tonguç Uysal, I, Zhao, J-X, Golding, SD, Lawrence, MG, Glikson, M and Collerson, KD (2007) Sm-Nd dating and rare-earth element tracing of calcite: implications for fluid-flow events in the Bowen Basin, Australia. Chemical Geology 238, 6371.CrossRefGoogle Scholar
Tostevin, R, Bradbury, HJ, Shields, GA, Wood, RA, Bowyer, F, Penny, AM and Turchyn, AV (2019) Calcium isotopes as a record of the marine calcium cycle versus carbonate diagenesis during the late Ediacaran. Chemical Geology 529, 119319.CrossRefGoogle Scholar
Toussaint, R, Aharonov, E, Koehn, D, Gratier, JP, Ebner, M, Baud, P, Rolland, A and Renard, F (2018) Stylolites: a review. Journal of Structural Geology 114, 163–95.CrossRefGoogle Scholar
Travé, A, Calvet, F, Sans, M, Verges, J and Thirlwall, M (2000) Fluid history related to the Alpine compression at the margin of the south-Pyrenean Foreland basin: the El Guix anticline. Tectonophysics 321, 73102.CrossRefGoogle Scholar
Travé, A, Labaume, P, Calvet, F and Soler, A (1997) Sediment dewatering and pore fluid migration along thrust faults in a foreland basin inferred from isotopic and elemental geochemical analyses (Eocene southern Pyrenees, Spain). Tectonophysics 282, 375–98.CrossRefGoogle Scholar
Travé, A, Labaume, P and Vergés, J (2007) Fluid systems in foreland fold-and-thrust belts: an overview from the Southern Pyrenees. In Thrust Belts and Foreland Basins. Frontiers in Earth Sciences (eds O Lacombe, F Roure, J Lavé and J Vergés) Heidelberg: Springer, Berlin. https://doi.org/10.1007/978-3-540-69426-7_5.Google Scholar
Uysal, IT, Feng, YX, Zhao, JX, Isik, V, P, Nuriel and SD, Golding (2009). Hydrothermal CO2 degassing in seismically active zones during the late Quaternary. Chemical Geology 265, 442–54.CrossRefGoogle Scholar
van der Pluijm, BA, Craddock, JP, Graham, BR and Harris, JH (1997) Paleostress in cratonic North America: implications for deformation of continental interiors. Science 277, 794–6.CrossRefGoogle Scholar
van der Pluijm, BA, Hall, CM, Vrolijk, PJ, Pevear, DR and Covey, MC (2001) The dating of shallow faults in the Earth’s crust. Nature 412, 172–5.CrossRefGoogle ScholarPubMed
van der Pluijm, BA, Vrolijk, PJ, Pevear, DR, Hall, CM and Solum, J (2006) Fault dating in the Canadian Rocky Mountains: evidence for late Cretaceous and early Eocene orogenic pulses. Geology 34, 837–40.CrossRefGoogle Scholar
van Geet, M, Swennen, R, Durmishi, C, Roure, F and Muchez, PH (2002) Paragenesis of Cretaceous to Eocene carbonate reservoirs in the Ionian fold and thrust belt (Albania): relation between tectonism and fluid flow. Sedimentology 49, 697718.CrossRefGoogle Scholar
Vandeginste, V, Swennen, R, Allaeys, M, Ellam, RM, Osadetz, K and Roure, F (2012) Challenges of structural diagenesis in foreland fold-and-thrust belts: a case study on paleofluid flow in the Canadian Rocky Mountains West of Calgary. Marine and Petroleum Geology 35, 235–51.CrossRefGoogle Scholar
Vannucchi, P, Remitti, F, Bettelli, G, Boschi, C and Dallai, L (2010) Fluid history related to the early Eocene-middle Miocene convergent system of the Northern Apennines (Italy): constraints from structural and isotopic studies. Journal of Geophysical Research: Solid Earth 115, 5405.CrossRefGoogle Scholar
Vass, A, Koehn, D, Toussaint, R, Ghani, I and Piazolo, S (2014) The importance of fracture-healing on the deformation of fluid-filled layered systems. Journal of Structural Geology 67, 94106.CrossRefGoogle Scholar
Vignaroli, G, Rossetti, F, Petracchini, L, Argante, V, Bernasconi, SM, Brilli, M, Giustini, F, Yu, T-L, Shen, C-C and Soligo, M (2022) Middle Pleistocene fluid infiltration with 10–15 ka recurrence within the seismic cycle of the active Monte Morrone Fault System (central Apennines, Italy) Tectonophysics 827, 229269. doi: 10.1016/j.tecto.2022.229269.CrossRefGoogle Scholar
Vilasi, N, Malandain, J, Barrier, L, Callot, JP, Amrouch, K, Guilhaumou, N, Lacombe, O, Muska, K, Roure, F and Swennen, R (2009) From outcrop and petrographic studies to basin-scale fluid flow modelling: the use of the Albanian natural laboratory for carbonate reservoir characterisation. Tectonophysics 474, 367–92.CrossRefGoogle Scholar
Vilasi, N, Swennen, R and Roure, F (2006) Diagenesis and fracturing of Paleocene–Eocene carbonate turbidite systems in the Ionian Basin: the example of the Kelçyra area (Albania). Journal of Geochemical Exploration 89, 409–13.CrossRefGoogle Scholar
Villemant, B and Feuillet, N (2003) Dating open systems by the 238U–234U–230Th method: application to Quaternary reef terraces. Earth and Planetary Science Letters 210, 105–18.CrossRefGoogle Scholar
Viola, G, Zwingmann, H, Mattila, J and Käpyaho, A (2013) K-Ar illite age constraints on the Proterozoic formation and reactivation history of a brittle fault in Fennoscandia. Terra Nova 25, 236–44.CrossRefGoogle Scholar
Wacker, U, Fiebig, J, Tödter, J, Schöne, BR, Bahr, A, Friedrich, O, Tütken, T, Gischler, E and Joachimski, MM (2014) Empirical calibration of the clumped isotope paleothermometer using calcites of various origins. Geochimica et Cosmochimica Acta 141, 127–44.CrossRefGoogle Scholar
Wang, X, Gao, J, He, S, He, Z, Zhou, Y, Tao, Z, Zhang, J and Wang, Y (2017) Fluid inclusion and geochemistry studies of calcite veins in Shizhu synclinorium, central China: record of origin of fluids and diagenetic conditions. Journal of Earth Science 28, 315–32.CrossRefGoogle Scholar
Weber, J, Cheshire, MC, Bleuel, M, Mildner, D, Chang, YJ, Ievlev, A, Littrell, KC, Ilavsky, J, Stack, AG and Anovitz, LM (2021) Influence of microstructure on replacement and porosity generation during experimental dolomitization of limestones. Geochimica et Cosmochimica Acta 303, 137–58.CrossRefGoogle Scholar
Wennberg, OP, Casini, G, Jonoud, S and Peacock, DCP (2016) The characteristics of open fractures in carbonate reservoirs and their impact on fluid flow: a discussion. Petroleum Geoscience 22, 91104.CrossRefGoogle Scholar
Xu, WG, Fan, HR, Hu, FF, Santosh, M, Yang, KF, Lan, TG and Wen, BJ (2015) Geochronology of the Guilaizhuang gold deposit, Luxi Block, eastern North China Craton: constraints from zircon U–Pb and fluorite-calcite Sm–Nd dating. Ore Geology Reviews 65, 390–9.CrossRefGoogle Scholar
Yan, H, Dreybrodt, W, Bao, H, Peng, Y, Wei, Y, Ma, S, Mo, B, Sun, H and Liu, Z (2021) The influence of hydrodynamics on the carbon isotope composition of inorganically precipitated calcite. Earth and Planetary Science Letters 565, 116932.CrossRefGoogle Scholar
Yardley, BWD and Graham, JT (2002) The origins of salinity in metamorphic fluids. Geofluids 2, 249–56.CrossRefGoogle Scholar
Yonkee, WA and Weil, AB (2015) Tectonic evolution of the Sevier and Laramide belts within the North American Cordillera orogenic system. Earth-Science Reviews 150, 531–93.CrossRefGoogle Scholar
Zaarur, S, Affek, HP and Brandon, MT (2013) A revised calibration of the clumped isotope thermometer. Earth and Planetary Science Letters 382, 4757.CrossRefGoogle Scholar
Zeebe, RE (2007) An expression for the overall oxygen isotope fractionation between the sum of dissolved inorganic carbon and water. Geochemistry, Geophysics, Geosystems 8, Q09002.CrossRefGoogle Scholar
Zheng, Y-F (1999) Oxygen isotope fractionation in carbonate and sulfate minerals. Geochemical Journal 33, 109–26.CrossRefGoogle Scholar
Zuluaga, LF, Rotevatn, A, Keilegavlen, E and Fossen, H (2016) The effect of deformation bands on simulated fluid flow within fault-propagation fold trap types: lessons from the San Rafael monocline, Utah. AAPG Bulletin 100, 1523–40.CrossRefGoogle Scholar
Zwingmann, H, Offler, R, Wilson, T and Cox, SF (2004) K–Ar dating of fault gouge in the northern Sydney Basin, NSW, Australia: implications for the breakup of Gondwana. Journal of Structural Geology 26, 2285–95.CrossRefGoogle Scholar
Figure 0

Fig. 1. Sketches representing a fold-and-thrust belt – foreland basin system (a) and a broken foreland (b). Stress regime, fluid system, source of fluids and engine of migrations, and mesoscale structures related to the folding event, from layer-parallel shortening to late stage of fold tightening, are reported.

Figure 1

Fig. 2. Field and microscope photographs of mesoscale structures studied for past fluid system reconstruction. (a, b) Veins in a fracture network in the Apennines, Italy (a) and in the Laramide province, USA (b). (c, d) Examples of vein textures: blocky calcite from the Apennines, Italy (c); elongated blocky calcite from southern Pyrenees, Spain (d). (e) Multiple fluid flow events and related cement precipitation in a crack seal, Laramide province, USA. (f, g) Cathodoluminescence images of a single phase of cementation in a vein whose borders were affected by dolomitization, Laramide province, USA (f), and of multiple diagenetic phases affecting a vein, Southeast Basin, France (g). Note pressure-solution (stylolite) along the vein border in (g). (h, i) Striated calcite steps along a fault surface, Laramide province USA (h), and fault breccia, associated to local overpressure, Jaca Basin, Spain (i).

Figure 2

Fig. 3. (a) Expected distribution of the mesoscale fracture pattern at the fold scale, including faults, joints and stylolites. Colours show the relation between the feature and the stage of deformation: brown relates to burial, grey to pre-orogenic, pink to outer arc extension during flexure, black to the layer-parallel shortening, green to fold growth, red to late stage of fold tightening. (b–e) Zoom on the expected fluid flow in the structures depicted in (a): (b) matrix-scale reactive fluid migration pathways; (c) migration pathways in joint network; (d) migration pathways in stylolites; (e) migration pathways in fault zones and cores, with the static state and the dynamic state. See text for details and references.

Figure 3

Fig. 4. Concept of geochemistry-assisted structural geology. Left-hand side: representation of all parameters (bold) that can be reconstructed from geochemical analysis (framed) on calcite, organized in a suggested workflow. The suggestion is to pick at least one of the analyses of each box to reconstruct the past fluid system with the least ambiguity. Right-hand side: structural implication of the outcome of the past fluid system reconstruction, including a direct appraisal of the reservoir hydrological structure, and an inferred model of the fluid conduit distribution in time and space.

Figure 4

Fig. 5. Simplified representation of δ18O vs δ13C plots obtained in syn-kinematic calcite cements from the fracture network (a, b, d) and in thrust fault zone (c), along with interpretation of the related fluid system. The published interpretations are reported on each graph. (a) The Sheep Mountain Anticline past fluid system (modified after Beaudoin et al. 2011). (b) The Puig-Reig Anticline past fluid system (modified after Cruset et al. 2016). (c) The Mount Tancia thrust fault zone past fluid system (modified after Curzi et al.,2020). (d) The past fluid system of the Albanide fold-and-thrust belt (modified after de Graaf et al. 2019). Note that the coloured frames in (a) locate the extension of plots (b–d). On all plots, the dashed grey frame represents the isotope values of the calcite fraction of the host rock when available. All values are given in ‰ PDB. T on plots refers to the measured temperatures of a given population, using an independent paleothermometer (T47 for Δ47CO2,Th for fluid inclusion homogenization temperatures). LPS stands for layer-parallel shortening.

Figure 5

Fig. 6. Representation of the first-order evolution of the past fluid system at the scale of the individual fold or/and thrust. From an initial state related to layer-parallel shortening (left-hand side), the fluid system evolved according to the fault distribution/the décollement nature, and to the structural style of the structure. The resulting fluid from which cement precipitates is symbolized by a hexagon whose colour is related to the nature of the fluid, several colours representing a mixing with no implication on the ratio. Fluid temperature with respect to the local temperature is symbolized by C when cooler, H when hotter and E when at equilibrium.

Figure 6

Fig. 7. Simplified cross-section and corresponding location of the Canadian Rocky Mountains fold-and-thrust belt where the past fluid system was reconstructed. The vertical exaggeration is c. 7:1. FTB stands for fold-and-thrust belt. The legend is transferable to the other cross-sections of Figures 8–10. In the present case, the decoupling level comprises argillites (Proterozoic). See text for details and references.

Figure 7

Fig. 8. Simplified cross-section and corresponding location of the Sevier range and Laramide province (USA) where the past fluid system was reconstructed. On the location map, the yellow colour indicates the intracratonic basins related to the exhumation of basement arches. The vertical exaggeration is about 2:1. WRB: Wind-River Basin; BHB: Bighorn Basin; PRB: Powder River Basin. See text for details and references. Key is given in Figure 7.

Figure 8

Fig. 9. Simplified cross-section and corresponding location of the southern Pyrenean FTB (Spain), where the past fluid system was reconstructed along the thrusts. The vertical exaggeration is c. 3:1. See text for details and references. Key is given in Figure 7. The nature of the décollement is evaporites (Triassic).

Figure 9

Fig. 10. Simplified cross-section and corresponding location of Umbria Marche Apennine Ridge (Italy), where the past fluid system was reconstructed. The vertical exaggeration is c. 2:1. See text for details and references. Key is given in Figure 7. The nature of the décollement is evaporites (Triassic).

Figure 10

Fig. 11. Schematic cross-sections in the fold-and-thrust belts with respect to the structural style, inherited structures and rheology of the sediments. The expected order of structural development (thrust activation) is shown as numbers, along with the expected fluid migration at that scale (blue arrows).

Figure 11

Fig. 12. Sketch illustrating the expected first-order evolution of the fluid system in FTB, without considering specificities related to structural style. (a) During layer-parallel shortening; (b) during folding/thrusting.

Figure 12

Fig. 13. Example of direct application of geochemistry-assisted structural geology, based on the case study of the Sheep Mountain Anticline (SMA; Beaudoin et al. 2011). (a) Geological map of the SMA, where the values of the δ18O of tectonic veins are reported as coloured points, and where iso-δ18O value curves are plotted. (b) Simplified δ18O vs δ13C plots of the cements in the fracture network. The colours of the dots and the corresponding frames relate to the deformation phase during which the vein developed. Red and orange vertical lines correspond to the past fluid system interpretation, in accordance with the curves plotted in (a). (c) Cross-section along the line located on (a), with the corresponding location of the iso-δ18O value curves, with a proposed subsurface geometry. (d) Interpretation of the deformation sequence accounting for a potential migration of the hinge during fold growth. (a) and (b) are modified after Beaudoin et al. (2011).