Skip to Main Content
Skip Nav Destination
This is an Open Access article distributed under the terms of the Creative Commons Attribution 4.0 License (http://creativecommons.org/licenses/by/4.0/).

The study of magmatic and metamorphic processes is challenged by geological complexities like geochemical variations, geochronological uncertainties and the presence/absence of fluids and/or melts. However, by integrating petrographic and microstructural studies with geochronology, geochemistry and phase equilibrium diagram investigations of different key mineral phases, it is possible to reconstruct insightful pressure–temperature–deformation–time histories. Using multiple geochronometers in a rock can provide a detailed temporal account of its evolution, as these geological clocks have different closure temperatures. Given the continuous improvement of existing and new in situ analytical techniques, this contribution provides an overview of frequently utilized petrochronometers such as garnet, zircon, titanite, allanite, rutile, monazite/xenotime and apatite, by describing the geological record that each mineral can retain and explaining how to retrieve this information. These key minerals were chosen as they provide reliable age information in a variety of rock types and, when coupled with their trace element (TE) composition, form powerful tools to investigate crustal processes at different scales. This review recommends best applications for each petrochronometer, highlights limitations to be aware of and discusses future perspectives. Finally, this contribution underscores the importance of integrating information retrieved by multi-petrochronometer studies to gain an in-depth understanding of complex thermal and deformation crustal processes.

Unravelling the composite tectono-magmatic and metamorphic evolution of terrains can be challenging because their multistage histories may have erased past information. One way to interrogate and address these complexities is to investigate geological processes throughout Earth's history that are encapsulated in key minerals such as garnet, zircon, titanite, allanite, rutile, monazite/xenotime and apatite. The mineral phases chosen for this contribution are the most used in petrochronological studies as they are reliable chronometers, and their chemical variability has been used to investigate metamorphic and magmatic crustal processes. As petrochronometers, they can retain information about the petrogenesis of their protoliths, (multiple) pressures (P), temperatures (T) and timing (t) at which they (re)crystallized in magmatic, metamorphic and/or hydrothermal events. Investigations employing these minerals have provided new breakthroughs in Earth Sciences. Scientific findings through petrochronological investigations include (but are not limited to) the study of the oldest minerals on Earth (Valley et al. 2014), the understanding of trace element (TE) partitioning processes between different key mineral phases (e.g. Rubatto 2002) and the study of fluid and magma sources during crust evolution (e.g. Valley 2003; Bouvier et al. 2012; Mulder et al. 2021). Magmatic and/or metamorphic overprinting stages are heterogeneously recorded by these petrochronometers, which preserve snapshots of a rock evolution. The timing, conditions of formation and distinct closure temperatures (i.e. Tc as the temperature of a geochronological system at the time corresponding to its apparent age; Dodson 1973) that characterize minerals and isotope systems (Fig. 1) allow, for example, the UPb system in garnet (Mezger et al. 1989a) to retain information at higher temperatures than in rutile, which resets its internal clock as a function of T, grain size and cooling rate. Thus, carrying out multi-mineral studies (Fig. 1) is an incredibly powerful tool to investigate complex magmatic and metamorphic histories (e.g. Cutts et al. 2014; Stearns et al. 2015; Manzotti et al. 2018; Volante et al. 2020c; Fumes et al. 2022; Odlum et al. 2022).

Over the past decades, our ability to interrogate the preserved rock record of crustal metamorphism and magmatism has significantly advanced thanks to continuously improving in situ analytical techniques, measurement methodologies and protocols, as well as availability of reference materials that have been calibrated through worldwide inter-laboratory collaborations. Most commonly used instruments include secondary ion mass spectrometry (SIMS) (e.g. Chamberlain et al. 2010; Ushikubo et al. 2014; Zhou et al. 2016; Chaussidon et al. 2017) and inductively coupled plasma mass spectrometers (single and multi-collector, sector-field, single and triple quadrupoles) associated with a laser ablation system (LA-ICP-MS, LA-MC-ICP-MS or LA-ICP-MS/MS) either used in tandem (split stream) or separately (e.g. Kylander-Clark et al. 2013; Hacker et al. 2015; Simpson et al. 2021). Additionally, improvements in microanalytical techniques, such as (but not limited to) electron backscatter diffraction (EBSD) mapping (e.g. Cavosie et al. 2015), TESCAN integrated mineral analyser (TIMA; e.g. Porter et al. 2020), field emission scanning electron microscope (FESEM; Tacchetto et al. 2022) for qualitative analysis acquisition and field emission gun (FEG) transmission electron microscope (TEM) for atomic lattice resolution imaging and nanoscale microanalysis (Reddy et al. 2020; Tacchetto et al. 2021), have significantly contributed in advancing the micro- to nanoscale petrographic, chemical, physical and structural characterization of sample material. The integration of continuously updated thermodynamic modelling studies of monazite (e.g. Janots et al. 2007; Kelsey et al. 2008; Spear and Pyle 2010; Hoschek 2016), allanite (Spear 2010), zircon (Kelsey et al. 2008; Kelsey and Powell 2011; Yakymchuk and Brown 2014) and apatite (Spear and Pyle 2010; Yakymchuk 2017) allows for the construction of a fundamental basis from which to overcome problems related to qualitative interpretations of accessory minerals within the P–T space, for which rutile (e.g. Fumes et al. 2022; Holtmann et al. 2022; Horton et al. 2022; Vanardois et al. 2022) and titanite (e.g. Kapp et al. 2009; Kirkland et al. 2016, 2020; Apen et al. 2020; Walters et al. 2022) already greatly contribute. The above-mentioned analytical advances combined with more traditional petro-structural (micro)analyses (Volante et al. 2020b) and continuously updated thermodynamic datasets used for modelling pressure (P), temperature (T) and composition (X) are continuously contributing to improve our understanding of crustal processes by allowing a more detailed characterization of mineral phases.

This review aims to showcase snapshots of recent analytical developments such as advancement in understanding the capabilities of LA-ICP-MS/MS and their applications, including new findings in the uptake of water in zircon crystal lattice and its influence on oxygen isotope composition, development and improvement in generating in situ quantitative elemental maps that can be directly linked with petrochronological and textural information using key mineral phases in igneous and metamorphic studies. In particular, we focus on the most employed minerals in the literature that can be used not only as chronometers in a great variety of rock lithologies but also as tracers of fluid–rock interactions at crustal conditions, and they include: garnet, zircon, titanite, rutile, allanite, monazite/xenotime and apatite (Fig. 2). In the following section, each mineral is described including its potential as chronometer, thermobarometer and chemical/isotopic tracer applied to crustal magmatic and metamorphic processes. Ultimately, the Discussion section provides examples of applications that use these petrochronometers in magmatic and metamorphic studies with a focus on the complementary set of information obtained when multi-mineral investigations are used to unravel different steps in the evolution of a volume of rock. Future perspectives on the use of these key mineral phases to investigate metamorphic and igneous processes are also included.

Fig. 2.

Representative textures for the key mineral phases described in this contribution. (a) Plane polarized light. (b) Backscattered electron (BSE) and (c) electron microprobe Mn elemental map images showing the early and late stages of fabric development in core and rims, respectively, during garnet growth. (d) Plane polarized light, (e) cathodoluminescence (CL) and (f) laser ablation ICP–MS U (ppm) elemental map images of zircon. (g) Plane polarized light, (h) BSE and (i) quantitative map of Al (a.p.f.u., atom per formula unit) images of titanite. (j) Plane polarized light, (k) BSE and (l) EBSD map images for rutile. (m) Plane and crossed polarized light, (n) BSE and (o) electron microprobe Th and Y elemental map images of monazite. (p) Plane and crossed polarized, (q) BSE and (r) LA–ICP–MS trace element map images of apatite. (s) Plane polarized light, (t) BSE and (u) X-ray elemental map images of allanite. Source: (c) Volante et al. (2020c); (f) Chew et al. (2017); (i) Walters et al. (2022); (l) Moore et al. (2020a, b); (n) Barrote et al. (2022a); (o) Volante et al. (2020c); (r) Henrichs et al. (2019); (u) Corti et al. (2020).

Fig. 2.

Representative textures for the key mineral phases described in this contribution. (a) Plane polarized light. (b) Backscattered electron (BSE) and (c) electron microprobe Mn elemental map images showing the early and late stages of fabric development in core and rims, respectively, during garnet growth. (d) Plane polarized light, (e) cathodoluminescence (CL) and (f) laser ablation ICP–MS U (ppm) elemental map images of zircon. (g) Plane polarized light, (h) BSE and (i) quantitative map of Al (a.p.f.u., atom per formula unit) images of titanite. (j) Plane polarized light, (k) BSE and (l) EBSD map images for rutile. (m) Plane and crossed polarized light, (n) BSE and (o) electron microprobe Th and Y elemental map images of monazite. (p) Plane and crossed polarized, (q) BSE and (r) LA–ICP–MS trace element map images of apatite. (s) Plane polarized light, (t) BSE and (u) X-ray elemental map images of allanite. Source: (c) Volante et al. (2020c); (f) Chew et al. (2017); (i) Walters et al. (2022); (l) Moore et al. (2020a, b); (n) Barrote et al. (2022a); (o) Volante et al. (2020c); (r) Henrichs et al. (2019); (u) Corti et al. (2020).

Garnet (Fig. 2a-c; Fig. 3) is a major rock-forming mineral, occurring in metamorphic (felsic and mafic compositions) and felsic peraluminous magmatic rocks (Villaros et al. 2009; Bartoli et al. 2013; Liu et al. 2014; Volante et al. 2020a; R. Li et al. 2021). Garnet is a key metamorphic indicator mineral (Barrow 1893, 1912) and its composition directly relates to the P–T conditions it experienced during growth (the garnet stability field extends from greenschist-facies to high-pressure (HP) and high-temperature (HT) conditions, dependent on rock composition), making it extremely useful as a geothermobarometer. Garnet grains can shield mineral inclusions whose identity or composition can indicate earlier assemblages/conditions experienced by the rock (e.g. St-Onge 1987; et al. 2008; Thomas and Davidson 2012; Pourteau et al. 2019; Schönig et al. 2019; Godet et al. 2022). These mineral inclusions can also preserve or trace earlier foliations, which can be used to infer older foliation trends (e.g. Sayab 2006; Aerden et al. 2013, 2021; Sayab et al. 2015, 2016; Volante et al. 2020b). Like other porphyroblasts, garnet can also infer kinematics via grain rotation (Fig. 3a; Passchier and Simpson 1986; Johnson 1999). Garnet can record a prolonged history, preserving a growth hiatus, where textural, chemical or chronological evidence shows that garnet grew during more than one metamorphic and/or magmatic event (Fig. 3b; e.g. Vance et al. 1998; Cutts et al. 2010; Ortolano et al. 2014; Kulhánek et al. 2021; Massonne and Li 2022).

Fig. 3.

(a) Garnet porphyroblast indicating the kinematics of deformation (arrows) as a sigma-clast and exhibiting inclusion trails of an internal foliation. (b) Polymetamorphic garnet grain from Polnish, Scotland (see Vance et al. 1998; Cutts et al. 2009). The core (Z1) and first rim (Z2) zones preserve the same early foliation (S1) but have cleavage domes and graphite inclusions on their border. A thin, inclusion-rich rim (Z3) which has been poorly preserved is observed. (c) Representative chondrite-normalized REE patterns for different types of garnet (Blereau 2017). (d) Partitioning between zircon and garnet: (1) Buick et al. (2006); (2) Hokada and Harley (2004); (3) GL7 in Rubatto (2002); (4) GP7 in Rubatto (2002); (5) Harley et al. (2001); (6) Kelly and Harley (2005); (7) Whitehouse and Platt (2003); (8) Hermann and Rubatto (2003). (e) Partitioning array plot of zircon/garnet compared to the experimental data of Taylor et al. (2015). The axes of this plot are: x, log(DYb); y, log(Dslope) = log(DYb/DGd) (Blereau 2017). (f) Sm–Nd dating of a large garnet porphyroblast by Pollington and Baxter (2010). The various garnet rims were obtained using a microdrill (the drill lines are marked in black on the garnet). Coloured zones correspond to the coloured isochrons and age points in the plots indicate that garnet grew over c. 8 myr, from 28 to 20 Ma. (g) TW diagram showing how the grain size and the amount of ablated U-rich inclusions affect garnet U–Pb dating. (h) Garnet compositional map of Ca, Fe, Mn and Y (from top left) of sample Verpeneset from Tamblyn et al. (2022). Black and white dots represent the position of core and rim analyses, respectively. (i and j) Lu/Hf isochrons obtained from garnet (i) core and (j) rim. Uncertainties consider overdispersion of the data and isochron ages are calculated with age offset (see Tamblyn et al. 2022 for full details). Source: (a) Photo courtesy of J. Pownall; (c–d) modified after Blereau (2017); (f) adapted from Pollington and Baxter (2010), Baxter et al. (2017); (g) modified after Millonig et al. (2020).

Fig. 3.

(a) Garnet porphyroblast indicating the kinematics of deformation (arrows) as a sigma-clast and exhibiting inclusion trails of an internal foliation. (b) Polymetamorphic garnet grain from Polnish, Scotland (see Vance et al. 1998; Cutts et al. 2009). The core (Z1) and first rim (Z2) zones preserve the same early foliation (S1) but have cleavage domes and graphite inclusions on their border. A thin, inclusion-rich rim (Z3) which has been poorly preserved is observed. (c) Representative chondrite-normalized REE patterns for different types of garnet (Blereau 2017). (d) Partitioning between zircon and garnet: (1) Buick et al. (2006); (2) Hokada and Harley (2004); (3) GL7 in Rubatto (2002); (4) GP7 in Rubatto (2002); (5) Harley et al. (2001); (6) Kelly and Harley (2005); (7) Whitehouse and Platt (2003); (8) Hermann and Rubatto (2003). (e) Partitioning array plot of zircon/garnet compared to the experimental data of Taylor et al. (2015). The axes of this plot are: x, log(DYb); y, log(Dslope) = log(DYb/DGd) (Blereau 2017). (f) Sm–Nd dating of a large garnet porphyroblast by Pollington and Baxter (2010). The various garnet rims were obtained using a microdrill (the drill lines are marked in black on the garnet). Coloured zones correspond to the coloured isochrons and age points in the plots indicate that garnet grew over c. 8 myr, from 28 to 20 Ma. (g) TW diagram showing how the grain size and the amount of ablated U-rich inclusions affect garnet U–Pb dating. (h) Garnet compositional map of Ca, Fe, Mn and Y (from top left) of sample Verpeneset from Tamblyn et al. (2022). Black and white dots represent the position of core and rim analyses, respectively. (i and j) Lu/Hf isochrons obtained from garnet (i) core and (j) rim. Uncertainties consider overdispersion of the data and isochron ages are calculated with age offset (see Tamblyn et al. 2022 for full details). Source: (a) Photo courtesy of J. Pownall; (c–d) modified after Blereau (2017); (f) adapted from Pollington and Baxter (2010), Baxter et al. (2017); (g) modified after Millonig et al. (2020).

Garnet is a powerful tool for linking accessory mineral geochronological and/or isotopic data to the evolution of the major mineral assemblage using trace elements. Linking these two data sources together is the cornerstone of petrochronology (Engi et al. 2017). Trace elements such as rare earth elements (REE) partition strongly into garnet (Bea et al. 1994; Hermann and Rubatto 2014), particularly the middle to heavy rare earth elements (M-HREE; Fig. 3c). Other minerals that also readily accommodate M-HREE, such as zircon and monazite, compete with garnet, resulting in modified trace element patterns when the accessory phase grows in the presence or absence of garnet. The use and effectiveness of these trace element partitioning relationships (DREE) were first presented for zircon and garnet (Fig. 3d; Rubatto 2002). Zircon grown in the presence of garnet has a M-HREE slope of c. 1 effectively flat, and a slightly curved to near flat DREE pattern (Whitehouse and Platt 2003; Hokada and Harley 2004; Kelly and Harley 2005; Harley and Kelly 2007; Wu et al. 2008a, b; Fornelli et al. 2014, 2018). Since these first studies, trace element partitioning has grown into a widely used tool in petrochronology (Whitehouse and Platt 2003; Hokada and Harley 2004; Baldwin and Brown 2008; Clark et al. 2009; Harley and Nandakumar 2014), and some studies conducted quantitative partitioning experiments to constrain temperature dependent patterns for DREE (Zrc/Grt) (Rubatto and Hermann 2007; Taylor et al. 2015), building upon empirical studies (Rubatto 2002; Hermann and Rubatto 2003; Whitehouse and Platt 2003; Hokada and Harley 2004; Kelly and Harley 2005; Buick et al. 2006). To address the challenge of handling extensive datasets generated by LA-ICP-MS, Taylor et al. (2017) developed an array plot for trace element partitioning data that can also show additional trends compared to the standard DREE plot (Fig. 3e). DREE (Mnz/Grt) is also a useful partitioning system (see Monazite section for more details). Trace element mapping of garnet via LA-ICP-MS is an area that is currently undergoing significant advances (e.g. Chew et al. 2017, 2021; Raimondo et al. 2017; Rubatto et al. 2020). M-HREE, Y and Cr zoning in garnet provides additional information about the growth history and mineral relationships not apparent in major element zonation (Raimondo et al. 2017).

Since the late 1980s, U–Pb and Sm–Nd geochronology (Mearns 1986; Mezger et al. 1989b) used multiple mineral phases to determine the age of whole-rock. The application of these methods led to the discovery of a favourable spread in Sm–Nd ratios between garnet and whole rock (Humphries and Cliff 1982). U–Pb garnet dating was found to be affected by U-rich inclusions such as zircon, monazite and apatite. Thus, for the following 30 years, garnet geochronology has been conducted using Sm–Nd and, subsequently, once instrumentation advanced sufficiently to measure low Hf contents in garnet, the Lu–Hf system was used for garnet dating (Duchêne et al. 1997). Inclusions such as monazite and apatite for the Sm–Nd system were treated with a HCl leaching step method that would remove significant monazite Nd components (Scherer et al. 2000), whereas Lu–Hf garnet geochronology is greatly affected by inclusions such as zircon, with samples containing significant contents of zircon being avoided for garnet whole-rock dating (Scherer et al. 2000). The following studies have compared the two isotopic systems for garnet dating via dissolution (Vervoort 2013), and they have shown that the Lu–Hf system (Tc c. 650–900°C) has slightly higher Tc than Sm–Nd (Tc c. 600–800°C) (Fig. 1; Scherer et al. 2000; Smit et al. 2013; Johnson et al. 2018). Traditionally, both Sm–Nd and Lu–Hf systems via isotope dilution–thermal ionization mass spectrometry (ID-TIMS) or MC-ICP-MS dissolution methods require the separation and purification of garnet prior to dissolution, column separation and measurement that is commonly associated with loss of the textural context of the analysed grains (cf. Pollington and Baxter 2010). Prolonged (e.g. Baxter and Scherer 2013; Bollen et al. 2022) or incremental garnet growth makes a single age from garnet separate, potentially geologically meaningless (e.g. Pollington and Baxter 2010). To overcome these issues, garnet has often been separated based on colour (e.g. Vance et al. 1998) or by using a microdrill to target garnet growth rings in order to determine the growth rate of garnet crystals (Fig. 3f; see Pollington and Baxter 2010, 2011; Dragovic et al. 2012, 2015; Schmidt et al. 2015; Tual et al. 2022). Results of such studies have showcased how effectively garnet can be used for geochronology, the caveat being that these methods are expensive, time consuming (c. 2 months) and the presence of inclusions must be considered and evaluated.

The recent introduction of ICP-MS/MS, which eliminates isotopic interferences for Lu–Hf measurements (i.e. 176Lu and 176Yb on 176Hf), coupled with an increase in machine sensitivity allowing to measure elements at very low concentrations (i.e. part per billion, ppb), has permitted significant advances in garnet geochronology. These analytical advances not only are at the foundation of the development of in situ Lu–Hf measurements in garnet (see Simpson et al. 2021; Brown et al. 2022; Ribeiro et al. 2022; Tamblyn et al. 2022; Simpson et al. 2023), but also resulted in revisiting of the U–Pb system in garnet with in situ analyses now possible using LA-ICP-MS or -SC (single collector) -MS (Seman et al. 2017; Gevedon et al. 2018; Millonig et al. 2020). Since garnet commonly has low U contents, the presence of U-rich inclusions remains an issue for in situ methods. Nonetheless, Uranium content in garnet can easily be tested via LA-ICP-MS spot analysis prior to U–Pb analysis and coupled with detailed textural imaging to identify potential U-rich inclusions. Grossular and andradite garnet tend to have enough U (>1 part per million, ppm), and many studies have used this method as a means of dating skarns related to mineralization (Seman et al. 2017; Gevedon et al. 2018; Wafforn et al. 2018; Duan et al. 2020). More recently, other works have also found almandine-pyrope-rich garnet with lower but enough U (>0.02 ppm) to produce reliable metamorphic ages (Millonig et al. 2020; Cerva-Alves et al. 2021; Schannor et al. 2021; O’Sullivan et al. 2023). Nonetheless, U-rich inclusions such as monazite, apatite and zircon shift data points on a Tera–Wasserburg (TW) diagram to the left, right or along the regression, reflecting older, younger or the same age as garnet, respectively (Fig. 3g; Millonig et al. 2020). Besides textural observations, to screen and filter out inclusion-dominated analyses it is crucial to monitor the masses on the ICP-MS such as Y and Zr for zircon, or light rare element (LREE) for monazite and allanite. The in situ U–Pb garnet dating method (Tc c. 650–900°C) has the advantage of minimal sample preparation, rapid analysis (2 hours) and the possibility of monitoring the signal. The Lu–Hf method is limited to garnet that contains sufficient Lu for measurement and relatively inclusion free, with zircon inclusions significantly affecting the age results (Simpson et al. 2021). This method has larger uncertainties than solution-based Lu–Hf analysis, but numerous advantages. Low Lu or high Hf (i.e. zircon) inclusions can be targeted to produce an isochron from a single garnet grain. The ablated spots (c. 40–120 µm) can target different chemical domains, including separate core and rim isochron ages (Fig. 3h–j; Tamblyn et al. 2022). The ability to produce an isochron from a single garnet grain means that this method can also be used to investigate detrital garnet (see Pereira et al. this volume, in press). An alternative to directly date garnet is targeting datable primary mineral inclusions such as monazite (e.g. Mahan et al. 2006; Cutts et al. 2010; Williams et al. 2017). However, possible preservation of detrital or older inclusions must be evaluated (Martin et al. 2007; Cutts et al. 2009, 2013; Peixoto et al. 2018).

Oxygen isotopes have been applied to whole-rock and mineral samples to trace the sources of metamorphic/metasomatic (e.g. Chamberlain and Conrad 1991; Crowe et al. 2001; Page et al. 2010; Martin et al. 2011, 2014; Vho et al. 2020) and magmatic/mantle fluids (e.g. King and Valley 2001; Lackey et al. 2006; Harris and Vogeli 2010). Initially, oxygen isotopes were applied to garnet to determine pressure estimates, based on the temperature dependent fractionation between co-existing mineral pairs (experimentally or empirically determined; Baxter et al. 2017 and references therein). This method utilizes slow diffusion of oxygen within garnet, thus peak temperature isotope fractionation is retained (Kohn and Valley 1998; Baxter et al. 2017). Improvements in analytical methods (analysis via SIMS) allowed for in situ analysis of oxygen isotopes in garnet to track the origin of metamorphic/metasomatic fluids (e.g. Raimondo et al. 2012, 2017; Russell et al. 2013). For example, garnet can preserve variation of fluid compositions as it grows, particularly in eclogites or skarns (D'Errico et al. 2012; Rubatto and Angiboust 2015). Oxygen in garnet is very similar to the bulk rock δ18O value, but deviations can occur following temperature variations affecting equilibration factors between minerals (i.e. garnet–quartz), changes in modal proportions and mineral assemblage during metamorphism as well as the presence of externally sourced fluids. Nonetheless, experimentally determined oxygen diffusion rates in garnet show that original isotope compositions can be retained despite partial re-equilibration of major elements (Scicchitano et al. 2021). Other stable isotopes such as Fe or Li may be applied in garnet (Bebout et al. 2014, 2022; An et al. 2017; Gerrits et al. 2019; Penniston-Dorland et al. 2020; Hoover et al. 2021, 2022), with limitations of these largely depending on analytical ability.

Geothermometry in garnet (Ferry and Spear 1978; Baxter et al. 2017) was initially applied via Fe–Mg exchange between garnet and biotite (Ferry and Spear 1978; Perchuk and Lavrent'Eva 1983; Ganguly and Saxena 1984) and geobarometry via the ‘GASP’ (garnet–aluminosilicate–silicate–plagioclase) end-member reaction, which is pressure sensitive due to the large molar volume difference between reactants and products (e.g. Ghent 1976; Newton and Haselton 1981; Holdaway 2001; Caddick and Thompson 2008). Later, large thermodynamic datasets (e.g. Berman 1988; Holland and Powell 1990, 1998, 2011) allowed for P–T pseudosection calculation, which was combined with the use of garnet end-member compositions to constrain the growth path or max P and T conditions (e.g. Spear et al. 1984; Vance and Mahar 1998). A caveat in the use of this method is the isolation of grown garnet (i.e. core domain) from the whole-rock composition used for P–T pseudosection calculation (e.g. Marmo et al. 2002; Evans 2004; Tinkham and Ghent 2005; Cutts et al. 2009, 2010). This approach is particularly useful where garnet growth domains are the result of different orogenic events (e.g. Cutts et al. 2014). Software for modelling evolving bulk compositions during garnet growth along a P–T path was developed to produce more reliable P–T estimates (Theria_G, Gaidies et al. 2008; GRTMod, Lanari et al. 2017). For P–T modelling, garnet compositions (i) may record prograde growth, which is typically defined by bell-shaped Mn enrichment in the garnet core (e.g. Hollister 1966; Atherton 1968; Cygan and Lasaga 1985; Schwandt et al. 1996; Dziggel et al. 2009; Lanari and Engi 2017; Spear 2017; Dempster et al. 2020), or (ii) alternatively, at high temperatures, garnet chemistry may be used to determine the max T conditions experienced by the sample where the prograde growth major element zonation is partially or completely re-equilibrated (Caddick et al. 2010). Also, fluid and/or melt inclusions hosted in peritectic garnet from migmatites have been used to evaluate the starting composition of the anatectic melt and fluid regime during anatexis (e.g. Cesare et al. 2009, 2011; Ferrero et al. 2012, 2021; Carvalho et al. 2019; Borghini et al. 2020). Inclusions such as zircon and quartz in garnet can also be used to produce pressure estimates using Raman barometry (Zhong et al. 2019; Spear and Wolfe 2020). After being trapped in a host mineral during part of the cooling and exhumation history of a rock, inclusions may develop residual pressures due to the differences in thermal expansivity and compressibility between the host and the inclusions (Gonzalez et al. 2019; Zhong et al. 2019, 2020). Within both metamorphic and magmatic studies, however, garnet-based thermobarometry should always be applied with a thorough petrological and microstructural relationships characterization and an appropriate estimate of equilibrium volume of the rock sample. Results that are inconsistent with the observed mineral assemblage should be crosschecked with other comparable tools and should be interpreted with care. Additionally, taking a step back and carefully reanalysing field and textural context of the investigated rocks, as well as reassessing the documentation of sample preparation, analytical procedure and data processing, may also reveal missed information.

The mineral zircon (ZrSiO4) has traditionally been the most robust accessory phase and widely used geochronological and geochemical tool in petrochronology to investigate crustal formation and evolution, and to untangle complex rock histories that most minerals fail to retain (e.g. Froude et al. 1983; Scherer et al. 2007; Guitreau et al. 2019). For instance, in the Archean Lewisian Gneiss Complex, NW Scotland, zircon is the only accessory mineral that has recorded the several Archean magmatic phases (cores) and the multiple ultra-high temperature (UHT) to HT/medium (M)T overprinting events (rims) (e.g. Whitehouse and Kemp 2010; Taylor et al. 2020; Fischer et al. 2021), whereas others such as monazite and titanite record only parts of the metamorphic events (e.g. Zhu and O'Nions 1999; Goodenough et al. 2013). Its resistance to erosive processes makes zircon a prime choice when studying detrital crystals to unravel past formation of continental crust in deeply eroded areas, or simply ancient terranes (e.g. Iizuka et al. 2005; Dhuime et al. 2012; Næraa et al. 2012; Nordsvan et al. 2018; Gardiner et al. 2019; Kirkland et al. 2021; Tedeschi et al. 2023; Pereira et al. this volume, in press). Over the years, many techniques have been improved to study zircon and retrieve various types of information from it (Table 1).

Table 1.

Synoptic table of most used tools for zircon investigations

Tool*InterestExample of reference
MorphologyNature of parental magma (alkaline, peraluminous or calc-alkaline)Pupin (1980); Vavra (1993) 
CL and BSE imagesPristineness of crystal lattice and crystallization conditions (igneous, metamorphic or weathered)Vavra (1990); Hanchar and Miller (1993); Corfu et al. (2003); Guitreau et al. (2018) 
Inclusions (mineral and melt)Nature of parental magma and/or fluids (for primary inclusions either as melt or mineral)Thomas et al. (2003); Hopkins et al. (2008); Bell et al. (2015) 
EBSDOrientation and deformation of crystalsTolometti et al. (2022); Cox et al. (2022) 
Trace element concentrationsNature of parental magma, magmatic or metamorphic origin and pristinenessGrimes et al. (2007, 2015); Laurent et al. (2021); Guitreau et al. (2022) 
(Ti)Crystallization temperatureWatson et al. (2006); Fu et al. (2008) 
(K) and (Ca)Primitive v. altered zirconBouvier et al. (2012); McCubbin et al. (2016) 
Th/UZircon crystallization conditions (e.g. magmatic v. metamorphic origin) and weatheringVavra et al. (1999); Kirkland et al. (2015); Yakymchuk et al. (2018); Guitreau and Flahaut (2019) 
Ce anomalyRedox state of parental magmaSmythe and Brenan (2016); Trail et al. (2012) 
U–PbDating (absolute age determination)Wetherill (1956, 1963); Davis et al. (2003 and reference therein); Schoene (2014) 
Lu–HfSource tracing (e.g. mantle or crust)Patchett (1983); Amelin et al. (2000); Gerdes and Zeh (2009); Guitreau et al. (2012) 
OSource tracing (e.g. involvement of clay-rich sediments)Valley (2003); Iizuka et al. (2013) 
Si isotopesSource tracing, parental magma SiO2 content and crystallization historyTrail et al. (2018); Guitreau et al. (2020, 2022)
Zr isotopesMagma crystallization historyIbañez-Mejia and Tissot (2019); Guo et al. (2020); Tompkins et al. (2020) 
Tool*InterestExample of reference
MorphologyNature of parental magma (alkaline, peraluminous or calc-alkaline)Pupin (1980); Vavra (1993) 
CL and BSE imagesPristineness of crystal lattice and crystallization conditions (igneous, metamorphic or weathered)Vavra (1990); Hanchar and Miller (1993); Corfu et al. (2003); Guitreau et al. (2018) 
Inclusions (mineral and melt)Nature of parental magma and/or fluids (for primary inclusions either as melt or mineral)Thomas et al. (2003); Hopkins et al. (2008); Bell et al. (2015) 
EBSDOrientation and deformation of crystalsTolometti et al. (2022); Cox et al. (2022) 
Trace element concentrationsNature of parental magma, magmatic or metamorphic origin and pristinenessGrimes et al. (2007, 2015); Laurent et al. (2021); Guitreau et al. (2022) 
(Ti)Crystallization temperatureWatson et al. (2006); Fu et al. (2008) 
(K) and (Ca)Primitive v. altered zirconBouvier et al. (2012); McCubbin et al. (2016) 
Th/UZircon crystallization conditions (e.g. magmatic v. metamorphic origin) and weatheringVavra et al. (1999); Kirkland et al. (2015); Yakymchuk et al. (2018); Guitreau and Flahaut (2019) 
Ce anomalyRedox state of parental magmaSmythe and Brenan (2016); Trail et al. (2012) 
U–PbDating (absolute age determination)Wetherill (1956, 1963); Davis et al. (2003 and reference therein); Schoene (2014) 
Lu–HfSource tracing (e.g. mantle or crust)Patchett (1983); Amelin et al. (2000); Gerdes and Zeh (2009); Guitreau et al. (2012) 
OSource tracing (e.g. involvement of clay-rich sediments)Valley (2003); Iizuka et al. (2013) 
Si isotopesSource tracing, parental magma SiO2 content and crystallization historyTrail et al. (2018); Guitreau et al. (2020, 2022)
Zr isotopesMagma crystallization historyIbañez-Mejia and Tissot (2019); Guo et al. (2020); Tompkins et al. (2020) 

*BSE, Backscattered electron; CL, cathodoluminescence; EBSD, electron backscatter diffraction.

When zircon crystallizes, it can develop various textures depending on the conditions at which it forms, such as sharp concentric growth zoning and/or marginal resorption if zircon grew in a magmatic environment or more irregular domains of homogeneous zoning cutting discordantly the growth domains in the post-magmatic environment (e.g. Corfu et al. 2003 and references therein). These textures are commonly revealed by routine imaging of cathodoluminescence (CL) or backscattered electrons (BSE) using a scanning electron microscope (SEM). Cathodoluminescence response is mostly due to interaction between the primary electron-beam and zircon lattice through a complex energy transfer reaction that ends with the emission of photons by tetravalent REE, especially Dy (e.g. Nasdala et al. 2003). Visible textures range from broad to fine oscillatory zoning in igneous zircon (Fig. 2d–f; Fig. 4a, b) to large homogeneous to patchy domains in metamorphic conditions (e.g. Corfu et al. 2003). Raman spectroscopy can be used to quantitatively map radiation damage zoning in single zircon crystals (i.e. metamictization; Balan et al. 2001) and, hence, evaluate the effects of radiation damage on the final U/Pb ratio (Anderson et al. 2020) or on any elemental and/or isotopic signature. Igneous zircon most commonly form when the activity of ZrO2 and SiO2 is optimal. This is the case in most magmas with SiO2 concentrations over 57 wt%, although rare crystals in magmas with lower SiO2 concentrations do exist (e.g. Aranovich et al. 2017; Fischer et al. 2021; Bea et al. 2022). Zircon precipitation from melts was investigated early on by Watson (1979) and Watson and Harrison (1984), and revisited by Boehnke et al. (2013).

Fig. 4.

(a) Cathodoluminescence image of a complexly zoned zircon crystal from the Napier Complex (Antarctica) showing a (recrystallized) bright metamorphic core surrounded by a fine-oscillatory zoned magmatic domain, itself resorbed and overgrown by three distinct metamorphic overgrowths. (b) Igneous zircon crystals exhibiting fine-oscillatory zoning from core to rim except for a thin metamorphic outer overgrowth. (c) Representative chondrite normalized REE patterns for different types of zircon. (1) Rubatto and Hermann (2007); (2) Hoskin and Schaltegger (2003); (3) Rubatto et al. (2013); (4) Taylor et al. (2014); (5) Li et al. (2018). (d) Modelled diffusion modified M-HREE compositions expected from c. 20 µm diameter SHRIMP spot analysis on the edge of a modelled theoretical zircon. C0, original unmodified composition; CEq, equilibrated composition. (e) M-HREE spread generated by incomplete modification of a modelled zircon after different T–t conditions. (f) Concordia diagram illustrating the interpretation of zircon UPb data in terms of ages depending on whether the UPb system evolved in a closed or open system, and when disturbance occurred. (g) Graphical representation of Lu–Hf isotope evolution of distinct geological reservoirs. This graph provides a theoretical framework for the interpretation of zircon Hf isotope signatures. Source: (d) and (e) after Blereau et al. (2022).

Fig. 4.

(a) Cathodoluminescence image of a complexly zoned zircon crystal from the Napier Complex (Antarctica) showing a (recrystallized) bright metamorphic core surrounded by a fine-oscillatory zoned magmatic domain, itself resorbed and overgrown by three distinct metamorphic overgrowths. (b) Igneous zircon crystals exhibiting fine-oscillatory zoning from core to rim except for a thin metamorphic outer overgrowth. (c) Representative chondrite normalized REE patterns for different types of zircon. (1) Rubatto and Hermann (2007); (2) Hoskin and Schaltegger (2003); (3) Rubatto et al. (2013); (4) Taylor et al. (2014); (5) Li et al. (2018). (d) Modelled diffusion modified M-HREE compositions expected from c. 20 µm diameter SHRIMP spot analysis on the edge of a modelled theoretical zircon. C0, original unmodified composition; CEq, equilibrated composition. (e) M-HREE spread generated by incomplete modification of a modelled zircon after different T–t conditions. (f) Concordia diagram illustrating the interpretation of zircon UPb data in terms of ages depending on whether the UPb system evolved in a closed or open system, and when disturbance occurred. (g) Graphical representation of Lu–Hf isotope evolution of distinct geological reservoirs. This graph provides a theoretical framework for the interpretation of zircon Hf isotope signatures. Source: (d) and (e) after Blereau et al. (2022).

The nature of mineral inclusions can provide a first order evidence of phases that co-precipitated with zircon and, thus, give information about the nature/composition of the melt in which it crystallized and the pressure and temperature conditions at which it grew (Delavault et al. 2016; Emo et al. 2018; Antoine et al. 2020). The same can be done with melt inclusions, but these can also be used to estimate partition coefficients for trace elements if they represent bulk-melt composition and their chemistry remained unmodified since entrapment (Thomas et al. 2003; Gudelius et al. 2020).

Zircon incorporates large quantities of trace elements from ppm to thousands of ppm level (e.g. Belousova et al. 2001; Hoskin and Schaltegger 2003), making it an easy target for analysis using electron microprobe (EPMA), ICP-MS, LA-(MC-)ICP-MS (Košler et al. 2005; Chew et al. 2017; Petrus et al. 2017) and secondary ionization mass spectrometers or sensitive high-resolution ion microprobe (SIMS/SHRIMP). The rapid technological improvements using LA-(MC-)ICP-MS allow now the acquisition and extraction of quantitative data from a mapped area, including chemical and isotopic information of accessory phases such as zircon (Petrus et al. 2017; Chew et al. 2021). Experiments and natural observations indicate very slow diffusion for most elements within the zircon lattice (see Lee et al. 1997; Cherniak and Watson 2003 for values). Pristine igneous zircon elemental variability is usually relatively small, making this mineral a poorly sensitive tracer of melt composition, except for sharp variations (i.e. strongly alkaline, felsic and mafic; e.g. Belousova et al. 2002; Grimes et al. 2007; Guitreau et al. 2022). However, when zircon undergoes significant radiation damage, it can be sensitive to thermal events that result in incorporation of measurable quantities of light REEs (e.g. Hoskin and Schaltegger 2003; Bouvier et al. 2012; Pidgeon et al. 2017). This modification of zircon composition is a good proxy for secondary alteration and/or weathering. Along the same line, non-formula elements, such as Ca, K and Al, can enter zircon easily once it becomes porous due to significant radiation damage accumulation (e.g. Holland and Gottfried 1955; Ewing et al. 2003; McCubbin et al. 2016; Pidgeon et al. 2019). Monitoring LREEs and non-formula elements allows for filtering of pristine zircon crystals or domains from altered and weathered ones. In contrast, rare earth element patterns in metamorphic zircon are more useful because equilibrium with other phases has a forcing effect on partition coefficients such that zircon cannot incorporate as much HREE as it would normally do (Fig. 4c). This is well illustrated when zircon forms in equilibrium with garnet (e.g. Kelly and Harley 2005; Blereau 2017), resulting in flattened to steepened heavy REE patterns (Fig. 4c). Whilst extremely sluggish under most conditions of metamorphism (Cherniak and Watson 2003), under extreme metamorphic conditions (>1000°C), the REE content in zircon may also be modified as a function of temperature and time (Watson 1996; Blereau et al. 2022), although others have suggested decoupling of REE from UPb above 850°C (Kunz et al. 2018; Jiao et al. 2020b; Durgalakshmi et al. 2021). REEs within zircon are impurities that are expelled at high-T resulting in the removal of internal zoning (Hoskin and Black 2000), with garnet acting as a sink for the released REEs (see Discussion). Diffusional modelling of REE-in-zircon shows that an initially igneous zircon (i.e. an inherited grain with steep M-HREE patterns) within a garnet-bearing metapelite is in disequilibrium with the garnet, and when exposed to metamorphic temperatures, the zircon attempts to reach equilibrium with garnet (i.e. a c. 1:1 DREE pattern) (Blereau et al. 2022; Fig. 4d). Zircon with a 50 µm radius can be re-equilibrated with the host metamorphic assemblage during both short (1100°C for 1–5 Ma) and extended periods (1050°C for 10–30 Ma or 1000°C for 200 Ma) of UHT metamorphism (Blereau et al. 2022; Fig. 4e). Conversely, unless diffusion is enhanced by fluids or other processes, below 900°C, zircon will largely preserve its pre-metamorphic REE signature, even when metamorphism is prolonged (>100 Ma) (Blereau et al. 2022; Fig. 4e). The change of valence of Ce and Eu to 4+ and 2+, respectively, in igneous zircon can be used to estimate redox conditions through Ce and Eu anomalies (e.g. Trail et al. 2012; Smythe and Brenan 2016; W.T. Li et al. 2021). However, some authors suggest using these proxies with caution because of melt cooling and chemical evolution (e.g. Loader et al. 2022). Europium anomalies can also be used to track feldspar fractionation and to estimate whether zircon crystallized in equilibrium with feldspar in igneous and metamorphic zircon, respectively (Rubatto 2002).

The ratio of Th over U is often used in zircon because it provides information about co-precipitating phases (e.g. Kunz et al. 2018). It can help discriminate igneous (Th/U c. 0.2–0.8) from metamorphic (Th/U < 0.1 or >1) zircon domains/crystals, distinguish between zircon formed under amphibolite-facies and granulite-facies and can trace low-temperature weathering of radiation-damaged zircon (e.g. Vavra et al. 1999; Hoskin and Black 2000; Kirkland et al. 2015; Yakymchuk et al. 2018; Guitreau et al. 2019; Guitreau and Flahaut 2019; Barrote et al. 2020). However, Th/U in zircon is not always a faithful recorder of metamorphic processes (Möller et al. 2003; Harley and Kelly 2007), and it should be used with caution on a case-by-case basis. Typically, U–Pb LA-ICP-MS geochronology routines do not include appropriate internal standard elements, therefore, the U, Th and Pb concentrations are semi-quantitative only.

Zircon is most commonly dated using the U–Pb isotope system (e.g. Schoene 2014) as, for this systematics, it has several advantages compared to other minerals. Zircon U–Pb isotope measurements are done using a great variety of techniques from solution-based such as TIMS for the most precise ages and MC-ICP-MS/ICP-MS to in situ microbeam techniques (i.e. SIMS, SHRIMP, LA-ICP-MS, LA-MC-ICP-MS). One of the main advantages is that Pb is essentially excluded from the zircon lattice because of its size and valence, resulting in virtually all measurable Pb in zircon being radiogenic – produced by the radioactive decay of U and Th isotopes. This means that the parent/daughter ratio measured in pristine zircon combined with U decay constants can be directly converted into an age (Schoene 2014). However, zircon lattice can become damaged by radioactive decay over time, which results in metamictization. This process makes zircon porous to external agents that may incorporate Pb, and other non-formula elements, into the crystal lattice, thus compromising the determined age. Another major benefit of the U–Pb isotope system compared to others is the fact that it contains two isotope systems (i.e. 238U–206Pb and 235U–207Pb) with distinct decay constants (e.g. Le Roux and Glendenin 1963; Jaffey et al. 1971), allowing open- v. closed-system evolution to be assessed. For easy visualization, U–Pb data are commonly plotted in 238U–206Pb and 235U–207Pb Wetherill concordia diagram (Fig. 4f; Wetherill 1956). When both dates lie on the concordia curve, the date is called concordant and is geologically meaningful since it likely reflects a closed U–Pb system evolution. By contrast, when both dates are different, the measured date is discordant and evaluation on the geological meaning will vary case by case. A third date can also be derived directly from the 207Pb/206Pb ratio that inherently assumes closed-system evolution. This date is the oldest of the three in the case of ‘normally’ discordant data (i.e. datapoint is located below the concordia curve). Unless information regarding the validity of intercepts in concordia diagram is available (e.g. multiple analyses from the same crystal or growth zones), 207Pb/206Pb ages are commonly more precise for zircons >1.5 Ga, whereas 238U–206Pb ages are used for crystals <1.5 Ga (Spencer et al. 2016).

Depending on when the U–Pb isotope system disturbance occurred, the memory of primary crystallization may or may not be preserved. If U–Pb disturbance is recent, U–Pb data are discordant and distribute along a discordia line that passes through zero (blue points). Since Pb isotopes are not fractionated in detectable proportions in such disturbance, 207Pb/206Pb ratios still provide the primary crystallization age, which corresponds to the discordia upper-intercept (Fig. 4f) and is identical to the purple-filled ellipse which represents a concordant analysis from either the same zircon or the same population that remained unmodified. In contrast, for an old U–Pb disturbance, the datapoints would align along a discordia line that connects a lower-intercept corresponding to the age of U–Pb disturbance and an upper-intercept representing the actual crystallization age of zircon (red-filled ellipses in Fig. 4f). Most cases of discordance correspond to Pb-loss, which graphically corresponds to a migration of datapoints towards the lower end of a discordia line, and less common cases are associated with reverse discordance which reflects Pb accumulation (e.g. Williams et al. 1984; Kusiak et al. 2013). The case in which zircon contains measurable amounts of common-Pb (Pbc) has not been presented here because, in most cases, it is a sign of zircon post-crystallization modification and/or advanced alteration. Consequently, data are generally discarded when 204Pb is detected, since the Pbc correction relies on knowledge of the isotopic composition of Pb when it entered the crystal, which is difficult to know in zircon.

The source of zircon parental magma can be assessed using multiple isotope systems, with the most common being Lu–Hf and O isotopes (e.g. Patchett 1983; Valley 2003). Lu–Hf isotopes in zircon are normally measured using MC-ICP-MS, either in solution or laser-ablation mode (e.g. Fisher et al. 2014a, b). The Lu–Hf isotope system is radiogenic and based on the decay of radioactive 176Lu into 176Hf with a half-life of c. 36 Ga (Scherer et al. 2001; Söderlund et al. 2004). The measured 176Hf/177Hf tracks the time-integrated fractionation of Lu from Hf in a reservoir (Fig. 4g). The principle of this technique is that during partial melting of most mantle and crustal lithologies, Hf and Lu are fractionated from each other, which results in magmas having Lu/Hf lower than that of the melting residue. Over time, melting residues (refractory mantle) develop very radiogenic (elevated) 176Hf/177Hf ratios (positive εHf), and crustal lithologies (former magmas) comparatively low radiogenic ratios (negative εHf) (Fig. 4g). For global interpretation, Hf isotope compositions are normalized to the Chondritic Uniform Reservoir (CHUR; Blichert-Toft and Albarède 1997; Bouvier et al. 2008; Iizuka et al. 2015) and transformed into epsilon Hf notation (εHf), with CHUR approximating the bulk silicate Earth composition. For instance, Figure 4g illustrates a magma source produced by partial melting of the mantle at 2600 Ma (green spot in Fig. 4g) and that differentiates into mafic and felsic lithologies, which are both characterized by specific 176Lu/177Hf content. These lithologies evolve after crystallization until 1500 Ma, when they are reworked by partial melting forming different granites (blue and red spots in Fig. 4g). Some of the generated melts mix with each other forming hybrid granites (grey spot in Fig. 4g). At 600 Ma, magmas extracted from the mantle (green spot in Fig. 4g) mix with crustal melts forming new hybrid granites (orange spot in Fig. 4g). Note that all crustal lithologies contain zircon, which allows the evolution of these reservoirs to be followed through time. Newly formed zircon domains (grains or overgrowths) within a single rock sample mostly have higher initial 176Hf/177Hf than older domains due to incomplete dissolution of detrital or magmatic zircon grains (hosting most of the non-radiogenic Hf). Also, metamorphic zircon incorporates additional radiogenic 176Hf formed by 176Lu decay in the rock's matrix between successive zircon growth events (Gerdes and Zeh 2009).

Oxygen isotopes are stable isotopes and the measured 18O/16O, and possibly 17O/16O when the triple-isotope system is used, tracks isotope fractionations due to magmatic processes and/or sources. These light isotopes do not fractionate much during magmatic processes (i.e. partial melting and fractional crystallization), but differences can be measured (e.g. Valley 2003; Bindeman 2008). Most O isotope studies in zircon use the source-tracing potential of O isotopes, taking advantage of large O isotope fractionation caused by low-temperature alteration of crustal lithologies inducing clay formation and resulting in enrichment of 18O relative to 16O. Much like the Lu–Hf isotope system, O isotope ratios are difficult to interpret as numerical values. Therefore, O isotope ratios are normalized relative to the international Vienna standard mean ocean water (VSMOW) and reported in the standard delta (δ18O) notation. Zircon formed from a reworked clay-rich crustal lithology may have δ18O up to +12‰ (Valley et al. 2005; Kemp et al. 2007), and mantle zircons have consistent δ18O of +5.3 ± 0.6‰ (Valley 2003). Zircon δ18O values below that of the mantle indicate that zircon parental melt interacted with meteoric fluids and/or was altered at high temperatures (Valley 2003 and references therein). Most studies using O isotopes in zircon interpret O isotope variations as evidence for mixtures between mantle- and crustal-derived melts and/or fluids or crustal contaminants (e.g. Kemp et al. 2007; Smithies et al. 2021). Recent studies have demonstrated that the uptake of water into the zircon crystal lattice can significantly modify its oxygen isotopic composition (Pidgeon et al. 2017; Liebmann et al. 2021). Therefore, monitoring of the 16O1H/16O ratio during SIMS oxygen isotope measurements to assess secondary modification of O isotope composition by water addition is recommended (Liebmann et al. 2021).

Recently, new stable isotope systems such as Si and Zr have been applied to zircon. Different techniques such as SIMS, MC-ICP-MS and LA-MC-ICP-MS have been employed to measure Si isotopes in zircon, which have resulted in good precision and closely mimic the natural variability of high-temperature processes (Trail et al. 2018, 2019; Guitreau et al. 2020, 2022). Si isotopes can be used to trace the origin of zircon parental magma due to the various silicon isotope signatures found in different types of igneous rocks (e.g. I, A, S, tonalite–trondhjemite–granodiorite (TTG); Savage et al. 2014; Deng et al. 2019). Si isotope compositions, expressed as permil deviations from a quartz standard (δ29Si and δ30Si), are sensitive to the SiO2 content, which reflects the degree of polymerization and crystallization temperature (Qin et al. 2016; Trail et al. 2019; Guitreau et al. 2022). This allows for reconstructions of the magma evolution. Changes in Si isotope compositions can occur due to metamorphic alteration and/or zircon recrystallization, which are dependent on the metamorphic grade (Guitreau et al. 2022).

Zr isotopes can be measured using conventional MC-ICP-MS instruments in solution or laser mode and applied to track fractional crystallization processes (Ibañez-Mejia and Tissot 2019; Zhang et al. 2019; Guo et al. 2020; Tian et al. 2020). Recent studies attributed measurable Zr isotope variations to kinetic fractionation in response to chemical gradients rather than equilibrium processes (Chen et al. 2020; Méheut et al. 2021). Zirconium isotope compositions of metamorphic zircon compared to igneous zircon may also result from chemical gradient effects (Zhang et al. 2019). This technique can hence provide insights into magmatic crystallization dynamics.

Zircon is used as a mineral pair thermometer based on incorporation of Ti into zircon at HT coupled with Zr substitution in rutile (Zack et al. 2004a; Watson and Harrison 2005; Watson et al. 2006; Ferry and Watson 2007). This technique requires the presence of both zircon and rutile within the mineral paragenesis, otherwise all temperatures are minimum estimates (Watson and Harrison 2005). Magmatic and metamorphic temperatures can also be overestimated at low pressures (<5 kbar) (Rubatto 2017). Temperature variations within an investigated sample have been interpreted to reflect waves of magmatic pulses within a magma chamber generating thermal and compositional heterogeneities (e.g. Collins et al. 2016; Volante et al. 2020a). Low diffusivity of Ti within zircon was previously measured perpendicular to the crystallographic c-axis (Cherniak and Watson 2007), making this thermometry a widely used tool. In contrast, recent experiments conducted parallel to the c-axis found significant anisotropy in the diffusivity of Ti (Bloch et al. 2022). When extrapolated, the resulting diffusivities are c. 7.5–11 times faster at 950–650°C than the original experiments, indicating that this thermometer can be modified under elevated crustal temperatures, slow cooling and/or small grain sizes (Bloch et al. 2022). Moreover, since this thermometer is dependent upon alpha-quartz and t-TiO2 activities, software such as Rhyolite-MELTS Gualda et al. (2012) have been proposed to improve temperature estimates’ precision (e.g. Schiller and Finger 2019). The latter tool is useful and intuitive to use, however its caveats (Volante et al. 2020a) and its relatively low precision of about 50°C (Guitreau et al. (2022)) should be considered.

Titanite (Ca[Ti,Al,Fe3+]SiO4[O,OH,F,Cl]) (Fig. 2g-i; Fig. 5) can occur as minor or accessory mineral and it is a particularly efficient chemical reactant with other major mineral phases that contain Ca and Ti, commonly leading to re-crystallization (Frost et al. 2001). Titanite commonly grows during magmatic (e.g. Ca-rich granitic rocks), metamorphic (e.g. calc-silicate, amphibolite) and hydrothermal events (Fig. 5a-c), with each different setting resulting in different titanite REE compositions that can be tracked by trace element fingerprinting (see reviews by Frost et al. 2001; Kohn 2017). Titanite commonly forms in mafic and calc-silicate rocks, and its growth can provide valuable time constraints in cases where other accessory minerals do not develop. Titanite crystallizes over a broad P–T range, including growth after breakdown of rutile, which is commonly stable at higher pressures (>1.4 GPa), ilmenite, Fe–Ti oxide stable at higher temperatures (>650/700°C), magnetite and/or clinozoisite (Frost et al. 2001; Kohn 2017). The combined acquisition of geochronological and geochemical data in texturally controlled titanite has robustly discriminated between distinct metamorphic, hydrothermal and magmatic events, and provided the opportunity to evaluate physical and chemical processes at the micro-scale in titanite-bearing rocks (e.g. Stearns et al. 2015; Garber et al. 2017; Olierook et al. 2019; Cavosie et al. 2022; Walters et al. 2022). However, petrological and chemical complexities can challenge U–Pb and P–T data interpretation. For instance, overgrowths of multiple metamorphic and hydrothermal titanite generations can significantly affect and modify the U–Pb system (e.g. Storey et al. 2006; Marsh and Smye 2017; Kirkland et al. 2018) and trace element compositions (e.g. Gordon et al. 2021), inducing decoupling of titanite U–Pb dates and trace element compositions (e.g. Romer and Rötzler 2001, 2011; Castelli and Rubatto 2002; Bonamici and Blum 2020; Walters et al. 2022).

Fig. 5.

Titanite. (a) titanite in orthogneisses. (b) Oscillatory zoning in igneous titanite. (c) Patchy zoning in metamorphic titanite. (d) REE patterns in magmatic and metamorphic titanite. (e) Slopes of REE patterns are affected by other mineral phases growing in equilibrium with (i) igneous Precambrian and (ii) metamorphic Caledonian titanite grains. (f) Al v. Fe. (g) Th/U v. Zr-in-titanite. (h) Th/Pb v. Zr-in-titanite. (i) Al/(Al + Fe) v. total REE discrimination diagrams to distinguish between igneous (in red), recrystallized (in purple) and metamorphic (in green) titanite. (j) Schematic Tera–Wasserburg concordia diagram for interpreting U–Pb titanite data. Source: (a), (c), (d) and (e) modified after Garber et al. (2017); (b), (i) and (j) modified after Olierook et al. (2019).

Fig. 5.

Titanite. (a) titanite in orthogneisses. (b) Oscillatory zoning in igneous titanite. (c) Patchy zoning in metamorphic titanite. (d) REE patterns in magmatic and metamorphic titanite. (e) Slopes of REE patterns are affected by other mineral phases growing in equilibrium with (i) igneous Precambrian and (ii) metamorphic Caledonian titanite grains. (f) Al v. Fe. (g) Th/U v. Zr-in-titanite. (h) Th/Pb v. Zr-in-titanite. (i) Al/(Al + Fe) v. total REE discrimination diagrams to distinguish between igneous (in red), recrystallized (in purple) and metamorphic (in green) titanite. (j) Schematic Tera–Wasserburg concordia diagram for interpreting U–Pb titanite data. Source: (a), (c), (d) and (e) modified after Garber et al. (2017); (b), (i) and (j) modified after Olierook et al. (2019).

Texturally, both magmatic and metamorphic (or hydrothermally altered) titanite can exhibit complex compositional zoning (Fig. 5a–c) and associated trace element (TE) patterns related to fluid–rock interaction (e.g. Smith et al. 2009; Garber et al. 2017; Olierook et al. 2019; Walter et al. 2021; Walters et al. 2022). Titanite preferentially incorporates minor and TE including U, high field-strength elements (HFSEs) such as Zr and REEs (e.g. Tiepolo et al. 2002; Lucassen et al. 2012), compared to other accessories (e.g. zircon, monazite). Trace element in titanite can provide information about crystallization pressure and temperature estimates (Hayden et al. 2008), oxygen fugacity (e.g. King et al. 2013; Cao et al. 2015) and fractionation processes (e.g. Piccoli et al. 2000; John et al. 2011). Geochemical and experimental studies on titanite suggest that it preferentially incorporates the medium rare earth element (MREE) in its crystal structure when in equilibrium with melt (e.g. Tiepolo et al. 2002; Prowatke and Klemme 2005; Olin and Wolff 2012), whereas recent petrochronological works show that melt-present and melt-absent metamorphic titanite have different REE patterns strongly depending on the presence and/or abundance of other phases in equilibrium with them (e.g. Garber et al. 2017; Walter et al. 2021; Walters et al. 2022). For instance, different TE uptake in titanite allowed Garber et al. (2017) to discriminate between (i) LREE-enriched (Precambrian) igneous cores, (ii) LREE-depleted and HREE-rich (Caledonian) recrystallized metamorphic rims likely associated with fluid or melt and (iii) (Caledonian) neocrystallized metamorphic titanite reflecting a negative or more positive REE slope based on whether titanite grew in equilibrium with hornblende and garnet or with allanite, apatite, plagioclase and biotite, respectively, during prograde and/or retrograde amphibolite-facies metamorphism (Fig. 5d, e; see also Cioffi et al. 2019; Walter et al. 2021).

Recent studies have highlighted the potential of TE in titanite as petrogenetic discriminator/indicator in magmatic, metamorphic and detrital studies (e.g. Ma et al. 2019; Olierook et al. 2019; Scibiorski and Cawood 2022). For example, Olierook et al. (2019) show that Fe concentrations and Th/U and Th/Pb ratios are systematically higher in magmatic than metamorphic titanite, reflecting useful petrogenetic discriminators when plotted against Al and Zr-in-titanite temperature (T°C), respectively (Fig. 5f–h). Also, normalized LREE/MREE, LREE/HREE and Eu anomalies reflect negative and/or positive REE slope correlations, which are useful to complement the characterization of titanite origin (Olierook et al. 2019). TE-based discrimination can be corroborated with detailed textural investigation of titanite grains, which are commonly characterized by oscillatory or sector zoning when magmatic in origin, with patchy and/or homogeneous zoning in metamorphic grains (e.g. Smith et al. 2009; Garber et al. 2017; Walters and Kohn 2017; Cioffi et al. 2019; Holder and Hacker 2019). Scibiorski and Cawood (2022) show that different titanite host-rock lithologies are reflected in the variation of TE chemistry, with low Zr/Y and high Fe in titanite from felsic host-rocks. This complements Al/Fe v. ΣLREE and U v. Th contents (ppm) used to discriminate magmatic v. metamorphic titanite (Olierook et al. 2019; Scibiorski and Cawood 2022; Fig. 5i) and between metamorphic, recrystallized and inherited igneous titanite (fig. 2 in Scibiorski and Cawood 2022), respectively. Also, systematic integration of titanite microstructural investigations by electron backscatter diffraction (EBSD) with in situ U–Pb petrochronology was proved to be a powerful tool not only to untangle and reconstruct complex deformation histories within crustal-scale high-strain zones but also to constrain deformation mechanism associated with shock and thermal metamorphism (Papapavlou et al. (2017); McGregor et al. 2021).

Titanite is a powerful U–Pb petrochronometer recording different primary and secondary geological processes (Spandler et al. 2016; Ma et al. 2019; Fisher et al. 2020; Barla 2021). However, titanite and other accessories including rutile, apatite and allanite tend to incorporate significant Pbc (e.g. Kirkland et al. 2017, 2018; Bonamici and Blum 2020), which is reflected in discordant U–Pb ratios (e.g. Marsh and Smye 2017), making U–Pb dates interpretation of these accessory phases challenging (e.g. Olierook et al. 2019; Walters et al. 2022). Figure 5j (modified after Olierook et al. 2019) represents a schematic TW diagram (Tera and Wasserburg 1972) illustrating how titanite U–Pb data (and the other phases that commonly have Pbc), including Pbc (see Storey et al. 2006; Kirkland et al. 2017), crystallization and recrystallization mixing lines can be interpreted (e.g. Spencer et al. 2013; Bonamici et al. 2015; Garber et al. 2017; Papapavlou et al. 2017; Holder and Hacker 2019; Mottram et al. 2019; Timms et al. 2020; Gordon et al. 2021). As a result of Pbc incorporation in the crystals, geochronological data typically form an intercept line (red line in Fig. 5j, ‘igneous titanite ages’) characterized by a 207Pb/206Pb upper intercept and lower concordia intercept (e.g. Spencer et al. 2013; Chew et al. 2014; Garber et al. 2017; Kirkland et al. 2017, 2018, 2020), with the igneous age being calculated using the lower one. However, when such intercept line is not statistically robust, then (i) a weighted mean of uncorrected dates can be calculated, if they are within 2SD uncertainty (e.g. Spencer et al. 2016; Olierook et al. 2019; Barrote et al. 2022b), or (ii) the upper intercept can be used to calculate concordant analyses with negligible Pbc (Olierook et al. 2019). U–Pb dates of samples affected by overprinting events would fall in a space in the TW diagram defined by a 207Pb/206Pb upper intercept, a first lower concordia intercept recording the first magmatic/metamorphic event and a second lower concordia intercept recording the subsequent event (green triangular shape in Fig. 5j). Titanite is also found to yield non-typical Pbc values, possibly due to inheritance of radiogenic Pb into the crystal structure (Kirkland et al. 2018; Mottram et al. 2019; Walters et al. 2022).

Additionally, an important factor to consider when dealing with titanite geochronology is Tc (Fig. 1; Dodson 1973). In the past two decades, it has been shown that titanite is much more retentive than previously envisaged (with Tc of c. 600–650°C, Mezger et al. 1991; 600–650°C, Scott and St-Onge 1995, Frost et al. 2001; 650 and 750°C, Cherniak 1993), increasing the temperature threshold for Pb and Zr volume diffusion in titanite as high as c. 750–840°C (c. 740°C, e.g. Schärer et al. 1994, Zhang and Schärer 1996; >750°C, e.g. Kylander-Clark et al. 2008, Kohn and Corrie 2011, Gao et al. 2012, Spencer et al. 2013, Stearns et al. 2015, Kohn 2017; >830°C, e.g. Hartnady et al. 2019, Holder et al. 2019, Kirkland et al. 2020) and challenging the assumption that titanite U–Pb dates commonly reflect cooling ages (Hartnady et al. 2019; Kirkland et al. 2020).

While less explored compared to other accessory minerals such as monazite, Sm–Nd isotope systematics in titanite (e.g. Yang et al. 2008; Amelin 2009; Fisher et al. 2011, 2020; Hammerli et al. 2014; Spandler et al. 2016; Ma et al. 2019; Zhang et al. 2021) has potential as a source tracer for understanding the formation and evolution of the crust (Amelin 2009; Fisher et al. 2020; Zhang et al. 2021). This isotopic system has been demonstrated to have high Tc (850–950°C; Cherniak 1995), surviving HT magmatic, metamorphic and hydrothermal conditions. In situ Sm–Nd isotopic systematics of titanite can be used to trace fluid, melt or crustal/juvenile rock sources (e.g. Lucassen et al. 2011; Hammerli et al. 2014; Spandler et al. 2016; Zhang et al. 2021).

Experimental studies have demonstrated that oxygen diffusion in accessory minerals is slower than in rock-forming minerals (e.g. Fortier and Giletti 1989), but oxygen diffusion in apatite is faster than in titanite, which is in turn faster than in zircon (Bruand et al. 2019). Oxygen isotopes in magmatic and metamorphic titanite have been used as a geochemical indicator to investigate magma petrogenesis as well as metamorphic and hydrothermal overprinting events (e.g. King et al. 2001; Bonamici et al. 2011, 2014, 2015; Bruand et al. 2019). For example, Bonamici et al. (2014) differentiated four generations of titanite with distinct δ18O values and internal textural zoning. Also, consistent results between δ18O in titanite and zircon indicate that titanite is a robust accessory mineral preserving the original magmatic δ18O composition (Bruand et al. 2019).

A thermometer using the Zr content in titanite (Zr-in-titanite) was developed by Hayden et al. (2008), relying on the direct substitution of Zr4+ for Ti4+, whereas pressures can be estimated from the net transfer reaction 2Ca2Al3Si3O12(OH) + TiO2 + SiO2 = 3CaAl2Si2O8 + CaTiSiO5 + H2O (referred to as TZARS; Kapp et al. 2009), using automated calculations in THERMOCALC (Holland and Powell 2011). The Zr-in-titanite thermometer covers a wide temperature range and is pressure dependent, resulting in large pressure uncertainties and rutile activity in rutile-absent rocks, limiting the reliability of temperature estimates. Therefore, modelling titanite within a mineral paragenesis using software such as THERMOCALC allows for pressure estimates with errors less than c. 0.1 GPa (Kohn et al. 2017).

Rutile (Fig. 2j-l; Fig. 6) is the high-temperature TiO2 polymorph (Dachille et al. 1968) and a common accessory mineral that occurs in metamorphic (U) high-pressure (UHP) mafic rocks (Fig. 6a; e.g. Zack et al. 2004b; Triebold et al. 2012; Zack and Kooijman 2017; Böhnke et al. 2019), in moderate-pressures and moderate-temepratures to HP and HT metapelitic rocks (Fig. 6b; e.g. Hart et al. 2018; Gonçalves et al. 2019) and in low-temperature/hydrothermal metapelitic rocks (Fig. 6c; e.g. Plavsa et al. 2018; Salama et al. 2018; Agangi et al. 2019; Porter et al. 2020; Schirra and Laurent 2021). In contrast, rutile occurrence in magmatic rocks is limited to HT and dry alkaline, kimberlite or pegmatitic rocks (e.g. Carruzzo et al. 2006; Cerny et al. 2007). At low-T (LT) and low-P (LP) conditions, rutile is no longer stable, and it is replaced by polymorphs anatase (tetragonal) and brookite (orthorhombic), respectively, whereas rutile HP polymorph is TiO2(II) (Dachille et al. 1968; Jamieson and Olinger 1969). During prograde metamorphism, rutile growth is attributed to breakdown of ilmenite, titanite and biotite (Zack and Kooijman 2017). Rutile popularity as a petrochronometer has exponentially increased due to its multifaceted potential to investigate tectonic processes as single-mineral thermometer (e.g. Zack et al. 2004a; Watson et al. 2006; Tomkins et al. 2007; Luvizotto and Zack 2009), geochronometer (e.g. Mezger et al. 1989a), geospeedometer (Cruz-Uribe et al. 2014; Kohn et al. 2016), as a provenance indicator (e.g. Triebold et al. 2012; Pereira and Storey 2023) and as an isotopic tracer (e.g. Ewing et al. 2011). Additionally, the increasing sensitivity of analytical equipment has enabled dating of low-U phases such as rutile (Luvizotto et al. 2009; Axelsson et al. 2018; Verberne et al. 2019; Moore et al. 2020b).

Fig. 6.

Rutile textures in (a) rutile inclusion in garnet from a HP mafic eclogite from the Sanbagawa belt, SW Japan. (b) HT rutile from an orthopyroxene–cordierite granulite from Madagascar. (c) LT rutile. Left: a BSE image of tabular rutile with a high W core (W-rt) from the Speewah carbonatite, Australia. Right: rutile grain from the Boddington Au–Cu deposit with patchy W zonation and coronitic titanite. (d) Nb–V–Cr ternary discrimination diagram for rutile. Binary plots of rutile (e) U/La v. Zr/Th and (f) Zr/Ba v. Sc/Y trace element ratios from orogenic gold deposits; (g–i) U–Pb data for LA-ICP-MS and SHRIMP analyses for porphyroblastic rutile. Weighted mean age of individual analyses for (g) LA-ICP-MS and (h) SHRIMP data. (i) U–Pb concordia diagram, ellipses representing the 2σ uncertainty; (j–l) TW plots and isochrons/errorchrons combined with Zr-in-rutile thermometry. Source: (a and b) photo courtesy of Pereira; (c) modified after Porter et al. (2020); (d) modified after Hart et al. (2018); (e) and (f) modified after Sciuba and Beaudoin (2021); (g–i) modified after Moore et al. (2020b); (j–l) modified after Bonnet et al. (2022).

Fig. 6.

Rutile textures in (a) rutile inclusion in garnet from a HP mafic eclogite from the Sanbagawa belt, SW Japan. (b) HT rutile from an orthopyroxene–cordierite granulite from Madagascar. (c) LT rutile. Left: a BSE image of tabular rutile with a high W core (W-rt) from the Speewah carbonatite, Australia. Right: rutile grain from the Boddington Au–Cu deposit with patchy W zonation and coronitic titanite. (d) Nb–V–Cr ternary discrimination diagram for rutile. Binary plots of rutile (e) U/La v. Zr/Th and (f) Zr/Ba v. Sc/Y trace element ratios from orogenic gold deposits; (g–i) U–Pb data for LA-ICP-MS and SHRIMP analyses for porphyroblastic rutile. Weighted mean age of individual analyses for (g) LA-ICP-MS and (h) SHRIMP data. (i) U–Pb concordia diagram, ellipses representing the 2σ uncertainty; (j–l) TW plots and isochrons/errorchrons combined with Zr-in-rutile thermometry. Source: (a and b) photo courtesy of Pereira; (c) modified after Porter et al. (2020); (d) modified after Hart et al. (2018); (e) and (f) modified after Sciuba and Beaudoin (2021); (g–i) modified after Moore et al. (2020b); (j–l) modified after Bonnet et al. (2022).

Rutile primarily incorporates high field strength elements (HFSE; i.e. Nb, Ta, Zr, Hf and Cr; Rudnick et al. 2000; Zack et al. 2002; Schmidt et al. 2009). These can be used (i) to characterize micro-scale processes associated with growth of rutile crystals (e.g. Hart et al. 2018; Verberne et al. 2022a); (ii) as a pressure-proxy, such that trace element budget (i.e. Na and Ta) in the melt rutile crystallized in reflects the depth at which partial melting of the source rock occurred (e.g. Foley et al. 2002; Moyen and Stevens 2006; Meyer et al. 2011; Kooijman et al. 2012); or (iii) as geochemical pathfinders for mineralized rocks (Clark and Williams-Jones 2004; Smythe et al. 2008; Pochon et al. 2017; Plavsa et al. 2018; Agangi et al. 2019; Ballouard et al. 2020; Porter et al. 2020; Sciuba and Beaudoin 2021). While HFSE are compatible in rutile, other trace elements including Sr, Th and REEs are incompatible (Klemme et al. 2005; Meyer et al. 2011). Trace element concentrations between the three TiO2 polymorphs (rutile, anatase and brookite) systematically differ, leading to erroneous results when applying the Zr-in-rutile thermometer or Cr and Nb discrimination diagrams (see below) to phases other than rutile (Triebold et al. 2012).

Lithology discrimination schemes from Triebold et al. (2012) and Meinhold et al. (2008) using Cr and Nb (ppm) content in rutile as well as Zr/Hf and Nb/Ta ratios are useful to identify the protolith composition (metapelitic v. metabasic). However, care must be taken when applying this technique for source discrimination in provenance studies due to documented unsystematic Nb/Cr ratios for rutile from amphibolite-facies rocks potentially reflecting TE disturbance during retrogression or prolonged HT metamorphism (Meyer et al. 2011; Kooijman et al. 2012). Additionally, the application of principal component analysis (PCA) on rutile from Precambrian UHT and Phanerozoic HP terranes (Hart et al. 2018), as well as on ore-bearing and barren metamorphic and magmatic rocks (Plavsa et al. 2018; Pereira et al. 2019; Porter et al. 2020) indicates that several TE can be used for lithological discrimination (van Schijndel et al. 2021) and to distinguish rutile from mineralized v. barren rocks. For example, the Nb–V–Cr ternary diagram shows that rutile from HP metamafic rocks is commonly V-rich, whereas in metapelites is more Nb-rich (Fig. 6d; Hart et al. 2018). When associated with mineralized systems, rutile contains anomalous concentrations of V, Sn, Sb, W, Ni, Cu, Cr, Ta, Nb and Fe (e.g. Scott and Radford 2007; Plavsa et al. 2018; Porter et al. 2020), where W and Cr variability is used to discriminate its origin from mineralized or barren rocks (Porter et al. 2020). It is also possible to identify rutile deriving from pegmatitic rocks (high Nb, Ta, Sn) and Au-ore, which are enriched in Sb. This can also be done by combining multi-element clustering of PCA analysis (Porter et al. 2020; Sciuba and Beaudoin 2021). However, due to the systematic differences in trace element contents between the three TiO2 polymorphs, a detailed characterization of the analysed grains is recommended using either EBSD or Raman spectroscopy (Plavsa et al. 2018; Porter et al. 2020; Sciuba and Beaudoin 2021). Multi-variant statistical analysis of rutile TE composition, particularly variations in Sc, REE, Y, Ca, Ba, Th, Zr, U and V, can reflect different metamorphic grades experienced by the country rock (Sciuba and Beaudoin 2021). For example, mafic and ultramafic, lower to middle greenschist-facies country rocks have lower Zr/Th and U/La ratios than intermediate and sedimentary greenschist facies one (Fig. 6e; Sciuba and Beaudoin 2021). A similar trend is observed for relative concentrations of Sc/Y and Zr/Ba, where lower and higher contents reflect lower and higher metamorphic grade, respectively (Fig. 6f; Sciuba and Beaudoin 2021). Also, the identification of localized TE enrichment along twin interface in rutile grains using atom probe tomography is interpreted to occur via volume diffusion during HT metamorphism (Verberne et al. 2022a).

Vry and Baker (2006) calculated the Tc of Pb diffusion in rutile to be between 500 and 540°C, based on natural samples (Fig. 1). For a spherical rutile of 200 µm, experimental data from Cherniak (2000a) predict whole-grain Tc of c. 600°C at an average cooling rate of 2–3°C/Ma, and above c. 640°C for Zr in rutile (2°C/Ma cooling rate; Cherniak et al. 2007; Dohmen et al. 2019). This moderate Tc for rutile implies that it is possible to find crystallization ages in cores of larger rutile crystals, especially if cooling is very fast. The preservation of a crystallization/growth age depends on several factors including max metamorphic temperatures, cooling rates and grain size (Kylander-Clark et al. 2008; Zack and Kooijman 2017; Moore et al. 2020b), and a multi-proxy approach may be needed to distinguish between different metamorphic and deformation stages. For example, Moore et al. (2020b) used rutile Zr thermometry, SHRIMP U–Pb age determination and electron backscatter diffraction (EBSD) microstructural analyses to identified two-stage rutile age populations which were not distinguishable using LA-ICP-MS data alone (Fig. 6g–i). Due to low-U content in rutile (<0.1 ppm; see Zack et al. 2011), it is important to first determine the U content to obtain meaningful metamorphic and/or magmatic ages (Zack et al. 2002, 2004b, 2011). Low-U content can lead to a high common v. radiogenic Pb ratio; therefore, monitoring the 206Pb/208Pb ratio and application of the 208Pb correction method is recommended (Zack et al. 2011). An application of this method allows the investigation of age variations within a single rutile grain, where transects from core to rim give younger ages towards the rim as a result of Pb diffusion during cooling and provide a temperature–time trajectory (Kooijman et al. 2010). In contrast to rutile geochronology obtained with LASS-ICP-MS, recent atom probe investigations indicate evidence for heterogeneous Pb and trace element distribution at the nanoscale (Verberne et al. 2020). Nonetheless, at the microscale (>20 µm), TE variations in a rutile single grain are negligible and concordant U–Pb dating is obtained, indicating that nanoscale defects do not significantly impact the micro-scale analysis (Verberne et al. 2020).

In situ analysis of Hf isotopes in rutile (Sláma et al. 2007; Ewing et al. 2011, 2014) has been applied to trace metasomatic processes in the lithospheric mantle (Choukroun et al. 2005; Aulbach et al. 2008) and recycling of continental material in the mantle (Ewing and Müntener 2018). Despite the relatively low Hf content (<300 ppm Hf), the 176Hf/177Hf of rutile can be accurately measured in situ by LA-MC-ICP-MS, provided care is taken with the analytical protocol and data reduction process (Ewing et al. 2011; Ewing and Müntener 2018). Matrix matched standards and a 176Hf signal intensity above 10 mV are necessary, requiring a large spot size of >160 µm (Yang et al. 2015; Ewing and Müntener 2018). This technique is particularly interesting for rocks that lack zircon but contain rutile, such as mafic lithologies, with low Hf concentrations (Ewing et al. 2011). In contrast, rutile from (U)HT felsic granulites can contain much higher Hf contents, ranging from 20 to 400 ppm (Ewing et al. 2013).

The solubility of ZrO2 in rutile is strongly temperature-dependent, and Zr-in-rutile has been identified as a useful thermometer (ZiR) when the rutile coexists with the appropriate buffer assemblage, i.e. zircon + quartz (Zack et al. 2004a; Watson et al. 2006; Tomkins et al. 2007; Hofmann et al. 2013). Underestimation of the calculated temperatures occurs when the rutile grows in the absence of zircon and/or in partially reset mineral assemblages (Zack et al. 2004a; Harley 2008). Possible biases include micro-inclusions of zircon (Zack et al. 2004a), prograde relict grains or incomplete equilibration with quartz (high Zr-rutile) or zircon (low-Zr rutile) due to slow diffusion along grain boundaries (see discussions in Taylor-Jones and Powell 2015; Kohn et al. 2016; Kohn 2020). Decoupling between Ti-in-zircon and Zr-in-rutile thermometry during UHT metamorphism is recorded in rutile that occurs as inclusions in zircon (Lei et al. 2020). Diffusion of Zr within rutile and Zr loss are closely related to the distribution of Zr, duration of UHT metamorphism and rutile grain size (Dohmen et al. 2019; Lei et al. 2020). The Zr concentration in rutile (Zr-in-rutile) is temperature-sensitive over a large range of geologically significant temperatures (e.g. Zack et al. 2004a; Ewing et al. 2013; Wawrzenitz et al. 2015; Pape et al. 2016; Böhnke et al. 2019; Clark et al. 2019; Moore et al. 2020b; Adlakha and Hattori 2021; Campomenosi et al. 2021; Bonnet et al. 2022). This relationship has been experimentally calibrated by Watson et al. (2006) and Ferry and Watson (2007), and a significant pressure effect has been calibrated by Tomkins et al. (2007). The Zr-in-rutile thermometer was recently refined and now predicts temperatures up to 40°C lower for T ≤550°C, and systematically higher temperatures for T >800°C (Kohn 2020). With the new calibrations, precisions of ±5°C and accuracy of c. ±15°C may be possible, although a variable rutile composition may lead to larger uncertainties (Kohn 2020). Taylor-Jones and Powell (2015) showed that Zr can leave rutile and move along grain boundaries towards existing distal zircon.

Zr-in-rutile temperatures can record HT events, whereas U–Pb in rutile records cooling ages due to low Tc (Fig. 1). Rutile U–Pb ages likely postdate Zr temperatures following high-grade metamorphism and subsequent simple cooling, although a more complex history of episodic cooling and reheating may lead to more significant decoupling between Zr temperatures and U–Pb ages (e.g. Ewing et al. 2015). Bonnet et al. (2022) show the combined use of Zr-in-rutile and U–Pb ages for rutile occurrence in subduction complexes that may be interpreted as crystallization ages for the units that experienced high-pressure, low-temperature metamorphism, but not for the high-grade units (Fig. 6j–l).

The term ‘allanite’ has been used to designate minerals from the allanite subgroup of the epidote group (Armbruster et al. 2006; Mills et al. 2009), which occurs in magmatic (Fig. 2m-o; Fig. 7a), metamorphic (from greenschist- to granulite-facies) and hydrothermal rocks (Janots et al. 2006, 2009; Rubatto et al. 2011; Airaghi et al. 2019). The allanite subgroup comprises a series of REE minerals with an ideal structural formula of Ca(LREE3+)(Al)2 (Fe2+, Fe3+)(SiO4)(Si2O7)O(OH). The presence of LREEs as major constituents, the various element substitutions, the broad P and T stability field and the preservation of growth stages, make allanite one of the burgeoning protagonists in petrochronology (e.g. Rubatto et al. 2011; Manzotti et al. 2018; Airaghi et al. 2019). Compositionally, two main factors are crucial for its use in petrochronology: (i) REEs’ sites can be occupied by Th4+ and U4+, pivotal elements for geochronology (Gieré and Sorensen 2004), and (ii) REEs are exceptional tracers of geological processes (Hermann 2002; Engi 2017). Allanite often exhibits complex chemical zoning (Fig. 7a; Romer and Xiao 2005; Rubatto et al. 2011; Airaghi et al. 2019), requiring investigation via in situ methods (e.g. Burn 2016; Zhang et al. 2022).

Fig. 7.

(a) Quantitative X-ray Map Analyser images (Q–XRMA; Ortolano et al. 2018) used to distinguish generations of metaigneous allanite grains recording different strain rates. (b) Classification diagram to distinguish magmatic from metamorphic allanite based on their Th/U ratio and the fraction of initial 206Pb (f206) values produced using machine learning (Random Forest algorithm). (c) Chondrite-normalized REE patterns of magmatic (pattern) v. metamorphic (filled) allanite: (1) Renna et al. (2007); (2) Gregory et al. (2012); (3) Zhang et al. (2022); (4) Regis et al. (2014); (5) Boston et al. (2017); (6) Vho et al. (2020); (7) Di Rosa et al. (2020). Use of geochemical data to distinguish magmatic and metamorphic allanite. Despite the degree of deformation of allanite crystals presented in (a) the same sequence of chemical zoning can be observed, with progressively deformed allanite crystals exhibiting variations of (d) Ca and (e) Ce contents in relation to LREE contents. Source: (a) from Corti et al. (2020); (b) modified after Di Rosa et al. (2020); (c) modified after Di Rosa et al. (2020).

Fig. 7.

(a) Quantitative X-ray Map Analyser images (Q–XRMA; Ortolano et al. 2018) used to distinguish generations of metaigneous allanite grains recording different strain rates. (b) Classification diagram to distinguish magmatic from metamorphic allanite based on their Th/U ratio and the fraction of initial 206Pb (f206) values produced using machine learning (Random Forest algorithm). (c) Chondrite-normalized REE patterns of magmatic (pattern) v. metamorphic (filled) allanite: (1) Renna et al. (2007); (2) Gregory et al. (2012); (3) Zhang et al. (2022); (4) Regis et al. (2014); (5) Boston et al. (2017); (6) Vho et al. (2020); (7) Di Rosa et al. (2020). Use of geochemical data to distinguish magmatic and metamorphic allanite. Despite the degree of deformation of allanite crystals presented in (a) the same sequence of chemical zoning can be observed, with progressively deformed allanite crystals exhibiting variations of (d) Ca and (e) Ce contents in relation to LREE contents. Source: (a) from Corti et al. (2020); (b) modified after Di Rosa et al. (2020); (c) modified after Di Rosa et al. (2020).

Trace element composition of allanite depends on the interplay between the bulk rock composition and fractionation of these elements during mineral reactions. For instance, a decrease in the LREE in magmatic allanite can be observed in Ca-rich rocks → diorite and granodiorite → granite → syenite (Smye et al. 2014; Engi 2017). The Th/U ratios can be often used as a complementary tool (together with initial Pbc lead values) to distinguish between magmatic (>100) and metamorphic (<50) grains or domains (Fig. 7b; Gregory et al. 2007, 2012; Di Rosa et al. 2020). REEs have also been used to discriminate distinct allanite growth stages in metamorphic and magmatic rocks (Fig. 7c; Manzotti et al. 2018; Corti et al. 2020). Corti et al. (2020) compared composition and internal structure of allanite crystals from metagranitoids recording different strain rates during HP and LT metamorphism, concluding that the matrix and allanite crystals accommodated plastic and brittle deformation, respectively. Allanite resistance to plastic deformation is a noteworthy characteristic, as evidenced by relicts preserved in the sheared eclogites from Monte Mucrone (Stünitz and Tullis 2001; Cenki-Tok et al. 2011). However, fracturing during brittle deformation can disturb its isotopic system (Burn 2016).

Despite the degree of deformation, the allanite grains exhibit the same sequence of chemical zoning pattern (evident for Ca and Ce), but with different textures and LREE contents, suggesting that deformation facilitates the release of LREEs (Fig. 7a, d and e; Corti et al. 2020). Gregory et al. (2012) used Th/U v. La/Sm and Eu/Eu* v. La/Sm discrimination diagrams to distinguish low- from high-temperature magmatic allanite, revealing useful information about the amount of melt present during allanite growth. However, no geochemical ratios have yet been found to systematically discriminate metamorphic from magmatic allanite (Di Rosa et al. 2020). In migmatitic rocks, allanite incorporates significant amounts of Th relative to melt (e.g. Hermann and Rubatto 2009), and it may exhibit intermediate Pbc values, between those for magmatic and metamorphic allanite (Gregory et al. 2012).

The favourable composition of allanite allows dating using the 232Th/208Pb, 238U/206Pb and 235U/207Pb systems. Allanite dating techniques range from single and multi-grain ID-TIMS (e.g. von Blackenburg 1992; Oberli et al. 2004; Smye et al. 2014; López-Moro et al. 2017) to in situ analysis using SHRIMP, SIMS or LA-ICP-MS (e.g. Catlos et al. 2000; Janots et al. 2009; Darling et al. 2012; Regis et al. 2014; Burn et al. 2017; Giuntoli et al. 2018; Liao et al. 2020; Vho et al. 2020). Challenges in the use of allanite reference material for LA-ICP-MS and SHRIMP methods (Burn 2016) include: (i) chemical and isotopic heterogeneities (e.g. Gregory et al. 2007; Boston et al. 2017; Giuntoli et al. 2018); (ii) excess 206Pb in magmatic allanite, which is the most widely used reference material (e.g. BONA, CAP and TARA allanite; Gregory et al. 2007; Burn et al. 2017; Yang et al. 2022); and (iii) use of non-matrix-matched reference materials (e.g. NIST610 glass by McFarlane 2016; Plešovice zircon in Burn et al. 2017). Nonetheless, robust LA-ICP-MS results have been obtained also by using zircon as primary reference material (Darling et al. 2012; Burn et al. 2017; Vho et al. 2020), which has a similar structure to allanite. Rastering (Darling et al. 2012) or a spot analyses routine (Burn 2016) can also be used to minimize matrix sensitivity in LA-ICP-MS analysis. As other accessory minerals, allanite may also incorporate non-radiogenic (204Pb) as well as radiogenic Pb (intermediate nuclei from 238U decay) affecting dates and uncertainties (Romer and Siegesmund 2003; Darling et al. 2012; Engi 2017). Furthermore, the structure of allanite can complicate geochronological procedures, as matrix matching reference materials is needed and the decay from 232Th, 235U and 238U causes structural damage to the crystal lattice (Burn 2016; McFarlane 2016). The destruction of the crystalline structure promotes Pb-loss and/or actinide remobilization with the formation of Th- and U-rich mineral phases (e.g. Barth et al. 1994; Smye et al. 2014). Additionally, deformation and interaction with fluids may open isotopic systems and play an important role in mineral re-equilibration, resorption and precipitation (Radulescu et al. 2009; Airaghi et al. 2019; Corti et al. 2020). Thus, syn-kinematic allanite has been used to date deformation processes at upper to middle crustal levels (Cenki-Tok et al. 2011). Investigations of magmatic allanite from an intensely deformed Mesoproterozoic granite in southern Norway indicate that higher mobility of Th than Pb during deformation processes results in the U–Pb system being more reliable than the Th–Pb system (Burn 2016). Nevertheless, allanite crystal structure may protractedly recover by annealing (Karioris et al. 1981), forming preserved (non-metamict) crystals that may yield younger dates (Catlos et al. 2000). More recently, allanite has been used as primary and secondary reference material for epidote dating (Peverelli et al. 2022).

When dating allanite, one of the three approaches discussed by Burn (2016) and Engi (2017) should be utilized to deal with initial or Pbc correction. (i) Consider Pbc evolution models. This approach is more often used for magmatic (Barth et al. 1994) than metamorphic (e.g. Radulescu et al. 2009; Rubatto et al. 2011) rocks due to the main issue of using global (silicate Earth or mantle) evolution models. Th, U and Pb contents differ from rock to rock, and their distribution within minerals is heterogeneous. These factors depend on several variables such as local effective bulk composition and fluids availability (Lanari and Engi 2017). (ii) Measuring Pbc in phases that coexist with allanite (e.g. Cenki-Tok et al. 2014), which can be hampered by the difficulty in interpreting coexisting phases. Finally, (iii) the ‘intercept approach’, in which intercepts from uncorrected TW and 206Pbc normalized Th–Pb isochron diagrams are used to estimate the initial Pbc (e.g. Janots and Rubatto 2014; Airaghi et al. 2019). Gregory et al. (2012) indicate that igneous allanite tends to have smaller amounts of non-radiogenic 208Pb than high-grade metamorphic rocks, whereas allanite crystals formed at subsolidus conditions exhibit the highest non-radiogenic 208Pb values.

Allanite enrichment in LREE and Sr allows for both in situ Nd and Sr isotopic-based petrogenetic information to be combined with U–Th–Pb dating, making allanite an important petrogenetic tool (Hoshino et al. 2007). Heterogeneous Pb and Sr isotopic concentrations in allanite are demonstrated to be inherited from precursor minerals involved in allanite-producing metamorphic reactions (Romer and Xiao 2005). Zhang et al. (2022) combined U–Th–Pb dating and Nd isotopes of allanite with U–Pb–Hf analyses of zircon, demonstrating good correlations between the two systems to investigate crustal formation and evolution. Nd isotopes in allanite analysis were used by Su et al. (2021) to unravel the hydrothermal history of an iron oxide copper–gold deposit from ore formation (multi-source) to the post-ore overprinting tectono-thermal events.

Monazite (LREE,Y,Th,Ca,Si)PO4 is a REE-rich phosphate mineral (Fig. 2p-r) that occurs in a variety of rock compositions and from diagenetic (Evans and Zalasiewicz 1996; Pereira et al. this volume, in press) to granulite-facies conditions (Black et al. 1984). Monazite is common in peraluminous granites (e.g. Montel 1993; Förster 1998), syenite, granitic and quartz veins (e.g. Piechocka et al. 2017) and carbonatitic plutons (e.g. Anenburg et al. 2021; Kamenetsky et al. 2021), and can exhibit various textures (from sector, firtree, oscillatory to lobate; Fig. 8a). However, under certain fluid conditions, in particular alkali-bearing fluids, monazite undergoes coupled dissolution–reprecipitation (CDR) (Vavra and Schaltegger 1999; Harlov and Hetherington 2010; Harlov et al. 2011; Kelly et al. 2012; Taylor et al. 2014, 2016; Bosse and Villa 2019; Weinberg et al. 2020; Salminen et al. 2022). CDR of monazite leaves very distinctive, lobate textures but can also modify U–Pb systematics (Fig. 8b; e.g. Vavra and Schaltegger 1999; Taylor et al. 2014; Blereau et al. 2016; Prent et al. 2020), cause U and Th loss (Williams et al. 2011) or Th gain (Harlov et al. 2011). Monazite is a common accessory mineral in metapelitic rocks experiencing low-amphibolite facies metamorphism (e.g. Rubatto 2002) to high-grade granulites, migmatites and charnockites (Fig. 8b; e.g. Laurent et al. 2018; Dev et al. 2021; Williams et al. 2022), whereas it is less common in HP rocks (Finger and Krenn 2007). Whilst geochronological disruption causes resetting of the oldest growth of the mineral, the reactivity of monazite is useful for tracing or determining the timing of fluid activity, reactivation of metamorphic processes and/or ore genesis. On the other hand, xenotime (Y, HREE)PO4 is a HREE carrier, which also makes it a critical source of HREEs (Strzelecki et al. 2022). Xenotime occurs in metapelitic, granitic and carbonatitic rocks, and is less abundant in mafic and calc-silicate rocks (Spear and Pyle 2002). Recent phase equilibria modelling studies have demonstrated that xenotime usually occurs at pressures lower than 8 kbar and temperatures lower than 750°C (e.g. Shrestha et al. 2019). In contrast, monazite is found to be stable at higher pressures and temperatures, where water availability strongly controls the monazite P–T stability field and its preservation along the prograde path at much higher P–T conditions than previously envisaged (Larson et al. 2022).

Fig. 8.

(a) Magmatic textures of monazite crystals from (1) Barrote et al. (2020); (2) Volante et al. (2020a); (3) Piechocka et al. (2017). (b) Metamorphic textures of monazite and xenotime crystals: (4) Volante et al. (2020c); (5) Photo courtesy of Cutts; (6) Cutts et al. (2018); (7) Blereau et al. (2016); (8) Laurent et al. (2018); (9) Manzotti et al. (2018); (10) Barrote et al. (2020). (c) Representative chondrite normalized REE patterns for different types of monazite and xenotime: (1) Manzotti et al. (2018); (2) Rasmussen and Muhling (2007); (3) Rubatto et al. (2013); (4) Taylor et al. (2014); (5) Buick et al. (2010). (d) U–Pb data for monazite grains from sapphirine–cordierite UHT gneisses in Norway. Monazite ages are colour-coded based on Y2O3 content. Source: (d) after Laurent et al. (2018).

Fig. 8.

(a) Magmatic textures of monazite crystals from (1) Barrote et al. (2020); (2) Volante et al. (2020a); (3) Piechocka et al. (2017). (b) Metamorphic textures of monazite and xenotime crystals: (4) Volante et al. (2020c); (5) Photo courtesy of Cutts; (6) Cutts et al. (2018); (7) Blereau et al. (2016); (8) Laurent et al. (2018); (9) Manzotti et al. (2018); (10) Barrote et al. (2020). (c) Representative chondrite normalized REE patterns for different types of monazite and xenotime: (1) Manzotti et al. (2018); (2) Rasmussen and Muhling (2007); (3) Rubatto et al. (2013); (4) Taylor et al. (2014); (5) Buick et al. (2010). (d) U–Pb data for monazite grains from sapphirine–cordierite UHT gneisses in Norway. Monazite ages are colour-coded based on Y2O3 content. Source: (d) after Laurent et al. (2018).

Both monazite and xenotime are REE-rich mineral phases, containing a critical amount of REEs in addition to Y (Schulz 2021 and references therein). Monazite REE patterns are usually characterized by a negative slope, whereas xenotime commonly exhibits a positive REE pattern more like zircon (Fig. 8c). Sr-enrichment in monazite has been interpreted to reflect HP monazite growing in the absence of feldspars (Finger and Krenn 2007). Additionally, a REE signature like HP zircon, with low HREE and absence of a negative Eu anomaly, has been identified in monazite grains from the Kokchetav UHP rocks in Kazakhstan, and UHP rocks in Norway, suggesting this as a geochemical signature for HP monazite (Hermann and Rubatto 2014; Hacker et al. 2015). Furthermore, REEs in monazite have been utilized as geochemical discriminators with U–Pb geochronology to distinguish between magmatic and metamorphic monazite in complex deformed and metamorphosed terranes (e.g. Pe-Piper et al. 2014; Prent et al. 2019; Itano et al. 2020).

Like zircon, monazite has a number of partitioning relationships that can be useful in integrating various data sources to the evolution of a sample's mineral paragenesis, including monazite/melt (Yurimoto et al. 1990; Ward et al. 1992; Bea et al. 1994; Stepanov et al. 2012), monazite/xenotime (Andrehs and Heinrich 1998), monazite/K-feldspar (Villaseca et al. 2003) and monazite/garnet (Hermann and Rubatto 2003). The most applied technique is the partitioning of REEs between monazite and garnet since these minerals frequently occur together (e.g. Buick et al. 2006; Rubatto et al. 2006; Kylander-Clark et al. 2013; Mottram et al. 2014; Taylor et al. 2014; Blereau et al. 2016; Hagen-Peter et al. 2016; Hacker et al. 2019; Warren et al. 2019). Despite monazite being relatively poor in HREEs compared to zircon and garnet, the presence of garnet still impacts the relative concentration of HREEs in monazite. Monazite growing syn- to post-garnet or modified in the presence of garnet (see also Discussion) typically shows a reduction in HREEs and Y compared to monazite grown in the absence of garnet (Hermann and Rubatto 2003; Rubatto et al. 2006). Recent studies have demonstrated more complicated processes associated with HREE partitioning in monazite, where partitioning coefficients between monazite and garnet within investigated metapelitic rocks did not reproduce the expected values (e.g. Larson et al. 2019, 2022; Shrestha et al. 2019). Their temperature dependence is also found to be more relevant than initially envisaged (Hacker et al. 2019; Warren et al. 2019; Jiao et al. 2021). Also, recent work has highlighted by modelling Sm–Eu–Gd partitioning in suprasolidus systems that even though fractionation of Eu by feldspar growth can dominantly control the Eu budget, at equilibrium, other factors such as oxygen fugacity were shown to play an important role (Holder et al. 2020).

Monazite has proved a reliable geochronometer using the U–Pb system (Parrish 1990; Williams et al. 2007). Generally, monazite does not incorporate Pb into its crystal structure but has high Th (often several wt%) and U (several thousand ppm), meaning that it has low Pbc and high radiogenic Pb. Monazite is also inferred to have a high Tc of up to 900°C (Fig. 1; Cherniak et al. 2004). However, monazite can be quite reactive (via CDR, see above), resulting in resetting of ages in fluid-dominated systems (Seydoux-Guillaume et al. 2002), whereas monazite in dry rocks (i.e. granulites) has been found to preserve detrital ages (e.g. Suzuki and Adachi 1994; Cutts et al. 2013; Guo et al. 2020). The high U and Th contents of monazite allow dating via chemical U–Th–Pb dating using EPMA (e.g. Suzuki and Adachi 1994; Montel et al. 1996, 2018). The advantages of this approach include in situ, non-destructive analysis and a small spot size allowing monazite grains <20 µm to be targeted (Ning et al. 2019; Williams and Jercinovic 2002). The main pitfall is that Pbc correction is not possible (Williams et al. 2017), involving the assumption that the analysed grain is concordant and contains no Pbc. The CHIME (chemical Th–U-total Pb isochron method) dating method is like chemical dating but targets multiple spots in an age domain to produce a ‘pseudo-isochron’. This method also uses compositional criteria to determine if the monazite age is concordant, resulting in a more robust method than traditional chemical dating (Suzuki and Kato 2008). Improvements in EPMA sensitivity and the method make this a powerful technique moving forward (Konečný et al. 2018; Montel et al. 2018; Ning et al. 2019). Prior to EPMA dating, many studies utilized ID-TIMS geochronology of monazite (i.e. Smith and Barreiro 1990), and later the LA-ICP-MS approach (Machado and Gauthier 1996; Poitrasson et al. 2000). Due to its high U and Th contents, monazite is extremely amenable to LA-ICP-MS, where small spot sizes (8–12 µm) can be used. Similarly, geochronology of xenotime can be analysed in situ by both SIMS (Cross 2009; Fielding et al. 2017) and LA-ICP-MS (Lawley et al. 2015; Simpson et al. 2021) to collect texturally contextualized isotopic dates. Currently, no matrix-matched reference materials for xenotime are available as no homogeneous natural xenotime has been found. Therefore, first-order matrix corrections have been applied on xenotime reference materials such as z6413 (Stern and Rayner 2003) and MG-1 (Fletcher et al. 2004) for U–Pb geochronology, whereas glass NIST 610 is currently used as primary reference material for in situ Lu–Hf dating (Simpson et al. 2021). Due to the wealth of information provided by REE data, TEs are commonly collected either separately or simultaneously with geochronological data using one or two mass spectrometers (Holder et al. 2013; e.g. Kylander-Clark et al. 2013; Hacker et al. 2015; Volante et al. 2020c; Barrote et al. 2022b). Nanoscale geochronological analysis of monazite (Fougerouse et al. 2020, 2021) and xenotime (Joseph et al. 2021) via atom probe (ATP) is also possible. Analytical development has allowed to investigate diffusion and migration of atoms along monazite grain boundaries and/or crystal defects, and relates e.g. intracrystalline deformation with age resetting due to radiogenic Pb loss (Fougerouse et al. 2021). Great potential to determine ages by using the ATP was shown also on small xenotime crystals (Joseph et al. 2021). In situ studies allow monazite ages to be directly related to mineral textures present in the rock, so the age of deformation can be attributed to metamorphic events in complex terranes (e.g. Smith and Barreiro 1990; Foster et al. 2002; Štípská et al. 2015; Piechocka et al. 2017; Prent et al. 2019, 2020; Jiao et al. 2020a; Volante et al. 2020c).

In granitic rocks, high LREE minerals, like monazite, are great competitors for trace elements, including Sm–Nd and REE, in the host magmas (e.g. Fisher et al. 2017; Hammerli and Kemp 2021). The Sm–Nd isotope tracer system in monazite can complement the more commonly applied Lu–Hf system in zircon (e.g. Fisher et al. 2017; Martin et al. 2020; Mulder et al. 2021; Barrote et al. 2022a; Volante et al. 2022) and whole-rock Sm–Nd and/or Lu–Hf (e.g. Mark 2001; Hoffmann et al. 2011; Caxito et al. 2021), providing insights into the formation, evolution and differentiation of the continental crust (Hammerli and Kemp 2021). Significant contributions have been made to improve in situ Sm–Nd precision using LASS-MC-ICP-MS (e.g. Barrote et al. 2022a) to achieve a comparable level of precision to Lu–Hf in zircon (e.g. Fisher et al. 2011, 2020; Goudie et al. 2014; Spencer et al. 2020). Furthermore, monazite is less prone to weathering than other Sm–Nd-bearing major minerals (e.g. plagioclase), making it a valuable tool to investigate primary isotopic signatures (e.g. Barrote et al. 2022a).

To estimate minimum magmatic temperatures, monazite thermometry relies on whole-rock REE content in peraluminous and metaluminous granitic rocks (Montel 1993; Plank et al. 2009; Stepanov et al. 2012). While the Montel (1993) equation is calibrated only at low pressure, Stepanov et al. (2012) corrected it by extending experiments to higher pressures. Recent studies have presented the importance of applying variable H2O contents for granitic rocks to determine monazite saturation temperatures, with this approach producing consistent zircon and monazite saturation temperatures across the same granitic samples (Volante et al. 2020a). The Y-HREE fractionation between monazite and xenotime results in an asymmetric miscibility gap in YPO4-(REE)PO4. This allows for the Y content of monazite to be used as a geothermometer, which is largely independent of pressure (e.g. Gratz and Heinrich 1997, 1998; Heinrich et al. 1997; Andrehs and Heinrich 1998). This thermometer can only be used when monazite has grown in the presence of xenotime. However, identifying and ascertaining that these two phases grew in equilibrium can be challenging (Pyle and Spear 1999). Like garnet, xenotime also preferentially incorporates HREE. Hence, xenotime is usually consumed during prograde metamorphism by garnet growth, and it commonly reappears during garnet breakdown at post-peak conditions or along the retrograde path (e.g. Hallett and Spear 2015). In this view, compositional variations in monazite (e.g. Regis et al. 2016; Manzotti et al. 2018) have been crucial to assess xenotime saturation within the rock system (e.g. Krenn and Finger 2010). For example, Y-rich monazite domains have been suggested to be plausible targets to use for thermometry (e.g. Viskupic and Hodges 2001; Krenn et al. 2012; Laurent et al. 2018). Y content can also be used to distinguish between different monazite age populations (i.e. Fig. 8d), and can be combined with thermometry to indicate the temperature–time evolution of the sample (e.g. Laurent et al. 2018). Some attempts were made to thermodynamically model monazite and xenotime (e.g. Kelsey et al. 2007, 2008; Spear and Pyle 2010; Shrestha et al. 2019). However, this approach only works with low-Ca bulk-rock compositions and requires the simplification of a complicated system.

Apatite Ca5(PO4)3(F,Cl,OH) is a calcium phosphate (Fig. 2s-u; Fig. 9) that is often present as an accessory mineral in several rock types, including various igneous rocks (ultramafic–felsic) to metamorphic and sedimentary, and even in meteorites (e.g. Harlov 2015; McCubbin and Jones 2015; Webster and Piccoli 2015). Apatite can readily incorporate large amounts of TE (e.g. U, Th, REE, Y, Sr), which replace Ca in the crystal lattice with charge compensation mechanisms (Engi 2017), making apatite important for the trace element budget of a rock. In addition, apatite commonly contains significant amounts of water and/or halogens and volatile species (e.g. F and Cl; Harrison and Watson 1984; Piccoli and Candela 2002), and records the evolution of metasomatic and hydrothermal fluids (e.g. Harlov 2015; Lin et al. 2023). Combined, the geochemical characteristics of apatite make it suitable for many geochemical and petrological applications, with measurements being possible with a large variety of instruments (e.g. SIMS, (LA)-ICP-MS, (LA)-MC-ICP-MS, EPMA), imaging with BSE (Fig. 9a) and TE mapping with LA-ICP-MS (Fig. 2r). Depending on its crystallization conditions, apatite can exhibit various textures visible in CL or BSE images (e.g. Mühlberg et al. 2021) and geochemical signatures that make this mineral a very interesting proxy for rock formation and evolution as well as for chronological information (Chew et al. 2011; Hammerli et al. 2014; Kirkland et al. 2017; Antoine et al. 2020; Fisher et al. 2020; Prent et al. 2020; Paul et al. 2021).

Fig. 9.

(a) BSE images for magmatic apatite. (b) Distinct REE patterns for apatite from different types of granitoids. (c) Discrimination diagram for apatite from various granitoids. BADR, basalt–andesite–dacite–rhyolite series. (d) Chondrite-normalized La/YbN v. YbN diagram of apatite from post-Archean granites and Archean TTGs. (e) Lithological discrimination diagram. Abbreviations: ALK, alkali-rich igneous rocks; IM, mafic I-type granitoids and mafic igneous rocks; LM, low- and medium-grade metamorphic and metasomatic; HM, partial-melts/leucosomes/high-grade metamorphic; S, S-type granitoids and high aluminum saturation index (ASI) ‘felsic’ I-types; UM, ultramafic rocks including carbonatites, lherzolites and pyroxenites. (f) Binary diagram U–Pb v. Lu–Hf ages (Ma) in apatite. (g) 87Sr/86Sr v. age diagram showing how initial apatite Sr isotope signatures can be interpreted in terms of source lithology, derived from time-integrated Rb/Sr ratios, and resulting model age (intersection between grey fans and BSE evolution). Depleted MORB (Mid-ocean ridge basalt) mantle (DMM). Source: (b) and (c) modified after Bruand et al. (2020); (d) modified after Antoine et al. (2020); (e) modified after O'Sullivan et al. (2020); (f) rock lithologies following discrimination diagram from O'Sullivan et al. (2020); (g) modified after Emo et al. (2018).

Fig. 9.

(a) BSE images for magmatic apatite. (b) Distinct REE patterns for apatite from different types of granitoids. (c) Discrimination diagram for apatite from various granitoids. BADR, basalt–andesite–dacite–rhyolite series. (d) Chondrite-normalized La/YbN v. YbN diagram of apatite from post-Archean granites and Archean TTGs. (e) Lithological discrimination diagram. Abbreviations: ALK, alkali-rich igneous rocks; IM, mafic I-type granitoids and mafic igneous rocks; LM, low- and medium-grade metamorphic and metasomatic; HM, partial-melts/leucosomes/high-grade metamorphic; S, S-type granitoids and high aluminum saturation index (ASI) ‘felsic’ I-types; UM, ultramafic rocks including carbonatites, lherzolites and pyroxenites. (f) Binary diagram U–Pb v. Lu–Hf ages (Ma) in apatite. (g) 87Sr/86Sr v. age diagram showing how initial apatite Sr isotope signatures can be interpreted in terms of source lithology, derived from time-integrated Rb/Sr ratios, and resulting model age (intersection between grey fans and BSE evolution). Depleted MORB (Mid-ocean ridge basalt) mantle (DMM). Source: (b) and (c) modified after Bruand et al. (2020); (d) modified after Antoine et al. (2020); (e) modified after O'Sullivan et al. (2020); (f) rock lithologies following discrimination diagram from O'Sullivan et al. (2020); (g) modified after Emo et al. (2018).

Igneous apatite has reasonably similar total REE contents regardless of the protolith lithology (e.g. Chu et al. 2009), except for apatite in alkali-rich rocks, which are commonly significantly REE-rich (e.g. Zirner et al. 2015). However, relative REE concentrations, and Y and Sr, are variable within different rock-types and particularly effective for discriminating magma types (e.g. Bea 1996; Belousova et al. 2001; Bruand et al. 2020). Negative Eu anomalies are present in apatite from most igneous rocks except ultramafics (e.g. Chakhmouradian et al. 2017; Bruand et al. 2020), and the magnitude of Eu anomalies is typically positively correlated with SiO2 content. In contrast, Sr content in apatite is negatively correlated with SiO2 content, with the highest Sr contents being observed in ultramafic rocks (Ihlen et al. 2014) and the lowest Sr contents in felsic rocks (Sha and Chappell 1999). Also, Sr in apatite strongly correlates with Sr content in the corresponding whole-rock (Belousova et al. 2001; Jennings et al. 2011; Bruand et al. 2014). Variations in Sr content in TTGs are interpreted to directly correlate with the melting depth of the TTGs’ source, where a deep source is in equilibrium with garnet and rutile, but no plagioclase (Moyen and Martin 2012). Moreover, apatite chemistry can be used as a tracer of different magmatic petrogenetic processes, reflecting geodynamic changes during crustal evolution (Bruand et al. 2020) such as (La/Yb)N v. YbN discrimination diagram for TTGs (Antoine et al. 2020; Fig. 9b–d).

Metamorphic apatite also exhibits contrasting signatures depending on its metamorphic grade and its textural context (e.g. O'Sullivan and Chew 2020; Prent et al. 2020). Total REE content increases with the metamorphic grade (e.g. El Korh et al. 2009), such that in high-grade rocks it is indistinguishable from that of igneous apatite (e.g. Bingen et al. 1996). In addition, negative Eu anomalies characterize apatite growing in low-grade metamorphic rocks (Henrichs et al. 2018). Also, metamorphic apatite contains higher Sr concentrations in greenschist- and blueschist-facies rocks (e.g. Nishizawa et al. 2005), and lower ones in migmatites and granulite-facies rocks (e.g. Nutman 2007). In felsic igneous and metamorphic rocks, apatite (La/Lu)N ratio is commonly ≤1, whereas in all other rocks (La/Lu)N ratio is >1, with the highest values recorded in ultramafic rocks (e.g. O'Reilly and Griffin 2000). A discrimination diagram using Sr/Y v. ΣLREE biplots from various igneous and metamorphic/metasomatic rocks (Fig. 9e) can be used to statistically categorize source lithologies (O'Sullivan et al. 2020). Additional discriminant diagrams based on the enrichment in LREEs and MREEs, Eu anomalies, the tetrad effect, Mn and Sr contents, total REEs and Y in apatite can be used to highlight the variability between different types of mineralization (Decrée et al. 2023). For example, bell-shaped REE patterns defined by MREE enrichment and positive Eu anomalies were found to be a unique indicator of hydrothermal apatite formed in a mineralized, reducing hydrothermal system (e.g. Krneta et al. 2018; Lin et al. 2023).

Apatite has long been the target for geochronological studies (e.g. Schoene and Bowring 2007; Chew et al. 2011; Glorie et al. 2022) either using the U–Th–Pb or Lu–Hf isotope system (e.g. Barfod et al. 2005; Chew et al. 2011; Simpson et al. 2021; Glorie et al. 2023), or Sm–Nd (Fisher et al. 2020). Apatite commonly incorporates significant amounts of Pbc that vary from limited amounts in igneous apatite to larger amounts in hydrothermal apatite crystals (e.g. Kirkland et al. 2017). Low-grade metamorphic apatite typically provides poor precision for U–Pb dating due to low U concentrations (<5 ppm) and high initial Pbc contents, while apatite from igneous rocks is typically U-rich (>20 ppm) and has lower Pbc, leading to more precise U–Pb ages (e.g. Henrichs et al. 2018, 2019; O'Sullivan et al. 2018). Because of its Pbc content, apatite can often produce U–Pb dates that are strongly discordant (also see Titanite section). This Pbc presents a particular challenge for young samples that have had little time to accumulate substantial radiogenic Pb (Pb*), or for apatite grains with low concentrations of U. Following to the development of apatite U–Th–Pb age reference materials (e.g. Chew et al. 2011; Thomson et al. 2012; Apen et al. 2022; Lana et al. 2022) and data reduction schemes employing 208Pb-, 207Pb- or 204Pb-based Pbc corrections to age reference materials and unknowns, it is now possible to routinely date apatite both precisely and accurately (e.g. Andersson et al. 2008, 2022; Chew et al. 2011; Thomson et al. 2012; Antoine et al. 2020; Prent et al. 2020; Glorie et al. 2022). However, some complications in data interpretation can occur in apatite crystals that have undergone metamorphism due to diffusion effects (e.g. Paul et al. 2019). Apatite U–Pb dates may be reset by deformation and fluids, causing the age to represent the last major deformation event or latest dissolution–precipitation and/or chemical exchange (Odlum et al. 2022).

Recent advances including the development of LA-ICP-MS/MS technology allowed resolution of isobaric interferences for Lu–Hf in situ dating of apatite crystals (e.g. Barfod et al. 2003, 2005; Larsson and Söderlund 2005; Simpson et al. 2021; Gillespie et al. 2022; Glorie et al. 2022, this volume, in press) and Sm–Nd (e.g. Hammerli et al. 2014, 2019; Doucelance et al. 2020; Fisher et al. 2020), which is extremely useful to investigate post-metamorphic cooling history and/or compare crystallization ages. Apatite is commonly combined with other mineral phases that exhibit various parent/daughter ratios (i.e. Lu/Hf and Sm/Nd) to obtain a more robust isotope isochron (e.g. Hammerli et al. 2014; Laurent et al. 2017; Simpson et al. 2021). Advantages of the Lu–Hf isotope system in apatite over U–Pb include higher Tc of the former (Fig. 1), resulting in an age closer to the apatite crystallization age (e.g. Henrichs et al. 2019). Comparisons of the two systems can have significant advantages (Fig. 9f; Glorie et al. 2022).

The Sm–Nd isotope system in apatite can provide insights into timing and source of apatite host-rock (e.g. Fisher et al. 2020). Among stable isotopes, oxygen isotopes appear to be proxies for magma sources and to record fluid circulation and metamorphism processes (e.g. Bruand et al. 2019). Chlorine and hydrogen isotopes have also been used as proxies for volatilization/condensation processes (e.g. Potts et al. 2018; Wudarska et al. 2020). Isotope systematics such as 87Rb–87Sr can be measured using either (LA-)MC-ICP-MS or LA-ICP-MS/MS, TIMS or SIMS, and coupled to age information (Fig. 9g; Emo et al. 2018; Ravindran et al. 2020; Gillespie et al. 2021). The 87Sr/86Sr ratio of apatite may be used together with K-rich or Rb-bearing minerals such as micas for in situ Rb–Sr geochronology to create a combined isochron (Olierook et al. 2020).

There is no currently developed thermometer for apatite like those for zircon (Ti-in-zircon) or rutile (Zr-in-rutile), but the U–Pb Tc (350–570°C; Fig. 1) can be used as an indirect proxy for tracking metamorphic or magmatic cooling in combination with other chronometers in apatite (Fig. 1; e.g. Cherniak 2000b; Barfod et al. 2003, 2005; Cochrane et al. 2014; Chew and Spikings 2015; Kirkland et al. 2018; Ferreira et al. 2022).

Geochronological and geochemical data retrieved from metamorphic accessory minerals are commonly integrated with information from the major mineral paragenesis, P–T constraints from phase equilibrium diagram calculations and microstructures. Figure 10 reflects an example of two synoptic, theoretical P–T–t evolutions where petrochronological information of metamorphic accessory minerals such as monazite, zircon and garnet is linked to P–T information. The first P–T path (Fig. 10a) represents common clockwise granulite-facies metamorphism of a pelitic protolith hosting inherited zircon (Fig. 10a, i). During prograde metamorphism, garnet growth begins and inherited zircon grains start to recrystallize with increasing temperatures (Fig. 10a ii). This prograde part of the metamorphic evolution is often difficult to accurately constrain due to open-system processes, overprinting and/or exceeding closure temperatures of earlier phases. Zircon has been shown to grow along the prograde path connected to the movement of locally-derived melts (e.g. Harley and Nandakumar 2014; Harley 2016; Weinberg et al. 2020) as well as injected, externally-derived melts (Andersson et al. 2002; Flowerdew et al. 2006; Wu et al. 2007). Once the sample crosses the solidus, partial melting begins. Inherited zircons may be consumed and or dissolved in the melt that becomes enriched in zirconium (Watson and Harrison 1984; Boehnke et al. 2013; Gervasoni et al. 2016), forming embayment and relict core textures in zircon. Monazite might also grow along the prograde and retrograde path (e.g. Rubatto 2002; Shrestha et al. 2019; Larson et al. 2022), but also during (Kelsey et al. 2008; Rubatto et al. 2009; Larson et al. 2022) and/or just after max T conditions are reached (Fig. 10a iii, v, vii; Clark et al. 2014; Shrestha et al. 2019), making it a more suitable time-capsule than zircon by preserving a greater P–T–t window at subsolidus conditions (e.g. Ambrose et al. 2015; Mottram et al. 2015; Hacker et al. 2019; Shrestha et al. 2020; Larson et al. 2022). Recent works modelled the P–T stability fields for monazite and xenotime, with monazite present during prograde growth of garnet, max T(/P) conditions and along the retrograde path during garnet breakdown at both sub- and suprasolidus conditions (Shrestha et al. 2019). In contrast, xenotime appears only at low P–T conditions during the early stages of the prograde path and/or along the retrograde path, when it would start to assimilate most of the Y and HREE, resulting in a decrease in monazite proportions (Shrestha et al. 2019). Prograde garnet sees further growth at suprasolidus conditions, combined with an additional generation of peritectic garnet (Fig. 10a iv) that could be distinguished based on different REE patterns and inclusions (Fig. 3f). Trace element partitioning is mostly utilized for max T(/P) to retrograde portion of the P–T–t history as this information is most likely preserved. For example, Figure 10a shows that garnet growth and metamorphic ‘soccer ball’ zircon (e.g. Vavra and Schaltegger 1999; Blereau et al. 2016; Taylor et al. 2016) grow in the same geochemical system and have a near 1:1 partitioning of M-HREE. How well these partitioning relationships are preserved depends on the conditions of metamorphism and how these affect the equilibration volume of the sample. Typically, this is largest at max T conditions, where rates of diffusion and potential fluids and/or melt increase. Approaching the solidus on the retrograde path, Zr-bearing phases may breakdown, e.g. garnet to cordierite (Fig. 10a vi) resulting in additional zircon growth (e.g. Fraser et al. 1997; Degeling et al. 2001; Wu et al. 2007; Kelsey et al. 2008), from rims to completely new grains (Fig. 10a vii). A subsequent fluid event, e.g. upon crystallization of local melt, can modify monazite and zircon textures, perturbing age and geochemical data (Poitrasson et al. 2000; Seydoux-Guillaume et al. 2012; Kröner et al. 2014; Taylor et al. 2014; Blereau et al. 2016; Prent et al. 2019). In more complex garnet textures, such as the unusual atoll garnet (e.g. Jonnalagadda et al. 2017; Kulhánek et al. 2021; Godet et al. 2022; Massonne and Li 2022), TE content recorded by subsequent growth of concentric garnet rings can also provide useful information about the interchange of TE within the system due to growth and breakdown of other accessory phases (Godet et al. 2022). For example, by using LA-ICP-MS trace element mapping, Godet et al. (2022) reported enrichment in V and Ti in the garnet inner rim compared to the core, which was attributed to rutile breakdown, whereas Cr, Y, LREE and MREE enrichment in the outer rim was interpreted to reflect allanite and monazite breakdown.

Fig. 10.

Synoptic P–T evolutions of a metapelite in different hypothetical scenarios with associated types of zircon, monazite, xenotime, garnet growth/breakdown textures and potential REE patterns. (a) Singular clockwise P–T evolution: (i) inherited zircon; (ii) primary garnet and recrystallization of inherited zircon; (iii) relict zircon core, initial monazite growth; (iv) growth of garnet rim and peritectic garnet; (v) neocrystallized ‘soccer ball’ zircon, additional monazite and recrystallization of zircon; (vi) garnet breakdown; (vii) melt crystallization; (viii) late fluid modification. (b) A residual protolith along a clockwise polymetamorphic P–T path: (i) recrystallization of zircon; (ii) prograde monazite growth in the presence of relict garnet; (iii) at the elevated solidus, minor melting causes additional monazite but no new zircon growth; (iv) garnet breakdown, final monazite growth and complete ‘ghost zoning’ to recrystallized zircon. After a localized hydrous retrograde event, the hydrated areas follow a secondary P–T path; (v) melt modifies zircon and monazite to relict cores stable with relict garnet grains; (vi) new zircon and monazite; (vii) final melt crystallization. (c) Synthetic Tera–Wasserburg plot with interpretation based on the nature of analysed textures; § enlarged inset of younger ages. Source: (a–c) modified after Blereau (2017).

Fig. 10.

Synoptic P–T evolutions of a metapelite in different hypothetical scenarios with associated types of zircon, monazite, xenotime, garnet growth/breakdown textures and potential REE patterns. (a) Singular clockwise P–T evolution: (i) inherited zircon; (ii) primary garnet and recrystallization of inherited zircon; (iii) relict zircon core, initial monazite growth; (iv) growth of garnet rim and peritectic garnet; (v) neocrystallized ‘soccer ball’ zircon, additional monazite and recrystallization of zircon; (vi) garnet breakdown; (vii) melt crystallization; (viii) late fluid modification. (b) A residual protolith along a clockwise polymetamorphic P–T path: (i) recrystallization of zircon; (ii) prograde monazite growth in the presence of relict garnet; (iii) at the elevated solidus, minor melting causes additional monazite but no new zircon growth; (iv) garnet breakdown, final monazite growth and complete ‘ghost zoning’ to recrystallized zircon. After a localized hydrous retrograde event, the hydrated areas follow a secondary P–T path; (v) melt modifies zircon and monazite to relict cores stable with relict garnet grains; (vi) new zircon and monazite; (vii) final melt crystallization. (c) Synthetic Tera–Wasserburg plot with interpretation based on the nature of analysed textures; § enlarged inset of younger ages. Source: (a–c) modified after Blereau (2017).

The second P–T–t evolution (Fig. 10b) represents a more residual scenario, where the protolith has already been partially melted. In this case, recrystallization and preservation of existing major and accessory minerals are promoted due to the lack and/or reduction in volume of partial melt being produced, which also limits the generation of new assemblages (Fig. 10b, i, ii, iii; Bea and Montero 1999; White and Powell 2002). Monazite and zircon contrast in their behaviour under these conditions. Monazite may grow in larger amounts than zircon, despite being in a melt poor environment, due to higher reactivity (e.g. Högdahl et al. 2012; Rubatto et al. 2013; Morrissey et al. 2016), potentially recording information completely missed by zircon. As in the first scenario (Fig. 10a), relict garnet breaks down during decompression to an intergrowth of plagioclase + orthopyroxene, another micro zircon permitting reaction (Fig. 10b). Local hydration or different amounts of melt loss (Morrissey et al. 2016; Larson et al. 2022) is reflected by a second P–T event (Fig. 10b, v, vi, vii) recorded only by more reactive high-strain microsites. The interpretation of geochronological data obtained from studies of high-grade metamorphic rocks can also be extremely challenging as most terranes record a spread of ages rather than statistical populations (e.g. Whitehouse and Kemp 2010; Farias et al. 2020; Taylor et al. 2020; Finch et al. 2021; Gutieva et al. 2021; Mulder and Cawood 2021; Salminen et al. 2022; Whitehouse et al. 2022). For example, in Figure 10c, the younger age population could represent either a single prolonged metamorphic event or potentially two events, with the second modifying the first and many other possible combinations. Determining the correct age interpretation depends on mineral textures, where the analyses and sample come together with complementary geochemical and isotopic information.

Metamorphic reactions between REE-rich phases such as allanite–monazite–xenotime–apatite are illustrated with those of titanite–rutile on the same theoretical P–T–t–(d) path, though these sets of minerals may occur in different chemical systems (Fig. 11). The timing and sequence of metamorphic reactions (i.e. growth of accessory phases) depend not only on the variation of pressure and temperature conditions, including the residence time of a volume of rock at certain conditions, but also on the bulk-rock composition of the protolith and on strain rate, which strongly influence the reaction sequence during fluid–rock interaction at different crustal levels. Reconstructing the formation and breakdown of these minerals and growth of other major and minor phases in the rock usually requires a detailed petrographic investigation that includes X-ray compositional maps and analysis of Y and REE partitioning between the phases (e.g. Janots et al. 2008; Garber et al. 2017; Manzotti et al. 2018; Airaghi et al. 2019). Additionally, tools such as TE mapping by LA-ICP-MS (e.g. Raimondo et al. 2017; Chew et al. 2021; Sliwinski and Stoll 2021; Godet et al. 2022) allow for greater precision in determining chemical composition effects on growth or break-down of minerals associated with different P–T conditions and fluid-present or absent processes.

Fig. 11.

Synoptic P–T evolution diagram including reactions of key petrochronometers such as garnet, allanite, monazite, rutile, apatite, xenotime and titanite. Source: modified after Manzotti et al. (2018) including key mineral reactions from Boston et al. (2017); Fumes et al. (2022); Manzotti et al. (2022).

Fig. 11.

Synoptic P–T evolution diagram including reactions of key petrochronometers such as garnet, allanite, monazite, rutile, apatite, xenotime and titanite. Source: modified after Manzotti et al. (2018) including key mineral reactions from Boston et al. (2017); Fumes et al. (2022); Manzotti et al. (2022).

During the prograde evolution of a time capsule mineral such as garnet, key inclusions can be incorporated and provide insights regarding the P and T conditions a rock volume experienced during the early stages of the evolution path (Fig. 11). For example, identification of rutile inclusions in garnet core (Fig. 11) and prograde allanite can indicate that rutile grew during the prograde stages towards max P conditions (Boston et al. 2017). In this case, prograde growth of garnet sequestered most of the Mn and HREE, resulting in allanite cores with a relatively HREE-depleted pattern in relation to the rim. Manzotti et al. (2018) investigated rocks in which Y-rich garnet cores enclosed a first generation of rutile and allanite inclusions, whereas Y-poor garnet rims included monazite grains that crystallized aligned parallel to the matrix foliation (Fig. 11). At this stage, monazite may form after breakdown of allanite or other phosphates such as apatite and/or xenotime (Engi 2017; Manzotti et al. 2018), recording peak pressure conditions (e.g. Boston et al. 2017; Manzotti et al. 2018; Volante et al. 2020c; Barrote et al. 2022b; Fumes et al. 2022). Monazite REE patterns may exhibit low Y content together with a steeply negative HREE slope, suggesting that garnet was still stable during monazite growth, whereas pronounced negative Eu anomalies commonly suggest plagioclase stability or oxygen fugacity variations (Fig. 11; e.g. Boston et al. 2017; Holder et al. 2020). At low- to medium-metamorphic grade, allanite and monazite may replace each other, and this process depends on different factors including bulk-rock (e.g. Wing et al. 2003; Janots et al. 2008; Spear 2010) and fluid (Budzyń et al. 2010) compositions, as well as oxygen fugacity (Janots et al. 2011) and pressure and temperature conditions (Janots et al. 2007). Post-metamorphic pressure peak, along the lower-pressure, possibly higher-temperature path, a second generation of allanite and apatite can form at the expenses of monazite grains commonly exhibiting complex coronitic dissolution textures (Fig. 11; e.g. Manzotti et al. 2018). At this stage, in mafic and Ca-rich systems, titanite crystals may grow as a replacement product of rutile, whereas ilmenite would grow in a Ca-poor system and at lower pressures. During retrograde, low-grade hydrothermal metamorphism in pelitic systems, new xenotime and monazite may crystallize following allanite, apatite, residual monazite relicts and garnet breakdown, with xenotime associated with chlorite and minor biotite aggregates replacing garnet porphyroblasts (Fig. 11; Manzotti et al. 2018). Most of the Y released by garnet would be incorporated by xenotime and monazite in minor amounts. New rutile crystals may grow during the retrograde cooling stage at the expense of titanite and as rutile exsolution, growing parallel to biotite cleavage during its replacement by chlorite (Fig. 11). During this late hydrothermal stage, rutile may grow with apatite and record the cooling stages (e.g. Apen et al. 2020).

In this section, we emphasize that a combined multi-mineral approach is a powerful tool to provide constraints on magma sources (e.g. mineralogy, composition), differentiation processes (e.g. mixing, fractional crystallization) and timing of magma formation and crystallization. The first scenario corresponds to simple fractional crystallization of a single magma batch (Fig. 12). The second is a more complex scenario wherein magma is contaminated by a distinct lithology and/or mingles with a different magma. The third scenario refers to Precambrian gneisses, wherein the objective is to recover the initial magmatic signatures despite the occurrence of a metamorphic overprinting event.

Fig. 12.

Hypothetical temperature–time paths for two distinct intrusions with one having cooled rapidly and the other one slowly. Note that closure temperatures do not overlap between techniques and minerals for illustrative purposes only. Also, cooling paths are presented as straight lines which may not accurately depict actual cooling paths.

Fig. 12.

Hypothetical temperature–time paths for two distinct intrusions with one having cooled rapidly and the other one slowly. Note that closure temperatures do not overlap between techniques and minerals for illustrative purposes only. Also, cooling paths are presented as straight lines which may not accurately depict actual cooling paths.

To constrain the crystallization age of a magmatic rock, zircon is arguably the easiest mineral to date among those presented here but it is not necessarily observed in all rock types. Consequently, the timing of igneous rock formation can be addressed using other phases such as titanite, rutile, monazite, apatite or allanite, provided that they record a magmatic age or fast cooling. In a scenario where a magmatic intrusion does not experience post-crystallization disturbance and contains most mineral phases presented in this contribution, multi-mineral dating becomes useful as these phases likely record different ages reflecting the magma cooling history (Fig. 12; e.g. Schaltegger et al. 2009; Jayananda et al. 2015). A potential issue is the precision on determined ages, as granite batholiths can cool over relatively short timescales (e.g. Coleman et al. 2004; Schaltegger et al. 2009; Barboni et al. 2013). In the same scenario, LuHf in zircon/apatite, SmNd in monazite/titanite/apatite and RbSr in apatite, combined, allow these source tracing radiogenic isotope systems to be directly compared, and coupled to geochronological data (i.e. UThPb and LuHf in apatite, and SmNd in monazite/titanite/apatite). Also, comparisons of these isotopic systems at the whole-rock and mineral scale can help constrain source composition or detect small discrepancies that may be important for understanding rock formation (e.g. Guitreau et al. 2012; Fisher et al. 2020; Barrote et al. 2022a; Volante et al. 2022; Zhang et al. 2022).

The second scenario includes a magmatic body that experienced mingling, mixing or contamination. Radiogenic systems allow end-members to be identified due to specific sensitivity to contamination, source characteristics and respective elemental concentrations. For example, Laurent et al. (2017) used RbSr and SmNd in apatite and titanite, respectively, from hybrid granitoids to show that these minerals did not crystallize at the same time and that while apatite formed in both magmas before mingling and after mingling (zonation), titanite crystallized only in the mingled magma. To identify end-member isotopic compositions, another geochemical study investigated zircon, apatite and titanite, which revealed mingling of mantle and crust-derived magmas (Sun et al. 2010). If one of the end-members is crustal derived, zircon xenocrysts can provide further information about crustal diversity (e.g. Villaros et al. 2012; Bea et al. 2021). Also, monazite can inherit Nd isotope heterogeneities from the source while losing the original UPb crystallization age (Fisher et al. 2017). Trace element patterns and ratios can provide further insights into magma evolution after and/or before mixing/mingling (e.g. Sun et al. 2010; Laurent et al. 2017). Melt and/or mineral inclusions in accessory minerals can also provide information about the melt source composition (e.g. Bruand et al. 2017; Antoine et al. 2020; Ferrero et al. 2021).

The third scenario considers Precambrian rocks, which often experienced polyphase histories that involved at least one metamorphic episode potentially compromising one or more isotope systems (e.g. Black 1988; Hammerli et al. 2014; Emo et al. 2018; Guitreau et al. 2018; Antoine et al. 2020; Fisher et al. 2020; Hammerli and Kemp 2021). For example, in situ SmNd investigation of apatite, allanite, titanite, xenotime and monazite in Precambrian metasedimentary rocks revealed re-equilibration of heterogeneous Nd isotope signature in apatite over 550°C, after which it retains its Nd isotope signature throughout anatexis (Hammerli et al. 2014). While titanite and allanite equilibrate at LT (350–400°C), REE-rich accessory phases exhibit homogeneous Nd isotopic signatures at HT (600°C) and behave as open systems during partial melting and anatexis (Hammerli et al. 2014).

From the conceptualization of petrochronology (as summarized in Engi et al. 2017; Kohn et al. 2017), significant analytical improvements have occurred in the field of geoscience and have been applied to all key mineral phases that are described in this review. New analytical methods have been developed (e.g. atom probe dating, Fougerouse et al. 2021; in situ Lu–Hf dating using LA-ICP-MS/MS, Simpson et al. 2021, 2022, 2023), previous analytical methods have been continuously updated and improved (e.g. in situ EPMA dating, Montel et al. 2018; in situ U–Pb dating, Millonig et al. 2020), new and updated numerical tools for quantitative petrology have been developed and enhanced, including imaging and modelling tools (e.g. XMapTools, Lanari et al. 2014, 2019; Q–XRMA, Ortolano et al. 2018, Zucali et al. 2021; Bingo-Antidote, Duesterhoeft and Lanari 2020) as well as the use of TE maps (e.g. Raimondo et al. 2017; George et al. 2018; Lanari and Piccoli 2020; Rubatto et al. 2020; Gaidies et al. 2021), and new boundaries have been pushed to investigate melting mechanisms (e.g. Ferrero et al. 2021). In this section, we briefly discuss the exciting and bright future in front of the discipline of petrochronology that has only started to push analytical advances to the limit of conventional methods, allowing for a better understanding of plate- to nanoscale crustal processes.

Garnet is the ultimate petrochronometer (Baxter et al. 2017), with the most exciting advance being the development of in situ LuHf geochronology (Simpson et al. 2021). The use of inclusions as proxies for whole-rock isotopic composition also means that this method may be suitable for detrital garnet studies (Maneiro et al. 2019; Mark et al. 2022). Future work should focus on enhancing accuracy and precision of LuHf and U–Pb measurement in secondary reference material e.g. Aysal et al. 2023 and finding primary reference material (i.e. isotopically homogeneous) for common Hf-bearing phases such as garnet (Simpson et al. 2021). Additionally, incorporation of Mn in the activity–composition relations for amphibole and clinopyroxene in the thermodynamic dataset for mafic systems (Green et al. 2016) would also greatly enhance accuracy for garnet modelling (White et al. 2014). Recent advances using in situ UPb dating of garnet have also demonstrated significant potential (e.g. Seman et al. 2017; Millonig et al. 2020; Schannor et al. 2021) and applicability to almandine compositions (e.g. Cerva-Alves et al. 2021; Schannor et al. 2021). Further advances aim to extend the range of metasomatic and metamorphic garnet composition that can be targeted using this technique (Millonig et al. 2022; O'Sullivan et al. 2023). Finally, the continuous development of quantitative elemental and mineral map tools (e.g. XMapTools) along with EPMA and LA-ICP-MS trace element maps (e.g. Chew et al. 2021) will prove an extremely powerful tool to enhance in situ microanalytical investigations and our understanding of complex micro-chemical processes. The Bingo-Antidote add-on to XMapTools performs thermodynamic calculations by comparing modelled and observed mineral assemblages, modes and compositions (Duesterhoeft and Lanari 2020).

Zircon can preserve a realm of information that can be retrieved from images (optical, SEM), mineral inclusions and elemental and isotope data (e.g. Th/U, REE, δ18O, εHf, δ30Si). Although zircon has long been used in igneous and metamorphic studies, recent technical and methodological advances allow to further extend our use of this mineral. These include analytical improvements in δ30Si and δ94Zr data acquisition to enhance and extend the use of these isotopes in zircon studies, and the integration of EBSD data with petrochronological investigation to better understand TE diffusion in zircon crystals under various P–T conditions. Pioneer work on elastic geobarometry for a noncubic host-inclusion system indicates potential of anisotropic quartz-in-zircon elastic model for elastic thermobarometry, and its potential wide applicability to crustal rocks (Gonzalez et al. 2021).

Monazite is an increasingly useful accessory mineral as it can be used to trace the presence of certain fluids (e.g. Weinberg et al. 2020; Salminen et al. 2022), and can potentially grow throughout metamorphism (e.g. Larson et al. 2022). With improvement of analytical and spatial resolution, even if highly zoned, monazite will only become more applicable for unravelling complex and multifaceted processes. However, even with the collection of in situ data, it can still be challenging to tie monazite growth to deformation and fluid-related processes. Further work into monazite behaviour under deformation combined with the use of EBSD may prove a powerful tool. Additionally, the integration of EBSD, crystallographic vorticity axis analysis and petrochronology should be applied to monazite within deformational structures as it has been done for zircon (Brown et al. 2022). More work needs to be done on monazite TE partitioning to better understand temperature dependencies (e.g. Larson et al. 2022). Some attempts have been made to thermodynamically model monazite and xenotime (e.g. Kelsey et al. 2007, 2008; Spear and Pyle 2010; Shrestha et al. 2019), however, this approach only works with low-Ca bulk-rock compositions and requires the simplification of a complicated system. These current limitations represent an extensive platform for improvements in this field.

Titanite is a very reactive mineral compared to other accessory minerals; however, its tendency to incorporate initial Pb requires combining U–Pb dating with trace-element composition, zoning and microstructural information to accurately filter out any additional, unrelated radiogenic Pb (Walters et al. 2022). Analytical improvement in acquisition of Sm–Nd and O isotopic analysis in titanite may be extremely beneficial to trace magmatic sources (e.g. Bonamici et al. 2014; Bruand et al. 2019, 2020) and the origin of fluids playing a major role in crustal-scale shear zones (Gordon et al. 2021; Moser et al. 2022) and ore deposit (e.g. Marfin et al. 2020) genesis. The assessment of ilmenite U–Pb dating by LA-ICP-MS (Thompson et al. 2021) may be a valuable contribution for evaluating residual initial Pb affecting titanite dating. It may also be useful to investigate the timing of replacement of these Ti-rich phases. Advancing and developing the recent application of elastic geothermobarometry of host-inclusion systems (e.g. Mazzucchelli et al. 2021) to titanite grains may provide a complementary non-destructive method to estimate the P and T at which inclusions were trapped in the hosting titanite (Nestola 2021).

Rutile is a crucial mineral to unravel the early history of rock's evolution, including the prograde, UHP, burial stage (e.g. Rezvukhina et al. 2021; Manzotti et al. 2022), to trace melt sources in magmatic systems using Lu–Hf as isotopic tracer (e.g. Ewing et al. 2011) and hydrothermal, metasomatic processes associated with ore deposit genesis (e.g. Agangi et al. 2019; Ballouard et al. 2020; Porter et al. 2020; Sciuba and Beaudoin 2021). Therefore, analytical advances in microanalytical and petrochronological investigations of this mineral phase are essential to contribute to our understanding of orogenic systems, tectonic environments and exploration strategies. A better micro- to nanoscale understanding of trace element mobility in rutile (Kooijman et al. 2012; Ewing et al. 2013; Kohn et al. 2016; Pape et al. 2016; Penniston-Dorland et al. 2018; Smye et al. 2018) may improve U–Pb dating and associated trace element distribution and isotopic information (Verberne et al. 2022a, b), including the investigation of Zr transport through the rock's matrix (Ewing et al. 2013; Kohn et al. 2016). Also, advancing the use of integrated EBSD investigations of rutile grains within different rock-types with LA-ICP-MS trace element mapping may significantly enhance our understanding of trace element partitioning and concentration in this mineral.

Allanite has overcome various challenges as a petrochronometer. Indeed, the highly variable composition of natural allanite complicates petrological interpretations and applications. Thermobarometric, isotopic (e.g. in situ oxygen) and TE-based petrological investigations have to be carried out using microanalytical techniques such as EBSD and/or LA-ICP-MS TE maps to explore these compositional variations. The use of compositional maps to distinguish different generations of allanite has proven to be an essential step in geochronological and petrological studies (e.g. Burn 2016; Airaghi et al. 2019; Corti et al. 2020), and the use of TE maps obtained with LA-ICP-MS can improve classification and understanding of allanite growth history. The chemical distinction between magmatic and metamorphic allanite has been enhanced by machine learning (e.g. Di Rosa et al. 2020), but more data are required to improve this classification. Nd and Sr isotopes are valuable geochemical tools that have been utilized in previous studies. Recent analytical advancements and developments have further strengthened the importance of these isotopic systems as tracers in understanding magmatic and metamorphic processes, especially in investigations of hydrothermal processes. Also, a better comprehension of the stability conditions (pressure, temperature and fluid) of allanite and correlated accessory minerals (e.g. monazite) is required. This would involve experiments, improvements on the thermodynamic dataset and investigation of inclusions (mineral, fluid and melt) in allanite.

Apatite is a useful tool for studying crustal processes by means of a variety of petrochronology methods, continuously enhanced in their accuracy and precision. Glorie et al. (2022, this volume, in press) show the promising apatite Lu–Hf geochronology method using LA ICP-MS/MS, which is more likely to reveal primary apatite growth ages in reworked terranes, due to the higher Tc, than the U–Pb system (Fig. 1). This may be important since Odlum et al. (2022) show that apatite REEs and U–Pb behaviour are decoupled in high-grade gneiss samples, suggesting REEs record higher-temperature processes than U–Pb isotopic systems. LA-ICP-MS trace element mapping is a promising technique to monitor the behaviour of trace elements in accessory mineral phases, and it should be more routinely applied. In apatite, it can be used to distinguish between metamorphic and magmatic grains or identify growth zones (Henrichs et al. 2018, 2019; Chew et al. 2021). Recent techniques such as age-depth profiles and laser ablation split stream make it possible to identify individual apatite generations and can aid with interpreting data from apatite domains with complex thermal histories (Kirkland et al. 2018). New apatite reference material is being characterized for in situ apatite U–Pb petrochronology and Sr–Nd isotope geochemistry (e.g. Apen et al. 2022; Kennedy et al. 2022) to obtain more precise and robust data.

In addition to the key mineral phases discussed in this contribution, we encourage and see potential in integrating other mineral phases in magmatic and metamorphic studies. Multi-mineral studies are paramount to combine major and accessory mineral phases that record different steps of the prograde and retrograde history of a volume of rock with lower temperature petrochronometers, which may retain useful information on the youngest evolution of the system. Critical and crucial advancement in petrochronology in the near future includes in situ investigations, where microanalytical studies are linked to the chemical and chronological information of different minerals and associated fabrics in a rock volume. Extreme potential is foreseen in the development and acquisition of quantitative LA-ICP-MS TE (e.g. George et al. 2018; Muñoz-Montecinos, J. et al. 2023) and geochronological maps (e.g. Chew et al. 2021). In this scenario, by enabling near-simultaneous detection, the time of flight (TOF) detector is a key component of LA-ICP-TOF-MS, a promising technique that captures the complete elemental mass spectrum for each laser pulse (Chew et al. 2021). Generating quantitative chemical and geochronological maps of a whole mineral using LA-ICP-MS can facilitate exploring the relationships between internal textures and trace and major elemental variation over time. This analytical technique will allow to discriminate distinct textural domains based on the ages and major, minor and trace elements, possibly avoiding issues such as mixing domains to help the interpretation of complex mineral textures. The advances made in LA-ICP-MS imaging have made it a critical analytical technique in the geosciences, as it enables the acquisition of high-resolution geochemical data that can help to constrain the timing and nature of igneous, metamorphic, ore-forming, sedimentary and diagenetic processes.

We thank Catherine Mottram and an anonymous reviewer for their constructive and insightful comments that greatly improved the manuscript. We also thank Inês Pereira for helpful comments and for her editorial handling. Martin Engi and Daniela Rubatto are thanked for insightful discussion on allanite. Vitor Barrote is thanked for long insightful discussion throughout the preparation of the manuscript about the pros and cons of using different analytical techniques in petrochronology applied to the mineral phases discussed in this manuscript. M. Tedeschi was supported by the National Council for Scientific and Technological Development – CNPq through a Research Productivity Grant 308733/2021-5. This is contribution no. 591 of the ClerVolc program of the International Research Center for Disaster Sciences and Sustainable Development of the University of Clermont Auvergne.

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

SV: conceptualization (lead), formal analysis (lead), project administration (lead), writing – original draft (lead), writing – review & editing (lead); EB: conceptualization (equal), formal analysis (equal), writing – original draft (equal), writing – review & editing (equal); MG: conceptualization (equal), formal analysis (equal), writing – original draft (equal), writing – review & editing (equal); MT: conceptualization (supporting), formal analysis (supporting), writing – original draft (supporting), writing – review & editing (supporting); VVS: conceptualization (supporting), data curation (supporting), writing – original draft (supporting), writing – review & editing (supporting); KC: conceptualization (supporting), data curation (supporting), writing – original draft (supporting), writing – review & editing (supporting).

This research received no specific grant from any funding agency in the public, commercial or not-for-profit sectors.

Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.

1.
Adlakha
E.
and
Hattori
K.
2021
.
Thermotectonic events recorded by U–Pb geochronology and Zr-in-rutile thermometry of Ti oxides in basement rocks along the P2 fault, eastern Athabasca Basin, Saskatchewan, Canada
.
GSA Bulletin
 ,
134
,
567
576
,
2.
Aerden
D.G.A.M.
,
Bell
T.H.
,
Puga
E.
,
Sayab
M.
,
Lozano
J.A.
and
Diaz de Federico
A.
2013
.
Multi-stage mountain building v. relative plate motions in the Betic Cordillera deduced from integrated microstructural and petrological analysis of porphyroblast inclusion trails
.
Tectonophysics
 ,
587
,
188
206
,
3.
Aerden
D.G.A.M.
,
Ruiz-Fuentes
A.
,
Sayab
M.
and
Forde
A.
2021
.
Kinematics of subduction in the Ibero-Armorican arc constrained by 3D microstructural analysis of garnet and pseudomorphed lawsonite porphyroblasts from Île de Groix (Variscan belt)
.
Solid Earth
 ,
12
,
971
992
,
4.
Agangi
A.
,
Reddy
S.M.
,
Plavsa
D.
,
Fougerouse
D.
,
Clark
C.
,
Roberts
M.
and
Johnson
T.E.
2019
.
Antimony in rutile as a pathfinder for orogenic gold deposits
.
Ore Geology Reviews
 ,
106
,
1
11
,
5.
Airaghi
L.
,
Janots
E.
,
Lanari
P.
,
de Sigoyer
J.
and
Magnin
V.
2019
.
Allanite petrochronology in fresh and retrogressed garnet–biotite metapelites from the Longmen Shan (Eastern Tibet)
.
Journal of Petrology
 ,
60
,
151
176
,
6.
Ambrose
T.K.
,
Larson
K.P.
,
Guilmette
C.
,
Cottle
J.M.
,
Buckingham
H.
and
Rai
S.
2015
.
Lateral extrusion, underplating, and out-of-sequence thrusting within the Himalayan metamorphic core, Kanchenjunga, Nepal
.
Lithosphere
 ,
7
,
441
464
,
7.
Amelin
Y.
2009
.
Sm–Nd and U–Pb systematics of single titanite grains
.
Chemical Geology
 ,
261
,
53
61
,
8.
Amelin
Y.
,
Lee
D.-C.
and
Halliday
A.N.
2000
.
Early–middle Archaean crustal evolution deduced from Lu–Hf and U–Pb isotopic studies of single zircon grains
.
Geochimica et Cosmochimica Acta
 ,
64
,
4205
4225
,
9.
An
Y.
,
Huang
J.-X.
,
Griffin
W.L.
,
Liu
C.
and
Huang
F.
2017
.
Isotopic composition of Mg and Fe in garnet peridotites from the Kaapvaal and Siberian cratons
.
Geochimica et Cosmochimica Acta
 ,
200
,
167
185
,
10.
Anderson
A.J.
,
Hanchar
J.M.
,
Hodges
K.V.
and
van Soest
M.C.
2020
.
Mapping radiation damage zoning in zircon using Raman spectroscopy: implications for zircon chronology
.
Chemical Geology
 ,
538
, article
119494
,
11.
Andersson
J.
,
Möller
C.
and
Johansson
L.
2002
.
Zircon geochronology of migmatite gneisses along the Mylonite Zone (S Sweden): a major Sveconorwegian terrane boundary in the Baltic Shield
.
Precambrian Research
 ,
114
,
121
147
,
12.
Andersson
J.
,
Bingen
B.
,
Cornell
D.
,
Johansson
L.
,
Söderlund
U.
and
Möller
C.
2008
.
The Sveconorwegian orogen of southern Scandinavia: setting, petrology and geochronology of polymetamorphic high-grade terranes
.
33rd International Geological Congress
,
6–14 August 2008
,
Oslo, Norway
,
Field Excursion Guide
.
13.
Andersson
J.B.H.
,
Logan
L.
et al
2022
.
U–Pb zircon–titanite–apatite age constraints on basin development and basin inversion in the Kiruna mining district, Sweden
.
Precambrian Research
 ,
372
, article
106613
,
14.
Andrehs
G.
and
Heinrich
W.
1998
.
Experimental determination of REE distributions between monazite and xenotime: potential for temperature-calibrated geochronology
.
Chemical Geology
 ,
149
,
83
96
,
15.
Anenburg
M.
,
Broom-Fendley
S.
and
Chen
W.
2021
.
Formation of rare earth deposits in carbonatites
.
Elements
 ,
17
,
327
332
.
16.
Antoine
C.
,
Bruand
E.
,
Guitreau
M.
and
Devidal
J.
2020
.
Understanding preservation of primary signatures in apatite by comparing matrix and zircon-hosted crystals from the Eoarchean Acasta Gneiss Complex (Canada)
.
Geochemistry, Geophysics, Geosystems
 ,
21
, article
e2020GC008923
,
17.
Apen
F.E.
,
Rudnick
R.L.
,
Cottle
J.M.
,
Kylander-Clark
A.R.C.
,
Blondes
M.S.
,
Piccoli
P.M.
and
Seward
G.
2020
.
Four-dimensional thermal evolution of the East African Orogen: accessory phase petrochronology of crustal profiles through the Tanzanian Craton and Mozambique Belt, northeastern Tanzania
.
Contributions to Mineralogy and Petrology
 ,
175
, article
97
,
18.
Apen
F.E.
,
Wall
C.J.
,
Cottle
J.M.
,
Schmitz
M.D.
,
Kylander-Clark
A.R.
and
Seward
G.G.
2022
.
Apatites for destruction: reference apatites from Morocco and Brazil for U–Pb petrochronology and Nd and Sr isotope geochemistry
.
Chemical Geology
 ,
590
, article
120689
,
19.
Aranovich
L.Y.
,
Bortnikov
N.
et al
2017
.
Morphology and impurity elements of zircon in the oceanic lithosphere at the Mid-Atlantic ridge axial zone (6–13 N): evidence of specifics of magmatic crystallization and postmagmatic transformations
.
Petrology
 ,
25
,
339
364
,
20.
Armbruster
T.
,
Bonazzi
P.
et al
2006
.
Recommended nomenclature of epidote-group minerals
.
European Journal of Mineralogy
 ,
18
,
551
567
,
21.
Asami
M.
,
Suzuki
K.
and
Grew
E.S.
2002
.
Chemical Th–U–total Pb dating by electron microprobe analysis of monazite, xenotime and zircon from the Archean Napier Complex, East Antarctica: evidence for ultra-high-temperature metamorphism at 2400 Ma
.
Precambrian Research
 ,
114
,
249
275
,
22.
Atherton
M.P.
1968
.
The variation in garnet, biotite and chlorite composition in medium grade pelitic rocks from the Dalradian, Scotland, with particular reference to the zonation in garnet
.
Contributions to Mineralogy and Petrology
 ,
18
,
347
371
,
23.
Aulbach
S.
,
O'Reilly
S.Y.
,
Griffin
W.
and
Pearson
N.J.
2008
.
Subcontinental lithospheric mantle origin of high niobium/tantalum ratios in eclogites
.
Nature Geoscience
 ,
1
,
468
472
,
24.
Axelsson
E.
,
Pape
J.
,
Berndt
J.
,
Corfu
F.
,
Mezger
K.
and
Raith
M.M.
2018
.
Rutile R632 – a new natural reference material for U–Pb and Zr determination
.
Geostandards and Geoanalytical Research
 ,
42
,
319
338
,
25.
Aysal
N.
,
Guillong
M.
et al
2023
.
A New Natural Secondary Reference Material for Garnet U-Pb Dating by TIMS and LA-ICP-MS
.
Geostandards and Geoanalytical Rsearch
 ,
26.
Balan
E.
,
Neuville
D.R.
,
Trocellier
P.
,
Fritsch
E.
,
Muller
J.-P.
and
Calas
G.
2001
.
Metamictization and chemical durability of detrital zircon
.
American Mineralogist
 ,
86
,
1025
1033
,
27.
Baldwin
J.A.
and
Brown
M.
2008
.
Age and duration of ultrahigh-temperature metamorphism in the Anápolis–Itauçu Complex, Southern Brasília Belt, central Brazil – constraints from U–Pb geochronology, mineral rare earth element chemistry and trace-element thermometry
.
Journal of Metamorphic Geology
 ,
26
,
213
233
,
28.
Ballouard
C.
,
Massuyeau
M.
,
Elburg
M.A.
,
Tappe
S.
,
Viljoen
F.
and
Brandenburg
J.-T.
2020
.
The magmatic and magmatic-hydrothermal evolution of felsic igneous rocks as seen through Nb–Ta geochemical fractionation, with implications for the origins of rare-metal mineralizations
.
Earth-Science Reviews
 ,
203
, article
103115
,
29.
Barboni
M.
,
Schoene
B.
,
Ovtcharova
M.
,
Bussy
F.
,
Schaltegger
U.
and
Gerdes
A.
2013
.
Timing of incremental pluton construction and magmatic activity in a back-arc setting revealed by ID-TIMS U/Pb and Hf isotopes on complex zircon grains
.
Chemical Geology
 ,
342
,
76
93
,
30.
Barfod
G.H.
,
Otero
O.
and
Albarède
F.
2003
.
Phosphate Lu–Hf geochronology
.
Chemical Geology
 ,
200
,
241
253
,
31.
Barfod
G.H.
,
Krogstad
E.J.
,
Frei
R.
and
Albarède
F.
2005
.
Lu–Hf and PbSL geochronology of apatites from Proterozoic terranes: a first look at Lu–Hf isotopic closure in metamorphic apatite
.
Geochimica et Cosmochimica Acta
 ,
69
,
1847
1859
,
32.
Barla
A.
2021
.
Titanite petrochronology by LA-ICPMS: method and significance in deciphering igneous processes
,
Msc Thesis
,
University of Arizona
,
USA
.
33.
Barrote
V.R.
,
McNaughton
N.J.
,
Tessalina
S.G.
,
Evans
N.J.
,
Talavera
C.
,
Zi
J.-W.
and
McDonald
B.J.
2020
.
The 4D evolution of the Teutonic Bore Camp VHMS deposits, Yilgarn Craton, Western Australia
.
Ore Geology Reviews
 ,
120
, article
103448
,
34.
Barrote
V.R.
,
Nebel
O.
,
Wainwright
A.N.
,
Cawood
P.A.
,
Hollis
S.P.
and
Raveggi
M.
2022a
.
Testing the advantages of simultaneous in-situ SmNd, UPb and elemental analysis of igneous monazite for petrochronological studies. An example from the late Archean, Penzance granite, Western Australia
.
Chemical Geology
 ,
594
, article
120760
,
35.
Barrote
V.R.
,
Volante
S.
,
Blereau
E.R.
,
Rosière
C.A.
and
Spencer
C.J.
2022b
.
Implications of the dominant LP–HT deformation in the Guanhães Block for the Araçuaí West-Congo Orogen evolution
.
Gondwana Research
 ,
107
,
154
175
,
36.
Barrow
G.
1893
.
On an intrusion of muscovite–biotite gneiss in the South-eastern Highlands of Scotland, and its accompanying metamorphism
.
Quarterly Journal of the Geological Society
 ,
49
,
330
358
,
37.
Barrow
G.
1912
.
On the geology of lower Dee-side and the southern Highland border
.
Proceedings of the Geologists’ Association
 ,
23
,
274–290
,
38.
Barth
S.
,
Oberli
F.
and
Meier
M.
1994
.
ThPb v. UPb isotope systematics in allanite from co-genetic rhyolite and granodiorite: implications for geochronology
.
Earth and Planetary Science Letters
 ,
124
,
149
159
,
39.
Bartoli
O.
,
Cesare
B.
,
Poli
S.
,
Bodnar
R.J.
,
Acosta-Vigil
A.
,
Frezzotti
M.L.
and
Meli
S.
2013
.
Recovering the composition of melt and the fluid regime at the onset of crustal anatexis and S-type granite formation
.
Geology
 ,
41
,
115
118
,
40.
Baxter
E.F.
and
Scherer
E.E.
2013
.
Garnet geochronology: timekeeper of tectonometamorphic processes
.
Elements
 ,
9
,
433
438
,
41.
Baxter
E.F.
,
Caddick
M.J.
and
Dragovic
B.
2017
.
Garnet: a rock-forming mineral petrochronometer
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
469
533
,
42.
Bea
F.
1996
.
Residence of REE, Y, Th and U in granites and crustal protoliths; implications for the chemistry of crustal melts
.
Journal of Petrology
 ,
37
,
521
552
,
43.
Bea
F.
and
Montero
P.
1999
.
Behavior of accessory phases and redistribution of Zr, REE, Y, Th, and U during metamorphism and partial melting of metapelites in the lower crust: an example from the Kinzigite Formation of Ivrea–Verbano, NW Italy
.
Geochimica et Cosmochimica Acta
 ,
63
,
1133
1153
,
44.
Bea
F.
,
Pereira
M.D.
and
Stroh
A.
1994
.
Mineral/leucosome trace-element partitioning in a peraluminous migmatite (a laser ablation-ICP-MS study)
.
Chemical Geology
 ,
117
,
291
312
,
45.
Bea
F.
,
Morales
I.
,
Molina
J.F.
,
Montero
P.
and
Cambeses
A.
2021
.
Zircon stability grids in crustal partial melts: implications for zircon inheritance
.
Contributions to Mineralogy and Petrology
 ,
176
,
1
13
,
46.
Bea
F.
,
Bortnikov
N.
et al
2022
.
Zircon crystallization in low-Zr mafic magmas: possible or impossible?
Chemical Geology
 ,
602
, article
120898
,
47.
Bebout
G.E.
,
Tsujimori
T.
,
Ota
T.
,
Shimaki
Y.
,
Kunihiro
T.
,
Carlson
W.D.
and
Nakamura
E.
2014
.
Lithium behavior during growth of metasedimentary garnets from the Cignana UHP locality, Italy
.
American Geophysical Union Fall Meeting
,
15–19 December 2014
,
San Francisco, CA
, Abstracts,
V31D-4783
.
48.
Bebout
G.E.
,
Ota
T.
,
Kunihiro
T.
,
Carlson
W.D.
and
Nakamura
E.
2022
.
Lithium in garnet as a tracer of subduction zone metamorphic reactions: the record in ultrahigh-pressure metapelites at Lago di Cignana, Italy
.
Geosphere
 ,
18
,
1020
1029
,
49.
Bell
E.A.
,
Boehnke
P.
,
Hopkins-Wielicki
M.D.
and
Harrison
T.M.
2015
.
Distinguishing primary and secondary inclusion assemblages in Jack Hills zircons
.
Lithos
 ,
234
,
15
26
,
50.
Belousova
E.A.
,
Walters
S.
,
Griffin
W.L.
and
O'Reilly
S.Y.
2001
.
Trace-element signatures of apatites in granitoids from the Mt Isa Inlier, northwestern Queensland
.
Australian Journal of Earth Sciences
 ,
48
,
603
619
,
51.
Belousova
E.
,
Griffin
W.L.
,
O'Reilly
S.Y.
and
Fisher
N.
2002
.
Igneous zircon: trace element composition as an indicator of source rock type
.
Contributions to Mineralogy and Petrology
 ,
143
,
602
622
,
52.
Berman
R.G.
1988
.
Internally-consistent thermodynamic data for minerals in the system Na2O–K2O–CaO–MgO–FeO–Fe2O3–Al2O3–SiO2–TiO2–H2O–CO2
.
Journal of Petrology
 ,
29
,
445
522
,
53.
Bindeman
I.
2008
.
Oxygen isotopes in mantle and crustal magmas as revealed by single crystal analysis
.
Reviews in Mineralogy and Geochemistry
 ,
69
,
445
478
,
54.
Bingen
B.
,
Demaiffe
D.
and
Hertogen
J.
1996
.
Redistribution of rare earth elements, thorium, and uranium over accessory minerals in the course of amphibolite to granulite facies metamorphism: the role of apatite and monazite in orthogneisses from southwestern Norway
.
Geochimica et Cosmochimica Acta
 ,
60
,
1341
1354
,
55.
Black
L.
P.
1988
.
Isotopic resetting of U–Pb zircon and Rb–Sr and Sm–Nd whole-rock systems in Enderby Land, Antarctica: implications for the interpretation of isotopic data from polymetamorphic and multiply deformed terrains
.
Precambrian Research
 ,
38
,
355
365
,
56.
Black
L.P.
,
Fitzgerald
J.D.
and
Harley
S.L.
1984
.
Pb isotopic composition, colour, and microstructure of monazites from a polymetamorphic rock in Antarctica
.
Contributions to Mineralogy and Petrology
 ,
85
,
141
148
,
57.
Blereau
E.
,
Clark
C.
,
Taylor
R.J.
,
Johnson
T.
,
Fitzsimons
I.
and
Santosh
M.
2016
.
Constraints on the timing and conditions of high-grade metamorphism, charnockite formation and fluid–rock interaction in the Trivandrum Block, southern India
.
Journal of Metamorphic Geology
 ,
34
,
527
549
,
58.
Blereau
E.
2017
.
A Petrochronological Investigation of Metamorphic, Melt and Fluid Related Processes in Lower Crustal Rocks from Southwestern Norway and Southern India
 . PhD Thesis,
Curtin University
,
Western Australia
.
59.
Blereau
E.
,
Clark
C.
,
Kinny
P.D.
,
Sansom
E.
,
Taylor
R.J.M.
and
Hand
M.
2022
.
Probing the history of ultra-high temperature metamorphism through rare earth element diffusion in zircon
.
Journal of Metamorphic Geology
 ,
40
,
329
357
,
60.
Blichert-Toft
J.
and
Albarède
F.
1997
.
The Lu–Hf isotope geochemistry of chondrites and the evolution of the mantle–crust system
.
Earth and Planetary Science Letters
 ,
148
,
243
258
,
61.
Bloch
E.
,
Jollands
M.
et al
2022
.
Diffusion anisotropy of Ti in zircon and implications for Ti-in-zircon thermometry
.
Earth and Planetary Science Letters
 ,
578
, article
117317
,
62.
Boehnke
P.
,
Watson
E.B.
,
Trail
D.
,
Harrison
T.M.
and
Schmitt
A.K.
2013
.
Zircon saturation re-revisited
.
Chemical Geology
 ,
351
,
324
334
,
63.
Böhnke
M.
,
Bröcker
M.
,
Maulana
A.
,
Klemd
R.
,
Berndt
J.
and
Baier
H.
2019
.
Geochronology and Zr-in-rutile thermometry of high-pressure/low temperature metamorphic rocks from the Bantimala complex, SW Sulawesi, Indonesia
.
Lithos
 ,
324–325
,
340
355
,
64.
Bollen
E.M.
,
Stowell
H.H.
,
Aronoff
R.F.
,
Stotter
S.V.
,
Daniel
C.G.
,
McFarlane
C.R.M.
and
Vervoort
J.D.
2022
.
Reconciling garnet Lu–Hf and Sm–Nd and monazite U–Pb ages for a prolonged metamorphic event, northern New Mexico
.
Journal of Petrology
 ,
63
, article
egac031
,
65.
Bonamici
C.E.
and
Blum
T.B.
2020
.
Reconsidering initial Pb in titanite in the context of in situ dating
.
American Mineralogist: Journal of Earth and Planetary Materials
 ,
105
,
1672
1685
,
66.
Bonamici
C.E.
,
Kozdon
R.
,
Ushikubo
T.
and
Valley
J.W.
2011
.
High-resolution P–T–t paths from δ18O zoning in titanite: a snapshot of late-orogenic collapse in the Grenville of New York
.
Geology
 ,
39
,
959
962
,
67.
Bonamici
C.E.
,
Kozdon
R.
,
Ushikubo
T.
and
Valley
J.W.
2014
.
Intragrain oxygen isotope zoning in titanite by SIMS: cooling rates and fluid infiltration along the Carthage-Colton Mylonite Zone, Adirondack Mountains, NY, USA
.
Journal of Metamorphic Geology
 ,
32
,
71
92
,
68.
Bonamici
C.E.
,
Fanning
C.M.
,
Kozdon
R.
,
Fournelle
J.H.
and
Valley
J.W.
2015
.
Combined oxygen-isotope and U–Pb zoning studies of titanite: new criteria for age preservation
.
Chemical Geology
 ,
398
,
70
84
,
69.
Bonnet
G.
,
Chopin
C.
,
Locatelli
M.
,
Kylander-Clark
A.R.C.
and
Hacker
B.R.
2022
.
Protracted subduction of the European hyperextended margin revealed by rutile U–Pb geochronology across the Dora-Maira Massif (Western Alps)
.
Tectonics
 ,
41
, article
e2021TC007170
,
70.
Borghini
A.
,
Ferrero
S.
,
O'Brien
P.J.
,
Laurent
O.
,
Günter
C.
and
Ziemann
M.A.
2020
.
Cryptic metasomatic agent measured in situ in Variscan mantle rocks: melt inclusions in garnet of eclogite, Granulitgebirge, Germany
.
Journal of Metamorphic Geology
 ,
38
,
207
234
,
71.
Bosse
V.
and
Villa
I.M.
2019
.
Petrochronology and hygrochronology of tectono-metamorphic events
.
Gondwana Research
 ,
71
,
76
90
,
72.
Boston
K.R.
,
Rubatto
D.
,
Hermann
J.
,
Engi
M.
and
Amelin
Y.
2017
.
Geochronology of accessory allanite and monazite in the Barrovian metamorphic sequence of the Central Alps, Switzerland
.
Lithos
 ,
286
,
502
518
,
73.
Bouvier
A.
,
Vervoort
J.D.
and
Patchett
P.J.
2008
.
The Lu–Hf and Sm–Nd isotopic composition of CHUR: constraints from unequilibrated chondrites and implications for the bulk composition of terrestrial planets
.
Earth and Planetary Science Letters
 ,
273
,
48
57
,
74.
Bouvier
A.-S.
,
Ushikubo
T.
,
Kita
N.T.
,
Cavosie
A.J.
,
Kozdon
R.
and
Valley
J.W.
2012
.
Li isotopes and trace elements as a petrogenetic tracer in zircon: insights from Archean TTGs and sanukitoids
.
Contributions to Mineralogy and Petrology
 ,
163
,
745
768
,
75.
Braun
I.
,
Montel
J.-M.
and
Nicollet
C.
1998
.
Electron microprobe dating of monazites from high-grade gneisses and pegmatites of the Kerala Khondalite Belt, southern India
.
Chemical Geology
 ,
146
,
65
85
,
76.
Brown
D.A.
,
Simpson
A.
,
Hand
M.
,
Morrissey
L.J.
,
Gilbert
S.
,
Tamblyn
R.
and
Glorie
S.
2022
.
Laser-ablation Lu–Hf dating reveals Laurentian garnet in subducted rocks from southern Australia
.
Geology
 ,
50
,
837
842
,
77.
Bruand
E.
,
Storey
C.
and
Fowler
M.
2014
.
Accessory mineral chemistry of high Ba–Sr granites from northern Scotland: constraints on petrogenesis and records of whole-rock signature
.
Journal of Petrology
 ,
55
,
1619
1651
,
78.
Bruand
E.
,
Fowler
M.
,
Storey
C.
and
Darling
J.
2017
.
Apatite trace element and isotope applications to petrogenesis and provenance
.
American Mineralogist
 ,
102
,
75
84
,
79.
Bruand
E.
,
Storey
C.
,
Fowler
M.
and
Heilimo
E.
2019
.
Oxygen isotopes in titanite and apatite, and their potential for crustal evolution research
.
Geochimica et Cosmochimica Acta
 ,
255
,
144
162
,
80.
Bruand
E.
,
Fowler
M.
et al
2020
.
Accessory mineral constraints on crustal evolution: elemental fingerprints for magma discrimination
.
Geochemical Perspectives Letters
 ,
13
,
7
12
,
81.
Budzyń
B.
,
Hetherington
C.J.
,
Williams
M.L.
,
Jercinovic
M.J.
and
Michalik
M.
2010
.
Fluid–mineral interactions and constraints on monazite alteration during metamorphism
.
Mineralogical Magazine
 ,
74
,
659
681
,
82.
Buick
I.S.
,
Hermann
J.
,
Williams
I.S.
,
Gibson
R.L.
and
Rubatto
D.
2006
.
A SHRIMP U–Pb and LA-ICP-MS trace element study of the petrogenesis of garnet–cordierite–orthoamphibole gneisses from the Central Zone of the Limpopo Belt, South Africa
.
Lithos
 ,
88
,
150
172
,
83.
Buick
I.S.
,
Clark
C.
,
Rubatto
D.
,
Hermann
J.
,
Pandit
M.
and
Hand
M.
2010
.
Constraints on the Proterozoic evolution of the Aravalli–Delhi Orogenic Belt (NW India) from monazite geochronology and mineral trace element geochemistry
.
Lithos
 ,
120
,
511
528
,
84.
Burn
M.
2016
.
LA-ICP-QMS Th–U/Pb allanite dating: methods and applications
, Doctoral dissertation,
Philosophisch-naturwissenschfatliche Fakultät der Universität Bern
,
Germany
.
85.
Burn
M.
,
Lanari
P.
,
Pettke
T.
and
Engi
M.
2017
.
Non-matrix-matched standardisation in LA-ICP-MS analysis: general approach, and application to allanite Th–U–Pb dating
.
Journal of Analytical Atomic Spectrometry
 ,
32
,
1359
1377
,
86.
Caddick
M.J.
and
Thompson
A.B.
2008
.
Quantifying the tectono-metamorphic evolution of pelitic rocks from a wide range of tectonic settings: mineral compositions in equilibrium
.
Contributions to Mineralogy and Petrology
 ,
156
,
177
195
,
87.
Caddick
M.J.
,
Konopásek
J.
and
Thompson
A.B.
2010
.
Preservation of garnet growth zoning and the duration of prograde metamorphism
.
Journal of Petrology
 ,
51
,
2327
2347
,
88.
Campomenosi
N.
,
Scambelluri
M.
et al
2021
.
Using the elastic properties of zircon–garnet host-inclusion pairs for thermobarometry of the ultrahigh-pressure Dora-Maira whiteschists: problems and perspectives
.
Contributions to Mineralogy and Petrology
 ,
176
, article
36
,
89.
Cao
M.
,
Qin
K.
,
Li
G.
,
Evans
N.J.
and
Jin
L.
2015
.
In situ LA-(MC)-ICP-MS trace element and Nd isotopic compositions and genesis of polygenetic titanite from the Baogutu reduced porphyry Cu deposit, Western Junggar, NW China
.
Ore Geology Reviews
 ,
65
,
940
954
,
90.
Carruzzo
S.
,
Clarke
D.B.
,
Pelrine
K.M.
and
MacDonald
M.A.
2006
.
Texture, composition, and origin of rutile in the South Mountain Batholith, Nova Scotia
.
The Canadian Mineralogist
 ,
44
,
715
729
,
91.
Carvalho
B.B.
,
Bartoli
O.
et al
2019
.
Anatexis and fluid regime of the deep continental crust: new clues from melt and fluid inclusions in metapelitic migmatites from Ivrea Zone (NW Italy)
.
Journal of Metamorphic Geology
 ,
37
,
951
975
,
92.
Castelli
D.
and
Rubatto
D.
2002
.
Stability of Al- and F-rich titanite in metacarbonate: petrologic and isotopic constraints from a polymetamorphic eclogitic marble of the internal Sesia Zone (Western Alps)
.
Contributions to Mineralogy and Petrology
 ,
142
,
627
639
,
93.
Catlos
E.
,
Sorensen
S.S.
and
Harrison
T.M.
2000
.
Th–Pb ion-microprobe dating of allanite
.
American Mineralogist
 ,
85
,
633
648
,
94.
Cavosie
A.J.
,
Erickson
T.M.
et al
2015
.
A terrestrial perspective on using ex situ shocked zircons to date lunar impacts
.
Geology
 ,
43
,
999
1002
,
95.
Cavosie
A.J.
,
Spencer
C.J.
,
Evans
N.
,
Rankenburg
K.
,
Thomas
R.J.
and
Macey
P.H.
2022
.
Granular titanite from the Roter Kamm crater in Namibia: product of regional metamorphism, not meteorite impact
.
Geoscience Frontiers
 ,
13
, article
101350
,
96.
Caxito
F.A.
,
Basto
C.F.
et al
2021
.
Neoproterozoic magmatic arc volcanism in the Borborema Province, NE Brazil: possible flare-ups and lulls and implications for western Gondwana assembly
.
Gondwana Research
 ,
92
,
1
25
,
97.
Cenki-Tok
B.
,
Oliot
E.
et al
2011
.
Preservation of Permian allanite within an Alpine eclogite facies shear zone at Mt Mucrone, Italy: mechanical and chemical behavior of allanite during mylonitization
.
Lithos
 ,
125
,
40
50
,
98.
Cenki-Tok
B.
,
Darling
J.R.
,
Rolland
Y.
,
Dhuime
B.
and
Storey
C.D.
2014
.
Direct dating of mid-crustal shear zones with synkinematic allanite: new in situ U–Th–Pb geochronological approaches applied to the Mont Blanc Massif
.
Terra Nova
 ,
26
,
29
37
,
99.
Cerny
P.
,
Novak
M.
,
Chapman
R.
and
Ferreira
K.J.
2007
.
Subsolidus behavior of niobian rutile from the Pisek region, Czech Republic: a model for exsolution in W- and Fe2 + >>Fe3 + -rich phases
.
Journal of Geosciences
 ,
52
,
143
159
,
100.
Cerva-Alves
T.
,
Hartmann
L.A.
,
Queiroga
G.N.
,
Lana
C.
,
Castro
M.P.
,
Maciel
L.A.C.
and
Remus
M.V.D.
2021
.
Metamorphic evolution of the juvenile Serrinha forearc basin in the southern Brasiliano Orogen
.
Precambrian Research
 ,
365
, article
106394
,
101.
Cesare
B.
,
Ferrero
S.
,
Salvioli-Mariani
E.
,
Pedron
D.
and
Cavallo
A.
2009
.
‘Nanogranite’ and glassy inclusions: the anatectic melt in migmatites and granulites
.
Geology
 ,
37
,
627
630
,
102.
Cesare
B.
,
Acosta-Vigil
A.
,
Ferrero
S.
,
Bartoli
O.
and
Forster
M.
2011
.
Melt inclusions in migmatites and granulites
.
Journal of the Virtual Explorer
 ,
38
, article
2
,
103.
Chakhmouradian
A.R.
,
Reguir
E.P.
et al
2017
.
Apatite in carbonatitic rocks: compositional variation, zoning, element partitioning and petrogenetic significance
.
Lithos
 ,
274
,
188
213
,
104.
Chamberlain
C.P.
and
Conrad
M.E.
1991
.
Oxygen isotope zoning in garnet
.
Science
 ,
254
,
403
406
,
105.
Chamberlain
K.R.
,
Schmitt
A.K.
et al
2010
.
In situ U–Pb SIMS (IN-SIMS) micro-baddeleyite dating of mafic rocks: method with examples
.
Precambrian Research
 ,
183
,
379
387
,
106.
Chaussidon
M.
,
Deng
Z.
,
Villeneuve
J.
,
Moureau
J.
,
Watson
B.
,
Richter
F.
and
Moynier
F.
2017
.
In situ analysis of non-traditional isotopes by SIMS and LA-MC-ICP-MS: key aspects and the example of Mg isotopes in olivines and silicate glasses
.
Reviews in Mineralogy and Geochemistry
 ,
82
,
127
163
,
107.
Chen
X.
,
Wang
W.
,
Zhang
Z.
,
Nie
N.X.
and
Dauphas
N.
2020
.
Evidence from Ab initio and transport modeling for diffusion-driven zirconium isotopic fractionation in igneous rocks
.
ACS Earth and Space Chemistry
 ,
4
,
1572
1595
,
108.
Cherniak
D.
J.
1993
.
Lead diffusion in titanite and preliminary results on the effects of radiation damage on Pb transport
.
Chemical Geology
 ,
110
,
177
194
,
109.
Cherniak
D.J.
1995
.
Sr and Nd diffusion in titanite
.
Chemical Geology
 ,
125
,
219
232
,
110.
Cherniak
D.J.
2000a
.
Pb diffusion in rutile
.
Contributions to Mineralogy and Petrology
 ,
139
,
198
207
,
111.
Cherniak
D.J.
2000b
.
Rare earth element diffusion in apatite
.
Geochimica et Cosmochimica Acta
 ,
64
,
3871
3885
,
112.
Cherniak
D.J.
2006
.
Pb and rare earth element diffusion in xenotime
.
Lithos
 ,
88
,
1
14
,
113.
Cherniak
D.J.
and
Watson
E.B.
2001
.
Pb diffusion in zircon
.
Chemical Geology
 ,
172
,
5
24
,
114.
Cherniak
D.J.
and
Watson
E.B.
2003
.
Diffusion in zircon
.
Reviews in Mineralogy and Geochemistry
 ,
53
,
113
143
,
115.
Cherniak
D.J.
and
Watson
E.
2007
.
Ti diffusion in zircon
.
Chemical Geology
 ,
242
,
470
483
,
116.
Cherniak
D.J.
,
Lanford
W.A.
and
Ryerson
F.J.
1991
.
Lead diffusion in apatite and zircon using ion implantation and Rutherford backscattering techniques
.
Geochimica et Cosmochimica Acta
 ,
55
,
1663
1673
,
117.
Cherniak
D.J.
,
Watson
E.B.
,
Grove
M.
and
Harrison
T.M.
2004
.
Pb diffusion in monazite: a combined RBS/SIMS study
.
Geochimica et Cosmochimica Acta
 ,
68
,
829
840
,
118.
Cherniak
D.J.
,
Manchester
J.
and
Watson
E.
2007
.
Zr and Hf diffusion in rutile
.
Earth and Planetary Science Letters
 ,
261
,
267
279
,
119.
Chew
D.M.
and
Spikings
R.A.
2015
.
Geochronology and thermochronology using apatite: time and temperature, lower crust to surface
.
Elements
 ,
11
,
189
194
,
120.
Chew
D.M.
,
Sylvester
P.J.
and
Tubrett
M.N.
2011
.
U–Pb and Th–Pb dating of apatite by LA-ICPMS
.
Chemical Geology
 ,
280
,
200
216
,
121.
Chew
D.M.
,
Petrus
J.A.
and
Kamber
B.S.
2014
.
U–Pb LA-ICPMS dating using accessory mineral standards with variable common Pb
.
Chemical Geology
 ,
363
,
185
199
,
122.
Chew
D.M.
,
Petrus
J.A.
,
Kenny
G.G.
and
McEvoy
N.
2017
.
Rapid high-resolution U–Pb LA-Q-ICPMS age mapping of zircon
.
Journal of Analytical Atomic Spectrometry
 ,
32
,
262
276
,
123.
Chew
D.
,
Drost
K.
,
Marsh
J.H.
and
Petrus
J.A.
2021
.
LA-ICP-MS imaging in the geosciences and its applications to geochronology
.
Chemical Geology
 ,
559
, article
119917
,
124.
Choukroun
M.
,
O'Reilly
S.Y.
,
Griffin
W.L.
,
Pearson
N.J.
and
Dawson
J.B.
2005
.
Hf isotopes of MARID (mica–amphibole–rutile–ilmenite–diopside) rutile trace metasomatic processes in the lithospheric mantle
.
Geology
 ,
33
,
45
48
,
125.
Chu
M.-F.
,
Wang
K.-L.
,
Griffin
W.L.
,
Chung
S.-L.
,
O'Reilly
S.Y.
,
Pearson
N.J.
and
Iizuka
Y.
2009
.
Apatite composition: tracing petrogenetic processes in Transhimalayan granitoids
.
Journal of Petrology
 ,
50
,
1829
1855
,
126.
Cioffi
C.R.
,
da Campos Neto
M.C.
,
Möller
A.
and
Rocha
B.C.
2019
.
Titanite petrochronology of the southern Brasília Orogen basement: effects of retrograde net-transfer reactions on titanite trace element compositions
.
Lithos
 ,
344–345
,
393
408
,
127.
Clark
C.
,
Collins
A.S.
,
Santosh
M.
,
Taylor
R.
and
Wade
B.P.
2009
.
The P–T–t architecture of a Gondwanan suture: REE, U–Pb and Ti-in-zircon thermometric constraints from the Palghat Cauvery shear system, South India
.
Precambrian Research
 ,
174
,
129
144
,
128.
Clark
C.
,
Kirkland
C.L.
,
Spaggiari
C.V.
,
Oorschot
C.
,
Wingate
M.T.D.
and
Taylor
R.J.
2014
.
Proterozoic granulite formation driven by mafic magmatism: an example from the Fraser Range Metamorphics, Western Australia
.
Precambrian Research
 ,
240
,
1
21
,
129.
Clark
C.
,
Taylor
R.J.M.
,
Johnson
T.E.
,
Harley
S.L.
,
Fitzsimons
I.C.W.
and
Oliver
L.
2019
.
Testing the fidelity of thermometers at ultrahigh temperatures
.
Journal of Metamorphic Geology
 ,
37
,
917
934
,
130.
Clark
J.R.
and
Williams-Jones
A.E.
2004
.
Rutile as a Potential Indicator Mineral for Metamorphosed Metallic Ore Deposits
 .
Rapport Final de DIVEX, Sous-projet SC2
,
Montréal, Canada
.
131.
Cochrane
R.
,
Spikings
R.A.
et al
2014
.
High temperature (>350°C) thermochronology and mechanisms of Pb loss in apatite
.
Geochimica et Cosmochimica Acta
 ,
127
,
39
56
,
132.
Coleman
D.S.
,
Gray
W.
and
Glazner
A.F.
2004
.
Rethinking the emplacement and evolution of zoned plutons: geochronologic evidence for incremental assembly of the Tuolumne Intrusive Suite, California
.
Geology
 ,
32
,
433
436
,
133.
Collins
W.J.
,
Huang
H.-Q.
and
Jiang
X.
2016
.
Water-fluxed crustal melting produces Cordilleran batholiths
.
Geology
 ,
44
,
143
146
,
134.
Copeland
P.
,
Parrish
R.R.
and
Harrison
T.M.
1988
.
Identification of inherited radiogenic Pb in monazite and its implications for U–Pb systematics
.
Nature
 ,
333
,
760
763
,
135.
Corfu
F.
,
Hanchar
J.M.
,
Hoskin
P.W.O.
and
Kinny
P.
2003
.
Atlas of zircon textures
.
Reviews in Mineralogy and Geochemistry
 ,
53
,
469
500
,
136.
Corti
L.
,
Zanoni
D.
,
Gatta
G.D.
and
Zucali
M.
2020
.
Strain partitioning in host rock controls light rare earth element release from allanite-(Ce) in subduction zones
.
Mineralogical Magazine
 ,
84
,
93
108
,
137.
Cox
M.A.
,
Cavosie
A.J.
et al
2022
.
Impact and habitability scenarios for early Mars revisited based on a 4.45-Ga shocked zircon in regolith breccia
.
Science Advances
 ,
8
, article
eabl7497
,
138.
Cross
A.J.
2009
.
SHRIMP U–Pb Xenotime Geochronology and Its Application to Dating Mineralisation, Sediment Deposition and Metamorphism
 . PhD thesis,
Australian National University
,
139.
Crowe
D.E.
,
Riciputi
L.R.
,
Bezenek
S.
and
Ignatiev
A.
2001
.
Oxygen isotope and trace element zoning in hydrothermal garnets: windows into large-scale fluid-flow behavior
.
Geology
 ,
29
,
479
482
,
140.
Cruz-Uribe
A.M.
,
Feineman
M.D.
,
Zack
T.
and
Barth
M.
2014
.
Metamorphic reaction rates at ∼650–800°C from diffusion of niobium in rutile
.
Geochimica et Cosmochimica Acta
 ,
130
,
63
77
,
141.
Cutts
K.A.
,
Hand
M.
,
Kelsey
D.E.
,
Wade
B.
,
Strachan
R.A.
,
Clark
C.
and
Netting
A.
2009
.
Evidence for 930 Ma metamorphism in the Shetland Islands, Scottish Caledonides: implications for Neoproterozoic tectonics in the Laurentia–Baltica sector of Rodinia
.
Journal of the Geological Society, London
 ,
166
,
1033
1047
,
142.
Cutts
K.A.
,
Kinny
P.D.
et al
2010
.
Three metamorphic events recorded in a single garnet: integrated phase modelling, in situ LA-ICPMS and SIMS geochronology from the Moine Supergroup, NW Scotland
.
Journal of Metamorphic Geology
 ,
28
,
249
267
,
143.
Cutts
K.A.
,
Kelsey
D.E.
and
Hand
M.
2013
.
Evidence for late Paleoproterozoic (ca 1690–1665 Ma) high- to ultrahigh-temperature metamorphism in southern Australia: implications for Proterozoic supercontinent models
.
Gondwana Research
 ,
23
,
617
640
,
144.
Cutts
K.A.
,
Stevens
G.
,
Hoffmann
J.E.
,
Buick
I.S.
,
Frei
D.
and
Münker
C.
2014
.
Paleo- to Mesoarchean polymetamorphism in the Barberton Granite–Greenstone Belt, South Africa: constraints from U–Pb monazite and Lu–Hf garnet geochronology on the tectonic processes that shaped the belt
.
GSA Bulletin
 ,
126
,
251
270
,
145.
Cutts
K.
,
Lana
C.
,
Alkmim
F.
and
Peres
G.G.
2018
.
Metamorphic imprints on units of the southern Araçuaí belt, SE Brazil: the history of superimposed Transamazonian and Brasiliano orogenesis
.
Gondwana Research
 ,
58
,
211
234
,
146.
Cygan
R.T.
and
Lasaga
A.C.
1985
.
Self-diffusion of magnesium in garnet at 750 degrees to 900 degrees C
.
American Journal of Science
 ,
285
,
328
350
,
147.
Dachille
F.
,
Simons
P.Y.
and
Roy
R.
1968
.
Pressure–temperature studies of anatase, brookite, rutile and TiO2-II
.
American Mineralogist
 ,
53
,
1929
1939
.
148.
Darling
J.R.
,
Storey
C.D.
and
Engi
M.
2012
.
Allanite U–Th–Pb geochronology by laser ablation ICPMS
.
Chemical Geology
 ,
292
,
103
115
,
149.
Davis
D.W.
,
Krogh
T.E.
and
Williams
I.S.
2003
.
Historical development of zircon geochronology
.
Reviews in Mineralogy and Geochemistry
 ,
53
,
145
181
,
150.
Decrée
S.
,
Coint
N.
,
Debaille
V.
,
Hagen-Peter
G.
,
Leduc
T.
and
Schiellerup
H.
2023
.
The potential for REEs in igneous-related apatite deposits in Europe
.
Geological Society, London, Special Publications
 ,
526
,
2021–2175
,
151.
Degeling
H.
,
Eggins
S.
and
Ellis
D.
2001
.
Zr budgets for metamorphic reactions, and the formation of zircon from garnet breakdown
.
Mineralogical Magazine
 ,
65
,
749
758
,
152.
Delavault
H.
,
Dhuime
B.
,
Hawkesworth
C.J.
,
Cawood
P.A.
,
Marschall
H.
and the
Edinburgh Ion Microprobe Facility
2016
.
Tectonic settings of continental crust formation: insights from Pb isotopes in feldspar inclusions in zircon
.
Geology
 ,
44
,
819
822
,
153.
Dempster
T.J.
,
Coleman
S.
,
Kennedy
R.
,
Chung
P.
and
Brown
R.W.
2020
.
Growth zoning of garnet porphyroblasts: grain boundary and microtopographic controls
.
Journal of Metamorphic Geology
 ,
38
,
1011
1027
,
154.
Deng
Z.
,
Chaussidon
M.
,
Guitreau
M.
,
Puchtel
I.S.
,
Dauphas
N.
and
Moynier
F.
2019
.
An oceanic subduction origin for Archaean granitoids revealed by silicon isotopes
.
Nature Geoscience
 ,
12
,
774
778
,
155.
D'Errico
M.
,
Lackey
J.
et al
2012
.
A detailed record of shallow hydrothermal fluid flow in the Sierra Nevada magmatic arc from low-δ18O skarn garnets
.
Geology
 ,
40
,
763
766
,
156.
Dev
J.A.
,
Tomson
J.K.
,
Sorcar
N.
and
Nandakumar
V.
2021
.
Combined U–Pb/Hf isotopic studies and phase equilibrium modelling of HT–UHT metapelites from Kambam ultrahigh-temperature belt, South India: constraints on tectonothermal history of the terrane
.
Lithos
 ,
406–407
, article
106531
,
157.
Dhuime
B.
,
Hawkesworth
C.J.
,
Cawood
P.A.
and
Storey
C.D.
2012
.
A change in the geodynamics of continental growth 3 billion years ago
.
Science
 ,
335
,
1334
1336
,
158.
Di Rosa
M.
,
Farina
F.
,
Lanari
P.
and
Marroni
M.
2020
.
Pre-Alpine thermal history recorded in the continental crust from Alpine Corsica (France): evidence from zircon and allanite LA-ICP-MS dating
.
Swiss Journal of Geosciences
 ,
113
, article
19
,
159.
Dodson
M.H.
1973
.
Closure temperature in cooling geochronological and petrological systems
.
Contributions to Mineralogy and Petrology
 ,
40
,
259
274
,
160.
Dohmen
R.
,
Marschall
H.R.
,
Ludwig
T.
and
Polednia
J.
2019
.
Diffusion of Zr, Hf, Nb and Ta in rutile: effects of temperature, oxygen fugacity, and doping level, and relation to rutile point defect chemistry
.
Physics and Chemistry of Minerals
 ,
46
,
311
332
,
161.
Doucelance
R.
,
Bruand
E.
,
Matte
S.
,
Bosq
C.
,
Auclair
D.
and
Gannoun
A.-M.
2020
.
In-situ determination of Nd isotope ratios in apatite by LA-MC-ICPMS: challenges and limitations
.
Chemical Geology
 ,
550
, article
119740
,
162.
Dragovic
B.
,
Samanta
L.M.
,
Baxter
E.F.
and
Selverstone
J.
2012
.
Using garnet to constrain the duration and rate of water-releasing metamorphic reactions during subduction: an example from Sifnos, Greece
.
Chemical Geology
 ,
314–317
,
9
22
,
163.
Dragovic
B.
,
Baxter
E.F.
and
Caddick
M.J.
2015
.
Pulsed dehydration and garnet growth during subduction revealed by zoned garnet geochronology and thermodynamic modeling, Sifnos, Greece
.
Earth and Planetary Science Letters
 ,
413
,
111
122
,
164.
Duan
Z.
,
Gleeson
S.A.
,
Gao
W.-S.
,
Wang
F.-Y.
,
Li
C.-J.
and
Li
J.-W.
2020
.
Garnet U–Pb dating of the Yinan Au–Cu skarn deposit, Luxi District, North China Craton: implications for district-wide coeval Au–Cu and Fe skarn mineralization
.
Ore Geology Reviews
 ,
118
, article
103310
,
165.
Duchêne
S.
,
Blichert-Toft
J.
,
Luais
B.
,
Télouk
P.
,
Lardeaux
J.-M.
and
Albarède
F.
1997
.
The Lu–Hf dating of garnets and the ages of the Alpine high-pressure metamorphism
.
Nature
 ,
387
,
586
589
,
166.
Duesterhoeft
E.
and
Lanari
P.
2020
.
Iterative thermodynamic modelling – part 1: a theoretical scoring technique and a computer program (Bingo-Antidote)
.
Journal of Metamorphic Geology
 ,
38
,
527
551
,
167.
Durgalakshmi
Sajeev
K.
et al
2021
.
The timing, duration and conditions of UHT metamorphism in remnants of the former eastern Gondwana
.
Journal of Petrology
 ,
62
, article
egab068
,
168.
Dziggel
A.
,
Wulff
K.
,
Kolb
J.
,
Meyer
F.M.
and
Lahaye
Y.
2009
.
Significance of oscillatory and bell-shaped growth zoning in hydrothermal garnet: evidence from the Navachab gold deposit, Namibia
.
Chemical Geology
 ,
262
,
262
276
,
169.
El Korh
A.
,
Schmidt
S.T.
,
Ulianov
A.
and
Potel
S.
2009
.
Trace element partitioning in HP–LT metamorphic assemblages during subduction-related metamorphism, Ile de Groix, France: a detailed LA-ICPMS study
.
Journal of Petrology
 ,
50
,
1107
1148
,
170.
Emo
R.B.
,
Smit
M.A.
et al
2018
.
Evidence for evolved Hadean crust from Sr isotopes in apatite within Eoarchean zircon from the Acasta Gneiss Complex
.
Geochimica et Cosmochimica Acta
 ,
235
,
450
462
,
171.
Engi
M.
2017
.
Petrochronology based on REE-minerals: monazite, allanite, xenotime, apatite
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
365
418
,
172.
Engi
M.
,
Lanari
P.
and
Kohn
M.J.
2017
.
Significant ages – an introduction to petrochronology
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
1
12
,
173.
Evans
J.
and
Zalasiewicz
J.
1996
.
UPb, PbPb and SmNd dating of authigenic monazite: implications for the diagenetic evolution of the Welsh Basin
.
Earth and Planetary Science Letters
 ,
144
,
421
433
,
174.
Evans
T.
P.
2004
.
A method for calculating effective bulk composition modification due to crystal fractionation in garnet-bearing schist: implications for isopleth thermobarometry
.
Journal of Metamorphic Geology
 ,
22
,
547
557
,
175.
Ewing
R.C.
,
Meldrum
A.
,
Wang
L.
,
Weber
W.J.
and
Corrales
L.R.
2003
.
Radiation effects in zircon
.
Reviews in Mineralogy and Geochemistry
 ,
53
,
387
425
,
176.
Ewing
T.A.
and
Müntener
O.
2018
.
The mantle source of island arc magmatism during early subduction: evidence from Hf isotopes in rutile from the Jijal Complex (Kohistan arc, Pakistan)
.
Lithos
 ,
308–309
,
262
277
,
177.
Ewing
T.A.
,
Rubatto
D.
,
Eggins
S.M.
and
Hermann
J.
2011
.
In situ measurement of hafnium isotopes in rutile by LA-MC-ICPMS: protocol and applications
.
Chemical Geology
 ,
281
,
72
82
,
178.
Ewing
T.A.
,
Hermann
J.
and
Rubatto
D.
2013
.
The robustness of the Zr-in-rutile and Ti-in-zircon thermometers during high-temperature metamorphism (Ivrea–Verbano Zone, northern Italy)
.
Contributions to Mineralogy and Petrology
 ,
165
,
757
779
,
179.
Ewing
T.A.
,
Rubatto
D.
and
Hermann
J.
2014
.
Hafnium isotopes and Zr/Hf of rutile and zircon from lower crustal metapelites (Ivrea–Verbano Zone, Italy): implications for chemical differentiation of the crust
.
Earth and Planetary Science Letters
 ,
389
,
106
118
,
180.
Ewing
T.A.
,
Rubatto
D.
,
Beltrando
M.
and
Hermann
J.
2015
.
Constraints on the thermal evolution of the Adriatic margin during Jurassic continental break-up: U–Pb dating of rutile from the Ivrea–Verbano Zone, Italy
.
Contributions to Mineralogy and Petrology
 ,
169
,
1
22
,
181.
Farias
P.
,
Weinberg
R.
,
Sola
A.
and
Becchio
R.
2020
.
From crustal thickening to orogen-parallel escape: the 120-myr-long HT–LP evolution recorded by titanite in the Paleozoic Famatinian Backarc, NW Argentina
.
Tectonics
 ,
39
, article
e2020TC006184
,
182.
Ferreira
J.A.
,
Pereira
I.
,
Bento dos Santos
T.
and
Mata
J.
2022
.
U–Pb age constraints on the protolith, cooling and exhumation of a Variscan middle crust migmatite complex from the Central Iberian Zone: insights into the Variscan metamorphic evolution and Ediacaran palaeogeographic implications
.
Journal of the Geological Society
 ,
179
,
183.
Ferrero
S.
,
Bartoli
O.
et al
2012
.
Microstructures of melt inclusions in anatectic metasedimentary rocks
.
Journal of Metamorphic Geology
 ,
30
,
303
322
,
184.
Ferrero
S.
,
Wannhoff
I.
et al
2021
.
Embryos of TTGs in Gore Mountain garnet megacrysts from water-fluxed melting of the lower crust
.
Earth and Planetary Science Letters
 ,
569
, article
117058
,
185.
Ferry
J.M.
and
Spear
F.
1978
.
Experimental calibration of the partitioning of Fe and Mg between biotite and garnet
.
Contributions to Mineralogy and Petrology
 ,
66
,
113
117
,
186.
Ferry
J.M.
and
Watson
E.B.
2007
.
New thermodynamic models and revised calibrations for the Ti-in-zircon and Zr-in-rutile thermometers
.
Contributions to Mineralogy and Petrology
 ,
154
,
429
437
,
187.
Fielding
I.O.H.
,
Johnson
S.P.
et al
2017
.
Using in situ SHRIMP U–Pb monazite and xenotime geochronology to determine the age of orogenic gold mineralization: an example from the Paulsens mine, southern Pilbara Craton
.
Economic Geology
 ,
112
,
1205
1230
,
188.
Finch
M.
,
Weinberg
R.
,
Barrote
V.
and
Cawood
P.
2021
.
Hf isotopic ratios in zircon reveal processes of anatexis and pluton construction
.
Earth and Planetary Science Letters
 ,
576
, article
117215
,
189.
Finger
F.
and
Krenn
E.
2007
.
Three metamorphic monazite generations in a high-pressure rock from the Bohemian Massif and the potentially important role of apatite in stimulating polyphase monazite growth along a PT loop
.
Lithos
 ,
95
,
103
115
,
190.
Fischer
S.
,
Prave
A.R.
,
Johnson
T.E.
,
Cawood
P.A.
,
Hawkesworth
C.J.
,
Horstwood
M.S.A.
and
EIMF
2021
.
Using zircon in mafic migmatites to disentangle complex high-grade gneiss terrains – terrane spotting in the Lewisian complex, NW Scotland
.
Precambrian Research
 ,
355
, article
106074
,
191.
Fisher
C.M.
,
McFarlane
C.R.M.
,
Hanchar
J.M.
,
Schmitz
M.D.
,
Sylvester
P.J.
,
Lam
R.
and
Longerich
H.P.
2011
.
Sm–Nd isotope systematics by laser ablation-multicollector-inductively coupled plasma mass spectrometry: methods and potential natural and synthetic reference materials
.
Chemical Geology
 ,
284
,
1
20
,
192.
Fisher
C.M.
,
Vervoort
J.D.
and
DuFrane
S.A.
2014a
.
Accurate Hf isotope determinations of complex zircons using the ‘laser ablation split stream’ method
.
Geochemistry, Geophysics, Geosystems
 ,
15
,
121
139
,
193.
Fisher
C.M.
,
Vervoort
J.D.
and
Hanchar
J.M.
2014b
.
Guidelines for reporting zircon Hf isotopic data by LA-MC-ICPMS and potential pitfalls in the interpretation of these data
.
Chemical Geology
 ,
363
,
125
133
,
194.
Fisher
C.M.
,
Hanchar
J.M.
,
Miller
C.F.
,
Phillips
S.
,
Vervoort
J.D.
and
Whitehouse
M.J.
2017
.
Combining Nd isotopes in monazite and Hf isotopes in zircon to understand complex open-system processes in granitic magmas
.
Geology
 ,
45
,
267
270
,
195.
Fisher
C.M.
,
Bauer
A.M.
et al
2020
.
Laser ablation split-stream analysis of the Sm–Nd and U–Pb isotope compositions of monazite, titanite, and apatite – improvements, potential reference materials, and application to the Archean Saglek Block gneisses
.
Chemical Geology
 ,
539
, article
119493
,
196.
Fletcher
I.R.
,
McNaughton
N.J.
,
Aleinikoff
J.A.
,
Rasmussen
B.
and
Kamo
S.L.
2004
.
Improved calibration procedures and new standards for U–Pb and Th–Pb dating of Phanerozoic xenotime by ion microprobe
.
Chemical Geology
 ,
209
,
295
314
,
197.
Flowerdew
M.
,
Millar
I.L.
,
Vaughan
A.
,
Horstwood
M.
and
Fanning
C.
2006
.
The source of granitic gneisses and migmatites in the Antarctic Peninsula: a combined U–Pb SHRIMP and laser ablation Hf isotope study of complex zircons
.
Contributions to Mineralogy and Petrology
 ,
151
,
751
768
,
198.
Foley
S.
,
Tiepolo
M.
and
Vannucci
R.
2002
.
Growth of early continental crust controlled by melting of amphibolite in subduction zones
.
Nature
 ,
417
,
837
840
,
199.
Fornelli
A.
,
Langone
A.
,
Micheletti
F.
,
Pascazio
A.
and
Piccarreta
G.
2014
.
The role of trace element partitioning between garnet, zircon and orthopyroxene on the interpretation of zircon U–Pb ages: an example from high-grade basement in Calabria (southern Italy)
.
International Journal of Earth Sciences
 ,
103
,
487
507
,
200.
Fornelli
A.
,
Langone
A.
,
Micheletti
F.
and
Piccarreta
G.
2018
.
REE partition among zircon, orthopyroxene, amphibole and garnet in a high-grade metabasic system
.
Geological Magazine
 ,
155
,
1705
1726
,
201.
Förster
H.-J.
1998
.
The chemical composition of REE–Y–Th–U-rich accessory minerals in peraluminous granites of the Erzgebirge–Fichtelgebirge region, Germany, part I: the monazite-(Ce)-brabantite solid solution series
.
American Mineralogist
 ,
83
,
259
272
,
202.
Fortier
S.M.
and
Giletti
B.J.
1989
.
An empirical model for predicting diffusion coefficients in silicate minerals
.
Science
 ,
245
,
1481
1484
,
203.
Foster
G.
,
Gibson
H.D.
,
Parrish
R.
,
Horstwood
M.
,
Fraser
J.
and
Tindle
A.
2002
.
Textural, chemical and isotopic insights into the nature and behaviour of metamorphic monazite
.
Chemical Geology
 ,
191
,
183
207
,
204.
Fougerouse
D.
,
Kirkland
C.L.
,
Saxey
D.W.
,
Seydoux-Guillaume
A.-M.
,
Rowles
M.R.
,
Rickard
W.D.A.
and
Reddy
S.M.
2020
.
Nanoscale isotopic dating of monazite
.
Geostandards and Geoanalytical Research
 ,
44
,
637
652
,
205.
Fougerouse
D.
,
Cavosie
A.J.
et al
2021
.
A new method for dating impact events – thermal dependency on nanoscale Pb mobility in monazite shock twins
.
Geochimica et Cosmochimica Acta
 ,
314
,
381
396
,
206.
Fraser
G.
,
Ellis
D.
and
Eggins
S.
1997
.
Zirconium abundance in granulite-facies minerals, with implications for zircon geochronology in high-grade rocks
.
Geology
 ,
25
,
607
610
,
207.
Frost
B.R.
,
Chamberlain
K.R.
and
Schumacher
J.C.
2001
.
Sphene (titanite): phase relations and role as a geochronometer
.
Chemical Geology
 ,
172
,
131
148
,
208.
Froude
D.O.
,
Ireland
T.R.
,
Kinny
P.D.
,
Williams
I.S.
,
Compston
W.
,
Williams
I.R.
and
Myers
J.S.
1983
.
Ion microprobe identification of 4,100–4,200 Myr-old terrestrial zircons
.
Nature
 ,
304
,
616
618
,
209.
Fu
B.
,
Page
F.Z.
et al
2008
.
Ti-in-zircon thermometry: applications and limitations
.
Contributions to Mineralogy and Petrology
 ,
156
,
197
215
,
210.
Fumes
R.A.
,
Luvizotto
G.L.
et al
2022
.
Petrochronology of high-pressure granulite facies rocks from Southern Brasília Orogen, SE Brazil: combining quantitative compositional mapping, single-element thermometry and geochronology
.
Journal of Metamorphic Geology
 ,
40
,
517
552
,
211.
Gaidies
F.
,
de Capitani
C.
and
Abart
R.
2008
.
THERIA_G: a software program to numerically model prograde garnet growth
.
Contributions to Mineralogy and Petrology
 ,
155
,
657
671
,
212.
Gaidies
F.
,
Morneau
Y.E.
,
Petts
D.C.
,
Jackson
S.E.
,
Zagorevski
A.
and
Ryan
J.J.
2021
.
Major and trace element mapping of garnet: unravelling the conditions, timing and rates of metamorphism of the Snowcap assemblage, west–central Yukon
.
Journal of Metamorphic Geology
 ,
39
,
133
164
,
213.
Ganguly
J.
and
Saxena
S.K.
1984
.
Mixing properties of aluminosilicate garnets: constraints from natural and experimental data, and applications to geothermo-barometry
.
American Mineralogist
 ,
69
,
88
97
.
214.
Gao
X.-Y.
,
Zheng
Y.-F.
,
Chen
Y.-X.
and
Guo
J.
2012
.
Geochemical and U–Pb age constraints on the occurrence of polygenetic titanites in UHP metagranite in the Dabie orogen
.
Lithos
 ,
136–139
,
93
108
,
215.
Garber
J.M.
,
Hacker
B.R.
,
Kylander-Clark
A.R.C.
,
Stearns
M.
and
Seward
G.
2017
.
Controls on trace element uptake in metamorphic titanite: implications for petrochronology
.
Journal of Petrology
 ,
58
,
1031
1057
,
216.
Gardiner
N.J.
,
Kirkland
C.L.
,
Hollis
J.
,
Szilas
K.
,
Steenfelt
A.
,
Yakymchuk
C.
and
Heide-Jørgensen
H.
2019
.
Building Mesoarchaean crust upon Eoarchaean roots: the Akia Terrane, West Greenland
.
Contributions to Mineralogy and Petrology
 ,
174
, article
20
,
217.
George
F.
,
Gaidies
F.
and
Boucher
B.
2018
.
Population-wide garnet growth zoning revealed by LA-ICP-MS mapping: implications for trace element equilibration and syn-kinematic deformation during crystallisation
.
Contributions to Mineralogy and Petrology
 ,
173
,
1
22
,
218.
Gerdes
A.
and
Zeh
A.
2009
.
Zircon formation v. zircon alteration – new insights from combined U–Pb and Lu–Hf in-situ LA-ICP-MS analyses, and consequences for the interpretation of Archean zircon from the Central Zone of the Limpopo Belt
.
Chemical Geology
 ,
261
,
230
243
,
219.
Gerrits
A.R.
,
Inglis
E.C.
,
Dragovic
B.
,
Starr
P.G.
,
Baxter
E.F.
and
Burton
K.W.
2019
.
Release of oxidizing fluids in subduction zones recorded by iron isotope zonation in garnet
.
Nature Geoscience
 ,
12
,
1029
1033
,
220.
Gervasoni
F.
,
Klemme
S.
,
Rocha-Júnior
E.R.V.
and
Berndt
J.
2016
.
Zircon saturation in silicate melts: a new and improved model for aluminous and alkaline melts
.
Contributions to Mineralogy and Petrology
 ,
171
, article
21
,
221.
Gevedon
M.
,
Seman
S.
,
Barnes
J.D.
,
Lackey
J.S.
and
Stockli
D.F.
2018
.
Unraveling histories of hydrothermal systems via U–Pb laser ablation dating of skarn garnet
.
Earth and Planetary Science Letters
 ,
498
,
237
246
,
222.
Ghent
E.D.
1976
.
Plagioclase–garnet–Al2SiO5–quartz; a potential geobarometer–geothermometer
.
American Mineralogist
 ,
61
,
710
714
.
223.
Gieré
R.
and
Sorensen
S.S.
2004
.
Allanite and other REE-rich epidote-group minerals
.
Reviews in Mineralogy and Geochemistry
 ,
56
,
431
493
,
224.
Gillespie
J.
,
Kinny
P.D.
,
Kirkland
C.L.
,
Martin
L.
,
Nemchin
A.A.
,
Cavosie
A.J.
and
Hasterok
D.
2021
.
Isotopic modelling of Archean crustal evolution from comagmatic zircon–apatite pairs
.
Earth and Planetary Science Letters
 ,
575
, article
117194
,
225.
Gillespie
J.
,
Kirkland
C.L.
,
Kinny
P.D.
,
Simpson
A.
,
Glorie
S.
and
Rankenburg
K.
2022
.
Lu–Hf, Sm–Nd, and U–Pb isotopic coupling and decoupling in apatite
.
Geochimica et Cosmochimica Acta
 ,
338
,
121
135
,
226.
Giuntoli
F.
,
Lanari
P.
,
Burn
M.
,
Kunz
B.E.
and
Engi
M.
2018
.
Deeply subducted continental fragments – part 2: insight from petrochronology in the central Sesia Zone (western Italian Alps)
.
Solid Earth
 ,
9
,
191
222
,
227.
Glorie
S.
,
Gillespie
J.
et al
2022
.
Detrital apatite Lu–Hf and U–Pb geochronology applied to the southwestern Siberian margin
.
Terra Nova
 ,
34
,
201
209
,
228.
Glorie
S.
et al
2023
.
Robust laser ablation Lu–Hf dating of apatite: an empirical evaluation
.
Geological Society, London, Special Publications
 ,
537
,
229.
Godet
A.
,
Raimondo
T.
and
Guilmette
C.
2022
.
Atoll garnet: insights from LA-ICP-MS trace element mapping
.
Contributions to Mineralogy and Petrology
 ,
177
, article
57
,
230.
Gonçalves
G.O.
,
Lana
C.
,
Buick
I.S.
,
Alkmim
F.F.
,
Scholz
R.
and
Queiroga
G.
2019
.
Twenty million years of post-orogenic fluid production and hydrothermal mineralization across the external Araçuaí orogen and adjacent São Francisco craton, SE Brazil
.
Lithos
 ,
342–343
,
557
572
,
231.
Gonzalez
J.P.
,
Thomas
J.B.
,
Baldwin
S.L.
and
Alvaro
M.
2019
.
Quartz-in-garnet and Ti-in-quartz thermobarometry: methodology and first application to a quartzofeldspathic gneiss from eastern Papua New Guinea
.
Journal of Metamorphic Geology
 ,
37
,
1193
1208
,
232.
Gonzalez
J.P.
,
Mazzucchelli
M.L.
,
Angel
R.J.
and
Alvaro
M.
2021
.
Elastic geobarometry for anisotropic inclusions in anisotropic host minerals: quartz-in-zircon
.
Journal of Geophysical Research: Solid Earth
 ,
126
, article
e2021JB022080
,
233.
Goodenough
K.M.
,
Crowley
Q.G.
,
Krabbendam
M.
and
Parry
S.F.
2013
.
New U–Pb age constraints for the Laxford Shear Zone, NW Scotland: evidence for tectono-magmatic processes associated with the formation of a Paleoproterozoic supercontinent
.
Precambrian Research
 ,
233
,
1
19
,
234.
Gordon
S.M.
,
Kirkland
C.L.
et al
2021
.
Deformation-enhanced recrystallization of titanite drives decoupling between U–Pb and trace elements
.
Earth and Planetary Science Letters
 ,
560
, article
116810
,
235.
Goudie
D.J.
,
Fisher
C.M.
,
Hanchar
J.M.
,
Crowley
J.L.
and
Ayers
J.C.
2014
.
Simultaneous in situ determination of U–Pb and Sm–Nd isotopes in monazite by laser ablation ICP-MS
.
Geochemistry, Geophysics, Geosystems
 ,
15
,
2575
2600
,
236.
Gratz
R.
and
Heinrich
W.
1997
.
Monazite–xenotime thermobarometry: experimental calibration of the miscibility gap in the binary system CePO4–YPO4
.
American Mineralogist
 ,
82
,
772
780
,
237.
Gratz
R.
and
Heinrich
W.
1998
.
Monazite–xenotime thermometry. III. Experimental calibration of the partitioning of gadolinium between monazite and xenotime
.
European Journal of Mineralogy
 ,
10
,
579
588
,
238.
Green
E.C.R.
,
White
R.W.
,
Diener
J.F.A.
,
Powell
R.
,
Holland
T.J.B.
and
Palin
R.M.
2016
.
Activity–composition relations for the calculation of partial melting equilibria in metabasic rocks
.
Journal of Metamorphic Geology
 ,
34
,
845
869
,
239.
Gregory
C.J.
,
Rubatto
D.
,
Allen
C.M.
,
Williams
I.S.
,
Hermann
J.
and
Ireland
T.
2007
.
Allanite micro-geochronology: a LA-ICP-MS and SHRIMP U–Th–Pb study
.
Chemical Geology
 ,
245
,
162
182
,
240.
Gregory
C.J.
,
Rubatto
D.
,
Hermann
J.
,
Berger
A.
and
Engi
M.
2012
.
Allanite behaviour during incipient melting in the southern Central Alps
.
Geochimica et Cosmochimica Acta
 ,
84
,
433
458
,
241.
Grimes
C.B.
,
John
B.E.
et al
2007
.
Trace element chemistry of zircons from oceanic crust: a method for distinguishing detrital zircon provenance
.
Geology
 ,
35
,
643
646
,
242.
Grimes
C.B.
,
Wooden
J.L.
,
Cheadle
M.J.
and
John
B.E.
2015
.
‘Fingerprinting’ tectono-magmatic provenance using trace elements in igneous zircon
.
Contributions to Mineralogy and Petrology
 ,
170
,
1
26
.
243.
Gualda
G.A.R.
,
Ghiorso
M.S.
,
Lemons
R.V.
and
Carley
T.L.
2012
.
Rhyolite-MELTS: a modified calibration of MELTS optimized for silica-rich, fluid-bearing magmatic systems
.
Journal of Petrology
 ,
53
,
875
890
,
244.
Gudelius
D.
,
Zeh
A.
,
Almeev
R.R.
,
Wilson
A.H.
,
Fischer
L.A.
and
Schmitt
A.K.
2020
.
Zircon melt inclusions in mafic and felsic rocks of the Bushveld Complex – constraints for zircon crystallization temperatures and partition coefficients
.
Geochimica et Cosmochimica Acta
 ,
289
,
158
181
,
245.
Guitreau
M.
and
Flahaut
J.
2019
.
Record of low-temperature aqueous alteration of Martian zircon during the late Amazonian
.
Nature Communications
 ,
10
,
1
9
,
246.
Guitreau
M.
,
Blichert-Toft
J.
,
Martin
H.
,
Mojzsis
S.J.
and
Albarède
F.
2012
.
Hafnium isotope evidence from Archean granitic rocks for deep-mantle origin of continental crust
.
Earth and Planetary Science Letters
 ,
337
,
211
223
,
247.
Guitreau
M.
,
Mora
N.
and
Paquette
J.
2018
.
Crystallization and disturbance histories of single zircon crystals from Hadean–Eoarchean Acasta gneisses examined by LA-ICP-MS U–Pb traverses
.
Geochemistry, Geophysics, Geosystems
 ,
19
,
272
291
,
248.
Guitreau
M.
,
Boyet
M.
et al
2019
.
Hadean protocrust reworking at the origin of the Archean Napier Complex (Antarctica)
.
Geochemical Perspectives Letters
 ,
12
,
7
11
,
249.
Guitreau
M.
,
Gannoun
A.
,
Deng
Z.
,
Marin-Carbonne
J.
,
Chaussidon
M.
and
Moynier
F.
2020
.
Silicon isotope measurement in zircon by laser ablation multiple collector inductively coupled plasma mass spectrometry
.
Journal of Analytical Atomic Spectrometry
 ,
35
,
1597
1606
,
250.
Guitreau
M.
,
Gannoun
A.
,
Deng
Z.
,
Chaussidon
M.
,
Moynier
F.
,
Barbarin
B.
and
Marin-Carbonne
J.
2022
.
Stable isotope geochemistry of silicon in granitoid zircon
.
Geochimica et Cosmochimica Acta
 ,
316
,
273
294
,
251.
Guo
J.-L.
,
Wang
Z.
,
Zhang
W.
,
Moynier
F.
,
Cui
D.
,
Hu
Z.
and
Ducea
M.N.
2020
.
Significant Zr isotope variations in single zircon grains recording magma evolution history
.
Proceedings of the National Academy of Sciences of the USA
 ,
117
,
21125
21131
,
252.
Gutieva
L.
,
Dziggel
A.
,
Volante
S.
and
Johnson
T.
2021
.
Zircon U–Pb and Lu–Hf record from the Archean Lewisian Gneiss Complex, NW Scotland
.
Goldschmidt Conference 2021
,
4–9 July 2021
,
253.
Hacker
B.R.
,
Kylander-Clark
A.R.
,
Holder
R.
,
Andersen
T.B.
,
Peterman
E.M.
,
Walsh
E.O.
and
Munnikhuis
J.K.
2015
.
Monazite response to ultrahigh-pressure subduction from U–Pb dating by laser ablation split stream
.
Chemical Geology
 ,
409
,
28
41
,
254.
Hacker
B.
,
Kylander-Clark
A.
and
Holder
R.
2019
.
REE partitioning between monazite and garnet: implications for petrochronology
.
Journal of Metamorphic Geology
 ,
37
,
227
237
,
255.
Hagen-Peter
G.
,
Cottle
J.M.
,
Smit
M.
and
Cooper
A.F.
2016
.
Coupled garnet Lu–Hf and monazite U–Pb geochronology constrain early convergent margin dynamics in the Ross orogen, Antarctica
.
Journal of Metamorphic Geology
 ,
34
,
293
319
,
256.
Hallett
B.W.
and
Spear
F.S.
2015
.
Monazite, zircon, and garnet growth in migmatitic pelites as a record of metamorphism and partial melting in the East Humboldt Range, Nevada
.
American Mineralogist
 ,
100
,
951
972
,
257.
Hammerli
J.
and
Kemp
T.I.S.
2021
.
Combined Hf and Nd isotope microanalysis of co-existing zircon and REE-rich accessory minerals: high resolution insights into crustal processes
.
Chemical Geology
 ,
581
, article
120393
,
258.
Hammerli
J.
,
Kemp
A.I.S.
and
Spandler
C.
2014
.
Neodymium isotope equilibration during crustal metamorphism revealed by in situ microanalysis of REE-rich accessory minerals
.
Earth and Planetary Science Letters
 ,
392
,
133
142
,
259.
Hammerli
J.
,
Kemp
A.I.S.
and
Whitehouse
M.J.
2019
.
In situ trace element and Sm–Nd isotope analysis of accessory minerals in an Eoarchean tonalitic gneiss from Greenland: implications for Hf and Nd isotope decoupling in Earth's ancient rocks
.
Chemical Geology
 ,
524
,
394
405
,
260.
Hanchar
J.M.
and
Miller
C.F.
1993
.
Zircon zonation patterns as revealed by cathodoluminescence and backscattered electron images: implications for interpretation of complex crustal histories
.
Chemical Geology
 ,
110
,
1
13
,
261.
Harley
S.L.
2008
.
Refining the P–T records of UHT crustal metamorphism
.
Journal of Metamorphic Geology
 ,
26
,
125
154
,
262.
Harley
S.L.
2016
.
A matter of time: the importance of the duration of UHT metamorphism
.
Journal of Mineralogical and Petrological Sciences
 ,
111
,
50
72
,
263.
Harley
S.L.
and
Kelly
N.M.
2007
.
The impact of zircon–garnet REE distribution data on the interpretation of zircon U–Pb ages in complex high-grade terrains: an example from the Rauer Islands, East Antarctica
.
Chemical Geology
 ,
241
,
62
87
,
264.
Harley
S.L.
and
Nandakumar
V.
2014
.
Accessory mineral behaviour in granulite migmatites: a case study from the Kerala Khondalite Belt, India
.
Journal of Petrology
 ,
55
,
1965
2002
,
265.
Harley
S.L.
,
Kinny
P.
,
Snape
I.
and
Black
L.P.
2001
.
Zircon chemistry and the definition of events in Archaean granulite terrains
.
Extended Abstracts of the 4th International Archaean Symposium
,
24–28 September
,
Perth, Australia
. AGSO-Geoscience Australia,
511
513
.
266.
Harlov
D.E.
2015
.
Apatite: a fingerprint for metasomatic processes
.
Elements
 ,
11
,
171
176
,
267.
Harlov
D.E.
and
Hetherington
C.J.
2010
.
Partial high-grade alteration of monazite using alkali-bearing fluids: experiment and nature
.
American Mineralogist
 ,
95
,
1105
1108
,
268.
Harlov
D.E.
,
Wirth
R.
and
Hetherington
C.J.
2011
.
Fluid-mediated partial alteration in monazite: the role of coupled dissolution–reprecipitation in element redistribution and mass transfer
.
Contributions to Mineralogy and Petrology
 ,
162
,
329
348
,
269.
Harris
C.
and
Vogeli
J.
2010
.
Oxygen isotope composition of garnet in the Peninsula Granite, Cape Granite Suite, South Africa: constraints on melting and emplacement mechanisms
.
South African Journal of Geology
 ,
113
,
401
412
,
270.
Harrison
T.M.
and
Watson
E.B.
1984
.
The behavior of apatite during crustal anatexis: equilibrium and kinetic considerations
.
Geochimica et Cosmochimica Acta
 ,
48
,
1467
1477
,
271.
Hart
E.
,
Storey
C.
,
Harley
S.L.
and
Fowler
M.
2018
.
A window into the lower crust: trace element systematics and the occurrence of inclusions/intergrowths in granulite-facies rutile
.
Gondwana Research
 ,
59
,
76
86
,
272.
Hartnady
M.I.H.
,
Kirkland
C.L.
,
Clark
C.
,
Spaggiari
C.V.
,
Smithies
R.H.
,
Evans
N.J.
and
McDonald
B.J.
2019
.
Titanite dates crystallization: slow Pb diffusion during super-solidus re-equilibration
.
Journal of Metamorphic Geology
 ,
37
,
823
838
,
273.
Hayden
L.A.
,
Watson
E.B.
and
Wark
D.A.
2008
.
A thermobarometer for sphene (titanite)
.
Contributions to Mineralogy and Petrology
 ,
155
,
529
540
,
274.
Heinrich
W.
,
Rehs
G.
and
Franz
G.
1997
.
Monazite–xenotime miscibility gap thermometry. I. An empirical calibration
.
Journal of Metamorphic Geology
 ,
15
,
3
16
,
275.
Henrichs
I.A.
,
O'Sullivan
G.
,
Chew
D.M.
,
Mark
C.
,
Babechuk
M.G.
,
McKenna
C.
and
Emo
R.
2018
.
The trace element and U–Pb systematics of metamorphic apatite
.
Chemical Geology
 ,
483
,
218
238
,
276.
Henrichs
I.A.
,
Chew
D.M.
,
O'Sullivan
G.J.
,
Mark
C.
,
McKenna
C.
and
Guyett
P.
2019
.
Trace element (Mn–Sr–Y–Th–REE) and U–Pb isotope systematics of metapelitic apatite during progressive greenschist- to amphibolite-facies Barrovian metamorphism
.
Geochemistry, Geophysics, Geosystems
 ,
20
,
4103
4129
,
277.
Hermann
J.
2002
.
Allanite: thorium and light rare earth element carrier in subducted crust
.
Chemical Geology
 ,
192
,
289
306
,
278.
Hermann
J.
and
Rubatto
D.
2003
.
Relating zircon and monazite domains to garnet growth zones: age and duration of granulite facies metamorphism in the Val Malenco lower crust
.
Journal of Metamorphic Geology
 ,
21
,
833
852
,
279.
Hermann
J.
and
Rubatto
D.
2009
.
Accessory phase control on the trace element signature of sediment melts in subduction zones
.
Chemical Geology
 ,
265
,
512
526
,
280.
Hermann
J.
and
Rubatto
D.
2014
. Subduction of continental crust to mantle depth: geochemistry of ultrahigh-pressure rocks. In:
Holland
H.D.
and
Turekian
K.K.
(eds)
Treatise on Geochemistry
 . 2nd edn,
309
340
, .
281.
Hoffmann
J.E.
,
Münker
C.
,
Polat
A.
,
Rosing
M.T.
and
Schulz
T.
2011
.
The origin of decoupled Hf–Nd isotope compositions in Eoarchean rocks from southern West Greenland
.
Geochimica et Cosmochimica Acta
 ,
75
,
6610
6628
,
282.
Hofmann
A.E.
,
Baker
M.B.
and
Eiler
J.M.
2013
.
An experimental study of Ti and Zr partitioning among zircon, rutile, and granitic melt
.
Contributions to Mineralogy and Petrology
 ,
166
,
235
253
,
283.
Högdahl
K.
,
Majka
J.
,
Sjöström
H.
,
Nilsson
K.P.
,
Claesson
S.
and
Konečný
P.
2012
.
Reactive monazite and robust zircon growth in diatexites and leucogranites from a hot, slowly cooled orogen: implications for the Palaeoproterozoic tectonic evolution of the central Fennoscandian Shield, Sweden
.
Contributions to Mineralogy and Petrology
 ,
163
,
167
188
,
284.
Hokada
T.
and
Harley
S.L.
2004
.
Zircon growth in UHT leucosome: constraints from zircon-garnet rare earth elements (REE) relations in Napier Complex, East Antarctica
.
Journal of Mineralogical and Petrological Sciences
 ,
99
,
180
190
,
285.
Holdaway
M.
2001
.
Recalibration of the GASP geobarometer in light of recent garnet and plagioclase activity models and versions of the garnet-biotite geothermometer
.
American Mineralogist
 ,
86
,
1117
1129
,
286.
Holder
R.M.
and
Hacker
B.R.
2019
.
Fluid-driven resetting of titanite following ultrahigh-temperature metamorphism in southern Madagascar
.
Chemical Geology
 ,
504
,
38
52
,
287.
Holder
R.M.
,
Hacker
B.R.
and
Kylander-Clark
A.R.
2013
.
Monazite petrochronology from the UHP Western Gneiss region, Norway
. American Geophysical Union Fall Meeting, 9–13 December 2013, San Francisco, CA, Abstracts,
2013AGUFM.V23A2759H
 .
288.
Holder
R.M.
,
Hacker
B.R.
,
Seward
G.G.E.
and
Kylander-Clark
A.R.C.
2019
.
Interpreting titanite U–Pb dates and Zr thermobarometry in high-grade rocks: empirical constraints on elemental diffusivities of Pb, Al, Fe, Zr, Nb, and Ce
.
Contributions to Mineralogy and Petrology
 ,
174
, article
42
,
289.
Holder
R.M.
,
Yakymchuk
C.
and
Viete
D.R.
2020
.
Accessory mineral Eu anomalies in suprasolidus rocks: beyond feldspar
.
Geochemistry, Geophysics, Geosystems
 ,
21
, article
e2020GC009052
,
290.
Holland
H.D.
and
Gottfried
D.
1955
.
The effect of nuclear radiation on the structure of zircon
.
Acta Crystallographica
 ,
8
,
291
300
,
291.
Holland
T.J.B.
and
Powell
R.
1990
.
An enlarged and updated internally consistent thermodynamic dataset with uncertainties and correlations: the system K2O–Na2O–CaO–MgO–MnO–FeO–Fe2O3–Al2O3–TiO2–SiO2–C–H2–O2
.
Journal of Metamorphic Geology
 ,
8
,
89
124
,
292.
Holland
T.J.B.
and
Powell
R.
1998
.
An internally consistent thermodynamic data set for phases of petrological interest
.
Journal of Metamorphic Geology
 ,
16
,
309
344
,
293.
Holland
T.J.B.
and
Powell
R.
2011
.
An improved and extended internally consistent thermodynamic dataset for phases of petrological interest, involving a new equation of state for solids
.
Journal of Metamorphic Geology
 ,
29
,
333
383
,
294.
Hollister
L.S.
1966
.
Garnet zoning: an interpretation based on the Rayleigh fractionation model
.
Science
 ,
154
,
1647
1651
,
295.
Holtmann
R.
,
Muñoz-Montecinos
J.
et al
2022
.
Cretaceous thermal evolution of the closing Neo-Tethyan realm revealed by multi-method petrochronology
.
Lithos
 ,
422–423
,
106731
,
296.
Hoover
W.F.
,
Penniston-Dorland
S.C.
,
Baumgartner
L.P.
,
Bouvier
A.-S.
,
Baker
D.
,
Dragovic
B.
and
Gion
A.
2021
.
A method for secondary ion mass spectrometry measurement of lithium isotopes in garnet: the utility of glass reference materials
.
Geostandards and Geoanalytical Research
 ,
45
,
477
499
,
297.
Hoover
W.F.
,
Penniston-Dorland
S.
et al
2022
.
Episodic fluid flow in an eclogite-facies shear zone: insights from Li isotope zoning in garnet
.
Geology
 ,
50
,
746
750
,
298.
Hopkins
M.
,
Harrison
T.M.
and
Manning
C.E.
2008
.
Low heat flow inferred from >4 Gyr zircons suggests Hadean plate boundary interactions
.
Nature
 ,
456
,
493
496
,
299.
Horton
F.
,
Holder
R.M.
and
Swindle
C.R.
2022
.
An extensive record of orogenesis recorded in a Madagascar granulite
.
Journal of Metamorphic Geology
 ,
40
,
287
305
, .
300.
Hoschek
G.
2016
.
Phase relations of the REE minerals florencite, allanite and monazite in quartzitic garnet–kyanite schist of the Eclogite Zone, Tauern Window, Austria
.
European Journal of Mineralogy
 ,
28
,
735
750
.
301.
Hoshino
M.
,
Kimata
M.
,
Arakawa
Y.
,
Shimizu
M.
,
Nishida
N.
and
Nakai
S.
2007
.
Allanite-(Ce) as an indicator of the origin of granitic rocks in Japan: importance of Sr–Nd isotopic and chemical composition
.
The Canadian Mineralogist
 ,
45
,
1329
1336
,
302.
Hoskin
P.W.O.
and
Black
L.P.
2000
.
Metamorphic zircon formation by solid-state recrystallization of protolith igneous zircon
.
Journal of Metamorphic Geology
 ,
18
,
423
439
,
303.
Hoskin
P.W.O.
and
Schaltegger
U.
2003
.
The composition of zircon and igneous and metamorphic petrogenesis
.
Reviews in Mineralogy and Geochemistry
 ,
53
,
27
62
,
304.
Humphries
F.J.
and
Cliff
R.A.
1982
.
Sm–Nd dating and cooling history of Scourian granulites, Sutherland
.
Nature
 ,
295
,
515
517
,
305.
Ibañez-Mejia
M.
and
Tissot
F.L.
2019
.
Extreme Zr stable isotope fractionation during magmatic fractional crystallization
.
Science Advances
 ,
5
, article
eaax8648
,
306.
Ihlen
P.M.
,
Schiellerup
H.
,
Gautneb
H.
and
Skår
Ø.
2014
.
Characterization of apatite resources in Norway and their REE potential – a review
.
Ore Geology Reviews
 ,
58
,
126
147
,
307.
Iizuka
T.
,
Hirata
T.
,
Komiya
T.
,
Rino
S.
,
Katayama
I.
,
Motoki
A.
and
Maruyama
S.
2005
.
U–Pb and Lu–Hf isotope systematics of zircons from the Mississippi River sand: implications for reworking and growth of continental crust
.
Geology
 ,
33
,
485
488
,
308.
Iizuka
T.
,
Campbell
I.H.
,
Allen
C.M.
,
Gill
J.B.
,
Maruyama
S.
and
Makoka
F.
2013
.
Evolution of the African continental crust as recorded by U–Pb, Lu–Hf and O isotopes in detrital zircons from modern rivers
.
Geochimica et Cosmochimica Acta
 ,
107
,
96
120
,
309.
Iizuka
T.
,
Yamaguchi
T.
,
Hibiya
Y.
and
Amelin
Y.
2015
.
Meteorite zircon constraints on the bulk Lu−Hf isotope composition and early differentiation of the Earth
.
Proceedings of the National Academy of Sciences of the USA
 ,
112
,
5331
5336
,
310.
Itano
K.
,
Ueki
K.
,
Iizuka
T.
and
Kuwatani
T.
2020
.
Geochemical discrimination of monazite source rock based on machine learning techniques and multinomial logistic regression analysis
.
Geosciences
 ,
10
, article
63
,
311.
Jaffey
A.
,
Flynn
K.
,
Glendenin
L.
,
Bentley
W.T.
and
Essling
A.
1971
.
Precision measurement of half-lives and specific activities of 235U and 238U
.
Physical Review C
 ,
4
, article
1889
,
312.
Jamieson
J.C.
and
Olinger
B.
1969
.
Pressure–temperature studies of anatase, brookite rutile, and Ti02(II): a discussion
.
American Mineralogist
 ,
54
,
1477
1481
.
313.
Janots
E.
and
Rubatto
D.
2014
.
U–Th–Pb dating of collision in the external Alpine domains (Urseren zone, Switzerland) using low temperature allanite and monazite
.
Lithos
 ,
184–187
,
155
166
,
314.
Janots
E.
,
Negro
F.
,
Brunet
F.
,
Goffé
B.
,
Engi
M.
and
Bouybaouène
M.L.
2006
.
Evolution of the REE mineralogy in HP–LT metapelites of the Sebtide Complex, Rif, Morocco: monazite stability and geochronology
.
Lithos
 ,
87
,
214
234
,
315.
Janots
E.
,
Brunet
F.
,
Goffé
B.
,
Poinssot
C.
,
Burchard
M.
and
Cemič
L.
2007
.
Thermochemistry of monazite-(La) and dissakisite-(La): implications for monazite and allanite stability in metapelites
.
Contributions to Mineralogy and Petrology
 ,
154
,
1
14
,
316.
Janots
E.
,
Engi
M.
,
Berger
A.
,
Allaz
J.
,
Schwarz
J.
and
Spandler
C.
2008
.
Prograde metamorphic sequence of REE minerals in pelitic rocks of the Central Alps: implications for allanite–monazite–xenotime phase relations from 250 to 610°C
.
Journal of Metamorphic Geology
 ,
26
,
509
526
,
317.
Janots
E.
,
Engi
M.
,
Rubatto
D.
,
Berger
A.
,
Gregory
C.
and
Rahn
M.
2009
.
Metamorphic rates in collisional orogeny from in situ allanite and monazite dating
.
Geology
 ,
37
,
11
14
,
318.
Janots
E.
,
Berger
A.
and
Engi
M.
2011
.
Physico-chemical control on the REE minerals in chloritoid-grade metasediments from a single outcrop (Central Alps, Switzerland)
.
Lithos
 ,
121
,
1
11
,
319.
Jayananda
M.
,
Chardon
D.
,
Peucat
J.-J.
and
Fanning
C.M.
2015
.
Paleo- to Mesoarchean TTG accretion and continental growth in the western Dharwar craton, southern India: constraints from SHRIMP U–Pb zircon geochronology, whole-rock geochemistry and Nd–Sr isotopes
.
Precambrian Research
 ,
268
,
295
322
,
320.
Jennings
E.S.
,
Marschall
H.
,
Hawkesworth
C.
and
Storey
C.
2011
.
Characterization of magma from inclusions in zircon: apatite and biotite work well, feldspar less so
.
Geology
 ,
39
,
863
866
,
321.
Jiao
S.
,
Fitzsimons
I.C.W.
,
Zi
J.-W.
,
Evans
N.J.
,
Mcdonald
B.J.
and
Guo
J.
2020a
.
Texturally controlled U–Th–Pb monazite geochronology reveals Paleoproterozoic UHT metamorphic evolution in the Khondalite Belt, North China Craton
.
Journal of Petrology
 ,
61
, article
egaa023
,
322.
Jiao
S.
,
Guo
J.
,
Evans
N.J.
,
Mcdonald
B.J.
,
Liu
P.
,
Ouyang
D.
and
Fitzsimons
I.C.W.
2020b
.
The timing and duration of high-temperature to ultrahigh-temperature metamorphism constrained by zircon U–Pb–Hf and trace element signatures in the Khondalite Belt, North China Craton
.
Contributions to Mineralogy and Petrology
 ,
175
, article
66
,
323.
Jiao
S.
,
Evans
N.J.
,
Mitchell
R.N.
,
Fitzsimons
I.C.W.
and
Guo
J.
2021
.
Heavy rare-earth element and Y partitioning between monazite and garnet in aluminous granulites
.
Contributions to Mineralogy and Petrology
 ,
176
, article
50
,
324.
John
T.
,
Klemd
R.
,
Klemme
S.
,
Pfänder
J.A.
,
Elis Hoffmann
J.
and
Gao
J.
2011
.
Nb–Ta fractionation by partial melting at the titanite–rutile transition
.
Contributions to Mineralogy and Petrology
 ,
161
,
35
45
,
325.
Johnson
S.E.
1999
.
Porphyroblast microstructures: a review of current and future trends
.
American Mineralogist
 ,
84
,
1711
1726
,
326.
Johnson
T.A.
,
Vervoort
J.D.
,
Ramsey
M.J.
,
Aleinikoff
J.N.
and
Southworth
S.
2018
.
Constraints on the timing and duration of orogenic events by combined Lu–Hf and Sm–Nd geochronology: an example from the Grenville orogeny
.
Earth and Planetary Science Letters
 ,
501
,
152
164
,
327.
Jonnalagadda
M.K.
,
Karmalkar
N.R.
,
Duraiswami
R.A.
,
Harshe
S.
,
Gain
S.
and
Griffin
W.L.
2017
.
Formation of atoll garnets in the UHP eclogites of the Tso Morari Complex, Ladakh, Himalaya
.
Journal of Earth System Science
 ,
126
, article
107
,
328.
Joseph
C.
,
Fougerouse
D.
,
Saxey
D.W.
,
Verberne
R.
,
Reddy
S.M.
and
Rickard
W.D.A.
2021
.
Xenotime at the nanoscale: U–Pb geochronology and optimisation of analyses by atom probe tomography
.
Geostandards and Geoanalytical Research
 ,
45
,
443
456
,
329.
Kamber
B.S.
,
Frei
R.
and
Gibb
A.J.
1998
.
Pitfalls and new approaches in granulite chronometry: an example from the Limpopo Belt, Zimbabwe
.
Precambrian Research
 ,
91
,
269
285
,
330.
Kamenetsky
V.S.
,
Doroshkevich
A.G.
,
Elliott
H.A.L.
and
Zaitsev
A.N.
2021
.
Carbonatites: contrasting, complex, and controversial
.
Elements
 ,
17
,
307
314
,
331.
Kapp
P.
,
Manning
C.E.
and
Tropper
P.
2009
.
Phase-equilibrium constraints on titanite and rutile activities in mafic epidote amphibolites and geobarometry using titanite–rutile equilibria
.
Journal of Metamorphic Geology
 ,
27
,
509
521
,
332.
Karioris
F.
,
Gowda
K.A.
and
Cartz
L.
1981
.
Heavy ion bombardment of monoclinic ThSiO4, ThO2 and monazite
.
Radiation Effects
 ,
58
,
1
3
,
333.
Kelly
N.M.
and
Harley
S.L.
2005
.
An integrated microtextural and chemical approach to zircon geochronology: refining the Archaean history of the Napier Complex, east Antarctica
.
Contributions to Mineralogy and Petrology
 ,
149
,
57
84
,
334.
Kelly
N.M.
,
Harley
S.L.
and
Möller
A.
2012
.
Complexity in the behavior and recrystallization of monazite during high-T metamorphism and fluid infiltration
.
Chemical Geology
 ,
322
,
192
208
,
335.
Kelsey
D.E.
and
Powell
R.
2011
.
Progress in linking accessory mineral growth and breakdown to major mineral evolution in metamorphic rocks: A thermodynamic approach in the Na2O-CaO-K2O-FeO-MgO-Al2O3-SiO2-H2O-TiO2-ZrO2 system
.
Journal of Metamorphic Geology
 ,
29
,
151
166
.
336.
Kelsey
D.E.
,
Hand
M.
,
Clark
C.
and
Wilson
C.J.L.
2007
.
On the application of in situ monazite chemical geochronology to constraining P–T–t histories in high-temperature (>850°C) polymetamorphic granulites from Prydz Bay, East Antarctica
.
Journal of the Geological Society, London
 ,
164
,
667
683
,
337.
Kelsey
D.E.
,
Clark
C.
and
Hand
M.
2008
.
Thermobarometric modelling of zircon and monazite growth in melt-bearing systems: examples using model metapelitic and metapsammitic granulites
.
Journal of Metamorphic Geology
 ,
26
,
199
212
,
338.
Kemp
A.
,
Hawkesworth
C.
et al
2007
.
Magmatic and crustal differentiation history of granitic rocks from Hf–O isotopes in zircon
.
Science
 ,
315
,
980
983
,
339.
Kennedy
A.K.
,
Wotzlaw
J.-F.
et al
2022
.
Apatite reference materials for SIMS microanalysis of isotopes and trace elements
.
Geostandards and Geoanalytical Research
 ,
47
,
215
467
,
340.
King
E.M.
and
Valley
J.W.
2001
.
The source, magmatic contamination, and alteration of the Idaho wbatholith
.
Contributions to Mineralogy and Petrology
 ,
142
,
72
88
,
341.
King
E.M.
,
Valley
J.W.
,
Davis
D.W.
and
Kowallis
B.J.
2001
.
Empirical determination of oxygen isotope fractionation factors for titanite with respect to zircon and quartz
.
Geochimica et Cosmochimica Acta
 ,
65
,
3165
3175
,
342.
King
P.L.
,
Sham
T.-K.
,
Gordon
R.A.
and
Dyar
M.D.
2013
.
Microbeam X-ray analysis of Ce3+/Ce4+ in Ti-rich minerals: a case study with titanite (sphene) with implications for multivalent trace element substitution in minerals
.
American Mineralogist
 ,
98
,
110
119
,
343.
Kingsbury
J.A.
,
Miller
C.F.
,
Wooden
J.L.
and
Harrison
T.M.
1993
.
Monazite paragenesis and U–Pb systematics in rocks of the eastern Mojave Desert, California, USA: implications for thermochronometry
.
Chemical Geology
 ,
110
,
147
167
,
344.
Kirkland
C.L.
,
Smithies
R.H.
,
Taylor
R.J.M.
,
Evans
N.
and
McDonald
B.
2015
.
Zircon Th/U ratios in magmatic environs
.
Lithos
 ,
212
,
397
414
,
345.
Kirkland
C.L.
,
Spaggiari
C.V.
et al
2016
.
Grain size matters: Implications for element and isotopic mobility in titanite
.
Precambrian Research
 ,
278
,
283
302
, .
346.
Kirkland
C.L.
,
Hollis
J.
,
Danisik
M.
,
Petersen
J.
,
Evans
N.
and
Mcdonald
B.
2017
.
Apatite and titanite from the Karrat Group, Greenland; implications for charting the thermal evolution of crust from the U–Pb geochronology of common Pb bearing phases
.
Precambrian Research
 ,
300
, 107–120,
347.
Kirkland
C.L.
,
Fougerouse
D.
,
Reddy
S.M.
,
Hollis
J.
and
Saxey
D.W.
2018
.
Assessing the mechanisms of common Pb incorporation into titanite
.
Chemical Geology
 ,
483
,
558
566
,
348.
Kirkland
C.L.
,
Yakymchuk
C.
,
Gardiner
N.J.
,
Szilas
K.
,
Hollis
J.
,
Olierook
H.
and
Steenfelt
A.
2020
.
Titanite petrochronology linked to phase equilibrium modelling constrains tectono-thermal events in the Akia Terrane, West Greenland
.
Chemical Geology
 ,
536
, article,
119467
,
349.
Kirkland
C.L.
,
Hartnady
M.I.H.
,
Barham
M.
,
Olierook
H.K.H.
,
Steenfelt
A.
and
Hollis
J.A.
2021
.
Widespread reworking of Hadean-to-Eoarchean continents during Earth's thermal peak
.
Nature Communications
 ,
12
, article
331
,
350.
Klemme
S.
,
Prowatke
S.
,
Hametner
K.
and
Günther
D.
2005
.
Partitioning of trace elements between rutile and silicate melts: implications for subduction zones
.
Geochimica et Cosmochimica Acta
 ,
69
,
2361
2371
,
351.
Kohn
M.J.
2017
.
Titanite petrochronology
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
419
441
,
352.
Kohn
M.J.
2020
.
A refined zirconium-in-rutile thermometer
.
American Mineralogist: Journal of Earth and Planetary Materials
 ,
105
,
963
971
,
353.
Kohn
M.J.
and
Corrie
S.L.
2011
.
Preserved Zr-temperatures and U–Pb ages in high-grade metamorphic titanite: evidence for a static hot channel in the Himalayan orogen
.
Earth and Planetary Science Letters
 ,
311
,
136
143
,
354.
Kohn
M.J.
and
Valley
J.W.
1998
.
Effects of cation substitutions in garnet and pyroxene on equilibrium oxygen isotope fractionations
.
Journal of Metamorphic Geology
 ,
16
,
625
639
,
355.
Kohn
M.J.
,
Penniston-Dorland
S.C.
and
Ferreira
J.C.S.
2016
.
Implications of near-rim compositional zoning in rutile for geothermometry, geospeedometry, and trace element equilibration
.
Contributions to Mineralogy and Petrology
 ,
171
, article
78
,
356.
Kohn
M.J.
,
Engi
M.
and
Lanari
P.
2017
.
Petrochronology: methods and applications
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
357.
Konečný
P.
,
Kusiak
M.A.
and
Dunkley
D.J.
2018
.
Improving U–Th–Pb electron microprobe dating using monazite age references
.
Chemical Geology
 ,
484
,
22
35
,
358.
Kooijman
E.
,
Mezger
K.
and
Berndt
J.
2010
.
Constraints on the U–Pb systematics of metamorphic rutile from in situ LA-ICP-MS analysis
.
Earth and Planetary Science Letters
 ,
293
,
321
330
,
359.
Kooijman
E.
,
Smit
M.A.
,
Mezger
K.
and
Berndt
J.
2012
.
Trace element systematics in granulite facies rutile: implications for Zr geothermometry and provenance studies
.
Journal of Metamorphic Geology
 ,
30
,
397
412
,
360.
Košler
J.
,
Wiedenbeck
M.
,
Wirth
R.
,
Hovorka
J.
,
Sylvester
P.
and
Míková
J.
2005
.
Chemical and phase composition of particles produced by laser ablation of silicate glass and zircon–implications for elemental fractionation during ICP-MS analysis
.
Journal of Analytical Atomic Spectrometry
 ,
20
,
402
409
,
361.
Krenn
E.
and
Finger
F.
2010
.
Unusually Y-rich monazite-(Ce) with 6–14 wt.% Y2O3 in a granulite from the Bohemian Massif: implications for high-temperature monazite growth from the monazite–xenotime miscibility gap thermometry
.
Mineralogical Magazine
 ,
74
,
217
225
,
362.
Krenn
E.
,
Schulz
B.
and
Finger
F.
2012
.
Three generations of monazite in Austroalpine basement rocks to the south of the Tauern Window: evidence for Variscan, Permian and Eo-Alpine metamorphic events
.
Swiss Journal of Geosciences
 ,
105
,
343
360
,
363.
Krneta
S.
,
Ciobanu
C.L.
,
Cook
N.J.
and
Ehrig
K.J.
2018
.
Numerical modeling of REE fractionation patterns in fluorapatite from the Olympic Dam Deposit (South Australia)
.
Minerals
 ,
8
, article
342
,
364.
Krogstad
E.J.
and
Walker
R.J.
1994
.
High closure temperatures of the U–Pb system in large apatites from the Tin Mountain pegmatite, Black Hills, South Dakota, USA
.
Geochimica et Cosmochimica Acta
 ,
58
,
3845
3853
,
365.
Kröner
A.
,
Wan
Y.
,
Liu
X.
and
Liu
D.
2014
.
Dating of zircon from high-grade rocks: which is the most reliable method?
Geoscience Frontiers
 ,
5
,
515
523
,
366.
Kulhánek
J.
,
Faryad
S.W.
,
Jedlicka
R.
and
Svojtka
M.
2021
.
Dissolution and reprecipitation of garnet during eclogite-facies metamorphism; major and trace element transfer during atoll garnet formation
.
Journal of Petrology
 ,
62
, article
egab077
,
367.
Kunz
B.E.
,
Regis
D.
and
Engi
M.
2018
.
Zircon ages in granulite facies rocks: decoupling from geochemistry above 850°C?
Contributions to Mineralogy and Petrology
 ,
173
, article
26
,
368.
Kusiak
M.A.
,
Whitehouse
M.J.
,
Wilde
S.A.
,
Nemchin
A.A.
and
Clark
C.
2013
.
Mobilization of radiogenic Pb in zircon revealed by ion imaging: implications for early Earth geochronology
.
Geology
 ,
41
,
291
294
,
369.
Kylander-Clark
A.R.C.
,
Hacker
B.R.
and
Mattinson
J.M.
2008
.
Slow exhumation of UHP terranes: titanite and rutile ages of the Western Gneiss Region, Norway
.
Earth and Planetary Science Letters
 ,
272
,
531
540
,
370.
Kylander-Clark
A.R.C.
,
Hacker
B.R.
and
Cottle
J.M.
2013
.
Laser-ablation split-stream ICP petrochronology
.
Chemical Geology
 ,
345
,
99
112
,
371.
Lackey
J.S.
,
Valley
J.W.
and
Hinke
H.J.
2006
.
Deciphering the source and contamination history of peraluminous magmas using δ18O of accessory minerals: examples from garnet-bearing plutons of the Sierra Nevada batholith
.
Contributions to Mineralogy and Petrology
 ,
151
,
20
44
,
372.
Lana
C.
,
Gonçalves
G.O.
et al
2022
.
Assessing the U–Pb, Sm–Nd and Sr–Sr isotopic compositions of the Sumé apatite as a reference material for LA-ICP-MS analysis
.
Geostandards and Geoanalytical Research
 ,
46
,
71
95
,
373.
Lanari
P.
and
Engi
M.
2017
.
Local bulk composition effects on metamorphic mineral assemblages
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
55
102
,
374.
Lanari
P
and
Piccoli
F
2020
.
IOP Conference Series: Materials Science and Engineering
 ,
891
,
012016
,
375.
Lanari
P.
,
Vidal
O.
,
De Andrade
V.
,
Dubacq
B.
,
Lewin
E.
,
Grosch
E.G.
and
Schwartz
S.
2014
.
XMapTools: a MATLAB©-based program for electron microprobe X-ray image processing and geothermobarometry
.
Computers & Geosciences
 ,
62
,
227
240
,
376.
Lanari
P.
,
Giuntoli
F.
,
Loury
C.
,
Burn
M.
and
Engi
M.
2017
.
An inverse modeling approach to obtain P–T conditions of metamorphic stages involving garnet growth and resorption
.
European Journal of Mineralogy
 ,
29
,
181
199
,
377.
Lanari
P.
,
Ferrero
S.
,
Goncalves
P.
and
Grosch
E.G.
2019
.
Metamorphic geology: progress and perspectives
.
Geological Society, London, Special Publications
 ,
478
,
1
12
,
378.
Larson
K.P.
,
Ali
A.
,
Shrestha
S.
,
Soret
M.
,
Cottle
J.M.
and
Ahmad
R.
2019
.
Timing of metamorphism and deformation in the Swat Valley, northern Pakistan: insight into garnet-monazite HREE partitioning
.
Geoscience Frontiers
 ,
10
,
849
861
,
379.
Larson
K.P.
,
Shrestha
S.
,
Cottle
J.M.
,
Guilmette
C.
,
Johnson
T.A.
,
Gibson
H.D.
and
Gervais
F.
2022
.
Re-evaluating monazite as a record of metamorphic reactions
.
Geoscience Frontiers
 ,
13
, article
101340
,
380.
Larsson
D.
and
Söderlund
U.
2005
.
Lu–Hf apatite geochronology of mafic cumulates: an example from a Fe–Ti mineralization at Smålands Taberg, southern Sweden
.
Chemical Geology
 ,
224
,
201
211
,
381.
Laurent
A.T.
,
Duchene
S.
,
Bingen
B.
,
Bosse
V.
and
Seydoux-Guillaume
A.-M.
2018
.
Two successive phases of ultrahigh temperature metamorphism in Rogaland, S. Norway: evidence from Y-in-monazite thermometry
.
Journal of Metamorphic Geology
 ,
36
,
1009
1037
,
382.
Laurent
O.
,
Zeh
A.
,
Gerdes
A.
,
Villaros
A.
,
Gros
K.
and
Słaby
E.
2017
.
How do granitoid magmas mix with each other? Insights from textures, trace element and Sr–Nd isotopic composition of apatite and titanite from the Matok pluton (South Africa)
.
Contributions to Mineralogy and Petrology
 ,
172
, article
80
,
383.
Laurent
O.
,
Moyen
J.-F.
,
Wotzlaw
J.-F.
,
Björnsen
J.
and
Bachmann
O.
2021
.
Early Earth zircons formed in residual granitic melts produced by tonalite differentiation
.
Geology
 ,
50
,
437
441
,
384.
Lawley
C.J.M.
,
Creaser
R.A.
et al
2015
.
Unraveling the western Churchill Province Paleoproterozoic gold metallotect: constraints from Re–Os arsenopyrite and U–Pb xenotime geochronology and LA-ICP-MS arsenopyrite trace element chemistry at the BIF-hosted Meliadine gold district, Nunavut, Canada
.
Economic Geology
 ,
110
,
1425
1454
,
385.
Lee
J.K.W.
,
Williams
I.S.
and
Ellis
D.J.
1997
.
Pb, U and Th diffusion in natural zircon
.
Nature
 ,
390
,
159
162
,
386.
Lei
H.
,
Xu
H.
and
Liu
P.
2020
.
Decoupling between Ti-in-zircon and Zr-in-rutile thermometry during ultrahigh temperature metamorphism of the Dabie Orogen, China
.
Geological Journal
 ,
55
,
6442
6449
,
387.
Le Roux
L.
and
Glendenin
L.
1963
.
Half-life of 232Th
. In:
Proceedings of the National Meeting on Nuclear Energy
,
Pretoria, South Africa
.
83
,
94
.
388.
Li
B.
,
Ge
J.
and
Zhang
B.
2018
.
Diffusion in garnet: a review
.
Acta Geochimica
 ,
37
,
19
31
,
389.
Li
D.
,
Fu
Y.
et al
2022
.
PL57 garnet as a new natural reference material for in situ U–Pb isotope analysis and its perspective for geological applications
.
Contributions to Mineralogy and Petrology
 ,
177
, article
19
,
390.
Li
R.
,
Collins
W.J.
,
Yang
J.-H.
,
Blereau
E.
and
Wang
H.
2021
.
Two-stage hybrid origin of Lachlan S-type magmas: a re-appraisal using isotopic microanalysis of lithic inclusion minerals
.
Lithos
 ,
402–403
, article
106378
,
391.
Li
W.-T.
,
Jiang
S.-Y.
,
Fu
B.
,
Liu
D.-L.
and
Xiong
S.-F.
2021
.
Zircon Hf–O isotope and magma oxidation state evidence for the origin of Early Cretaceous granitoids and porphyry Mo mineralization in the Tongbai–Hong'an–Dabie orogens, eastern China
.
Lithos
 ,
398
, article
106281
,
392.
Liao
X.
,
Li
Q.
,
Whitehouse
M.J.
,
Yang
Y.
and
Liu
Y.
2020
.
Allanite U–Th–Pb geochronology by ion microprobe
.
Journal of Analytical Atomic Spectrometry
 ,
35
,
489
497
,
393.
Liebmann
J.
,
Spencer
C.J.
,
Kirkland
C.L.
,
Xia
X.-P.
and
Bourdet
J.
2021
.
Effect of water on δ18O in zircon
.
Chemical Geology
 ,
574
, article
120243
,
394.
Lin
S.
,
Hu
K.
,
Cao
J.
,
Liu
Y.
,
Liu
S.
and
Zhang
B.
2023
.
Geochemistry and origin of hydrothermal apatite in Carlin-type Au deposits, southwestern China (Gaolong deposit)
.
Ore Geology Reviews
 ,
154
, article
105312
,
395.
Liu
Z.-C.
,
Wu
F.-Y.
,
Ji
W.-Q.
,
Wang
J.-G.
and
Liu
C.-Z.
2014
.
Petrogenesis of the Ramba leucogranite in the Tethyan Himalaya and constraints on the channel flow model
.
Lithos
 ,
208–209
,
118
136
,
396.
Loader
M.A.
,
Nathwani
C.L.
,
Wilkinson
J.J.
and
Armstrong
R.N.
2022
.
Controls on the magnitude of Ce anomalies in zircon
.
Geochimica et Cosmochimica Acta
 ,
328
,
242
257
,
397.
López-Moro
F.J.
,
Romer
R.
,
López-Plaza
M.
and
Sanchez
M.G.
2017
.
Zircon and allanite U–Pb ID-TIMS ages of vaugnerites from the Calzadilla pluton, Salamanca (Spain): dating mantle-derived magmatism and post-magmatic subsolidus overprint
.
Geologica Acta
 ,
15
,
395
408
,
398.
Z.
,
Zhang
L.
,
Du
J.
and
Bucher
K.
2008
.
Coesite inclusions in garnet from eclogitic rocks in western Tianshan, northwest China: convincing proof of UHP metamorphism
.
American Mineralogist
 ,
93
,
1845
1850
,
399.
Lucassen
F.
,
Franz
G.
,
Dulski
P.
,
Romer
R.L.
and
Rhede
D.
2011
.
Element and Sr isotope signatures of titanite as indicator of variable fluid composition in hydrated eclogite
.
Lithos
 ,
121
,
12
24
,
400.
Lucassen
F.
,
Franz
G.
and
Rhede
D.
2012
.
Small-scale transport of trace elements Nb and Cr during growth of titanite: an experimental study at 600°C, 0.4 GPa
.
Contributions to Mineralogy and Petrology
 ,
164
,
987
997
,
401.
Luvizotto
G.L.
and
Zack
T.
2009
.
Nb and Zr behavior in rutile during high-grade metamorphism and retrogression: an example from the Ivrea–Verbano Zone
.
Chemical Geology
 ,
261
,
303
317
,
402.
Luvizotto
G.L.
,
Zack
T.
et al
2009
.
Rutile crystals as potential trace element and isotope mineral standards for microanalysis
.
Chemical Geology
 ,
261
,
346
369
,
403.
Ma
Q.
,
Evans
N.J.
,
Ling
X.-X.
,
Yang
J.-H.
,
Wu
F.-Y.
,
Zhao
Z.-D.
and
Yang
Y.-H.
2019
.
Natural titanite reference materials for in situ U–Pb and Sm–Nd isotopic measurements by LA-(MC)-ICP-MS
.
Geostandards and Geoanalytical Research
 ,
43
,
355
384
,
404.
Machado
N.
and
Gauthier
G.
1996
.
Determination of 207Pb/206Pb ages on zircon and monazite by laser-ablation ICPMS and application to a study of sedimentary provenance and metamorphism in southeastern Brazil
.
Geochimica et Cosmochimica Acta
 ,
60
,
5063
5073
,
405.
Mahan
K.H.
,
Goncalves
P.
,
Williams
M.L.
and
Jercinovic
M.J.
2006
.
Dating metamorphic reactions and fluid flow: application to exhumation of high-P granulites in a crustal-scale shear zone, western Canadian Shield
.
Journal of Metamorphic Geology
 ,
24
,
193
217
,
406.
Maneiro
K.A.
,
Baxter
E.F.
,
Samson
S.D.
,
Marschall
H.R.
and
Hietpas
J.
2019
.
Detrital garnet geochronology: application in tributaries of the French Broad River, southern Appalachian Mountains, USA
.
Geology
 ,
47
,
1189
1192
,
407.
Manzotti
P.
,
Bosse
V.
,
Pitra
P.
,
Robyr
M.
,
Schiavi
F.
and
Ballèvre
M.
2018
.
Exhumation rates in the Gran Paradiso Massif (Western Alps) constrained by in situ U–Th–Pb dating of accessory phases (monazite, allanite and xenotime)
.
Contributions to Mineralogy and Petrology
 ,
173
, article
24
,
408.
Manzotti
P.
,
Schiavi
F.
,
Nosenzo
F.
,
Pitra
P.
and
Ballèvre
M.
2022
.
A journey towards the forbidden zone: a new, cold, UHP unit in the Dora-Maira Massif (Western Alps)
.
Contributions to Mineralogy and Petrology
 ,
177
,
1
22
,
409.
Marfin
A.E.
,
Ivanov
A.V.
,
Kamenetsky
V.S.
,
Abersteiner
A.
,
Yakich
T.Y.
and
Dudkin
T.V.
2020
.
Contact metamorphic and metasomatic processes at the Kharaelakh intrusion, Oktyabrsk deposit, Norilsk–Talnakh ore district: application of LA-ICP-MS dating of perovskite, apatite, garnet, and titanite
.
Economic Geology
 ,
115
,
1213
1226
,
410.
Mark
C.
,
Stutenbecker
L.
et al
2022
.
Detrital garnet Lu–Hf and U–Pb geochronometry coupled with compositional analysis: possibilities and limitations as a sediment provenance indicator
 . EGU General Assembly, Conference Abstracts,
23–27 May 2022, Vienna, Austria, EGU22-6405, https://doi.org/10.5194/egusphere-egu22-6405
.
411.
Mark
G.
2001
.
Nd isotope and petrogenetic constraints for the origin of the Mount Angelay igneous complex: implications for the origin of intrusions in the Cloncurry district, NE Australia
.
Precambrian Research
 ,
105
,
17
35
,
412.
Marmo
B.
,
Clarke
G.
and
Powell
R.
2002
.
Fractionation of bulk rock composition due to porphyroblast growth: effects on eclogite facies mineral equilibria, Pam Peninsula, New Caledonia
.
Journal of Metamorphic Geology
 ,
20
,
151
165
,
413.
Marsh
J.H.
and
Smye
A.J.
2017
.
U–Pb systematics and trace element characteristics in titanite from a high-pressure mafic granulite
.
Chemical Geology
 ,
466
,
403
416
,
414.
Martin
A.J.
,
Gehrels
G.E.
and
DeCelles
P.G.
2007
.
The tectonic significance of (U, Th)/Pb ages of monazite inclusions in garnet from the Himalaya of central Nepal
.
Chemical Geology
 ,
244
,
1
24
,
415.
Martin
E.L.
,
Collins
W.J.
and
Spencer
C.J.
2020
.
Laurentian origin of the Cuyania suspect terrane, western Argentina, confirmed by Hf isotopes in zircon
.
GSA Bulletin
 ,
132
,
273
290
,
416.
Martin
L.A.J.
,
Ballèvre
M.
,
Boulvais
P.
,
Halfpenny
A.
,
Vanderhaeghe
O.
,
Duchêne
S.
and
Deloule
E.
2011
.
Garnet re-equilibration by coupled dissolution–reprecipitation: evidence from textural, major element and oxygen isotope zoning of ‘cloudy’ garnet
.
Journal of Metamorphic Geology
 ,
29
,
213
231
,
417.
Martin
L.A.J.
,
Rubatto
D.
,
Crépisson
C.
,
Hermann
J.
,
Putlitz
B.
and
Vitale-Brovarone
A.
2014
.
Garnet oxygen analysis by SHRIMP-SI: matrix corrections and application to high-pressure metasomatic rocks from Alpine Corsica
.
Chemical Geology
 ,
374
,
25
36
,
418.
Massonne
H.-J.
and
Li
B.
2022
.
Eclogite with unusual atoll garnet from the southern Armorican Massif, France: pressure–temperature path and geodynamic implications
.
Tectonophysics
 ,
823
, article
229183
,
419.
Mazzucchelli
M.L.
,
Angel
R.J.
and
Alvaro
M.
2021
.
EntraPT: an online platform for elastic geothermobarometry
.
American Mineralogist
 ,
106
,
830
837
,
420.
McCubbin
F.M.
and
Jones
R.H.
2015
.
Extraterrestrial apatite: planetary geochemistry to astrobiology
.
Elements
 ,
11
,
183
188
,
421.
McCubbin
F.M.
,
Boyce
J.W.
et al
2016
.
Geologic history of Martian regolith breccia Northwest Africa 7034: evidence for hydrothermal activity and lithologic diversity in the Martian crust
.
Journal of Geophysical Research: Planets
 ,
121
,
2120
2149
,
422.
McFarlane
C.R.
2016
.
Allanite U–Pb geochronology by 193 nm LA ICP-MS using NIST610 glass for external calibration
.
Chemical Geology
 ,
438
,
91
102
,
423.
McGregor
M.
,
Erickson
T.M.
et al
2021
.
High-resolution EBSD and SIMS U–Pb geochronology of zircon, titanite, and apatite: insights from the Lac La Moinerie impact structure, Canada
.
Contributions to Mineralogy and Petrology
 ,
176
, article
76
,
424.
Mearns
E.W.
1986
.
Sm–Nd ages for Norwegian garnet peridotite
.
Lithos
 ,
19
,
269
278
,
425.
Méheut
M.
,
Ibañez-Mejia
M.
and
Tissot
F.L.
2021
.
Drivers of zirconium isotope fractionation in Zr-bearing phases and melts: the roles of vibrational, nuclear field shift and diffusive effects
.
Geochimica et Cosmochimica Acta
 ,
292
,
217
234
,
426.
Meinhold
G.
,
Anders
B.
,
Kostopoulos
D.
and
Reischmann
T.
2008
.
Rutile chemistry and thermometry as provenance indicator: an example from Chios Island, Greece
.
Sedimentary Geology
 ,
203
,
98
111
,
427.
Meyer
M.
,
John
T.
,
Brandt
S.
and
Klemd
R.
2011
.
Trace element composition of rutile and the application of Zr-in-rutile thermometry to UHT metamorphism (Epupa Complex, NW Namibia)
.
Lithos
 ,
126
,
388
401
,
428.
Mezger
K.
,
Hanson
G.N.
and
Bohlen
S.R.
1989a
.
High-precision U–Pb ages of metamorphic rutile: application to the cooling history of high-grade terranes
.
Earth and Planetary Science Letters
 ,
96
,
106
118
,
429.
Mezger
K.
,
Hanson
G.N.
and
Bohlen
S.R.
1989b
.
U–Pb systematics of garnet: dating the growth of garnet in the Late Archean Pikwitonei granulite domain at Cauchon and Natawahunan lakes, Manitoba, Canada
.
Contributions to Mineralogy and Petrology
 ,
101
,
136
148
,
430.
Mezger
K.
,
Rawnsley
C.M.
,
Bohlen
S.R.
and
Hanson
G.N.
1991
.
U–Pb garnet, sphene, monazite, and rutile ages: implications for the duration of high-grade metamorphism and cooling histories, Adirondack Mts, New York
.
The Journal of Geology
 ,
99
,
415
428
,
431.
Millonig
L.J.
,
Albert
R.
,
Gerdes
A.
,
Avigad
D.
and
Dietsch
C.
2020
.
Exploring laser ablation U–Pb dating of regional metamorphic garnet – The Straits Schist, Connecticut, USA
.
Earth and Planetary Science Letters
 ,
552
, article
116589
,
432.
Millonig
L.J.
,
Beranoaguirre
A.
,
Albert
R.
,
Marschall
H.
,
Baxter
E.
and
Gerdes
A.
2022
. Garnet U–Pb dating by LA-ICPMS: opportunities, limitations, and applications.
EGU General Assembly Abstracts
 ,
Vienna, Austria, EUG22-7077
,
433.
Mills
S.J.
,
Hatert
F.
,
Nickel
E.H.
and
Ferraris
G.
2009
.
The standardisation of mineral group hierarchies: application to recent nomenclature proposals
.
European Journal of Mineralogy
 ,
21
,
1073
1080
,
434.
Möller
A.
,
O'Brien
P.J.
,
Kennedy
A.
and
Kröner
A.
2003
.
Linking growth episodes of zircon and metamorphic textures to zircon chemistry: an example from the ultrahigh-temperature granulites of Rogaland (SW Norway)
.
Geological Society, London, Special Publications
 ,
220
,
65
81
,
435.
Montel
J.-M.
1993
.
A model for monazite/melt equilibrium and application to the generation of granitic magmas
.
Chemical Geology
 ,
110
,
127
146
,
436.
Montel
J.-M.
,
Foret
S.
,
Veschambre
M.
,
Nicollet
C.
and
Provost
A.
1996
.
Electron microprobe dating of monazite
.
Chemical Geology
 ,
131
,
37
53
,
437.
Montel
J.-M.
,
Kato
T.
,
Enami
M.
,
Cocherie
A.
,
Finger
F.
,
Williams
M.
and
Jercinovic
M.
2018
.
Electron-microprobe dating of monazite: the story
.
Chemical Geology
 ,
484
,
4
15
,
438.
Moore
J.
,
Beinlich
A.
,
Piazolo
S.
,
Austrheim
H.
and
Putnis
A.
2020a
.
Metamorphic differentiation via enhanced dissolution along high permeability zones
.
Journal of Petrology
 ,
61
, article
egaa096
,
439.
Moore
J.
,
Beinlich
A.
et al
2020b
.
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
127
,
440.
Morrissey
L.J.
,
Hand
M.
,
Lane
K.
,
Kelsey
D.E.
and
Dutch
R.A.
2016
.
Upgrading iron–ore deposits by melt loss during granulite facies metamorphism
.
Ore Geology Reviews
 ,
74
,
101
121
,
441.
Moser
A.C.
,
Hacker
B.R.
,
Gehrels
G.E.
,
Seward
G.G.E.
,
Kylander-Clark
A.R.C.
and
Garber
J.M.
2022
.
Linking titanite U–Pb dates to coupled deformation and dissolution–reprecipitation
.
Contributions to Mineralogy and Petrology
 ,
177
, article
42
,
442.
Mottram
C.M.
,
Warren
C.J.
,
Regis
D.
,
Roberts
N.M.W.
,
Harris
N.B.W.
,
Argles
T.W.
and
Parrish
R.R.
2014
.
Developing an inverted Barrovian sequence; insights from monazite petrochronology
.
Earth and Planetary Science Letters
 ,
403
,
418
431
,
443.
Mottram
C.M.
,
Parrish
R.R.
,
Regis
D.
,
Warren
C.J.
,
Argles
T.W.
,
Harris
N.B.W.
and
Roberts
N.M.W.
2015
.
Using U–Th–Pb petrochronology to determine rates of ductile thrusting: time windows into the Main Central Thrust, Sikkim Himalaya
.
Tectonics
 ,
34
,
1355
1374
,
444.
Mottram
C.M.
,
Cottle
J.M.
and
Kylander-Clark
A.R.C.
2019
.
Campaign-style U–Pb titanite petrochronology: along-strike variations in timing of metamorphism in the Himalayan metamorphic core
.
Geoscience Frontiers
 ,
10
,
827
847
,
445.
Moyen
J.F.
and
Martin
H.
2012
.
Forty years of TTG research
.
Lithos
 ,
148
,
312
336
.
446.
Moyen
J.F.
and
Stevens
G.
2006
.
Experimental constraints on TTG petrogenesis: implications for Archean geodynamics
. AGU
Geophysical Monograph Series
 ,
164
,
149–175
,
447.
Mühlberg
M.
,
Stevens
G.
,
Moyen
J.-F.
,
Kisters
A.F.
and
Lana
C.
2021
.
Thermal evolution of the Stolzburg Block, Barberton granitoid–greenstone terrain, South Africa: implications for Paleoarchean tectonic processes
.
Precambrian Research
 ,
359
, article
106082
,
448.
Mulder
J.A.
and
Cawood
P.A.
2021
.
Evaluating preservation bias in the continental growth record against the monazite archive
.
Geology
 ,
50
,
243
247
,
449.
Mulder
J.A.
,
Nebel
O.
,
Gardiner
N.J.
,
Cawood
P.A.
,
Wainwright
A.N.
and
Ivanic
T.J.
2021
.
Crustal rejuvenation stabilised Earth's first cratons
.
Nature Communications
 ,
12
,
1
7
,
450.
Muñoz-Montecinos
J.
,
Angiboust
S.
,
Garcia-Casco
A.
and
Raimondo
T.
2023
.
Shattered veins elucidate brittle creep processes in the deep slow slip and tremor region
.
Tectonics
 ,
42
,
e2022TC007605
,
451.
Næraa
T.
,
Scherstén
A.
,
Rosing
M.T.
,
Kemp
A.
,
Hoffmann
J.
,
Kokfelt
T.
and
Whitehouse
M.
2012
.
Hafnium isotope evidence for a transition in the dynamics of continental growth 3.2 Gyr ago
.
Nature
 ,
485
,
627
630
,
452.
Nasdala
L.
,
Zhang
M.
,
Kempe
U.
,
Panczer
G.
,
Gaft
M.
,
Andrut
M.
and
Plötze
M.
2003
.
Spectroscopic methods applied to zircon
.
Reviews in Mineralogy and Geochemistry
 ,
53
,
427
467
,
453.
Nestola
F.
2021
.
How to apply elastic geobarometry in geology
.
American Mineralogist
 ,
106
,
669
671
,
454.
Newton
R.
and
Haselton
H.
1981
. Thermodynamics of the garnet–plagioclase–Al2SiO5–quartz geobarometer. In:
Newton
R.C.
,
Navrotsky
A.
and
Wood
B.J.
(eds)
Thermodynamics of Minerals and Melts
 .
Springer
,
131
147
.
455.
Ning
W.
,
Wang
J.
,
Xiao
D.
,
Li
F.
,
Huang
B.
and
Fu
D.
2019
.
Electron probe microanalysis of monazite and its applications to U–Th–Pb dating of geological samples
.
Journal of Earth Science
 ,
30
,
952
963
,
456.
Nishizawa
M.
,
Takahata
N.
,
Terada
K.
,
Komiya
T.
,
Ueno
Y.
and
Sano
Y.
2005
.
Rare-earth element, lead, carbon, and nitrogen geochemistry of apatite-bearing metasediments from the ∼3.8 Ga Isua Supracrustal Belt, West Greenland
.
International Geology Review
 ,
47
,
952
970
,
457.
Nordsvan
A.R.
,
Collins
W.J.
et al
2018
.
Laurentian crust in northeast Australia: implications for the assembly of the supercontinent Nuna
.
Geology
 ,
46
,
251
254
,
458.
Nutman
A.P.
2007
.
Apatite recrystallisation during prograde metamorphism, Cooma, southeast Australia: implications for using an apatite–graphite association as a biotracer in ancient metasedimentary rocks
.
Australian Journal of Earth Sciences
 ,
54
,
1023
1032
,
459.
Oberli
F.
,
Meier
M.
,
Berger
A.
,
Rosenberg
C.L.
and
Gieré
R.
2004
.
U–Th–Pb and 230Th/238U disequilibrium isotope systematics: precise accessory mineral chronology and melt evolution tracing in the Alpine Bergell intrusion
.
Geochimica et Cosmochimica Acta
 ,
68
,
2543
2560
,
460.
Odlum
M.L.
,
Levy
D.A.
,
Stockli
D.F.
,
Stockli
L.D.
and
DesOrmeau
J.W.
2022
.
Deformation and metasomatism recorded by single-grain apatite petrochronology
.
Geology
 ,
50
,
697
703
,
461.
Olierook
H.K.H.
,
Taylor
R.J.M.
et al
2019
.
Unravelling complex geologic histories using U–Pb and trace element systematics of titanite
.
Chemical Geology
 ,
504
,
105
122
,
462.
Olierook
H.K.H.
,
Rankenburg
K.
et al
2020
.
Resolving multiple geological events using in situ Rb–Sr geochronology: implications for metallogenesis at Tropicana, Western Australia
.
Geochronology
 ,
2
,
283
303
,
463.
Olin
P.H.
and
Wolff
J.A.
2012
.
Partitioning of rare earth and high field strength elements between titanite and phonolitic liquid
.
Lithos
 ,
128–131
,
46
54
,
464.
O'Reilly
S.Y.
and
Griffin
W.
2000
.
Apatite in the mantle: implications for metasomatic processes and high heat production in Phanerozoic mantle
.
Lithos
 ,
53
,
217
232
,
465.
Ortolano
G.
,
Visalli
R.
,
Cirrincione
R.
and
Rebay
G.
2014
.
PT-path reconstruction via unraveling of peculiar zoning pattern in atoll shaped garnets via image assisted analysis: an example from the Santa Lucia del Mela garnet micaschists (northeastern Sicily-Italy)
.
Periodico di Mineralogia
 ,
83
,
257
297
.
466.
Ortolano
G.
,
Visalli
R.
,
Godard
G.
and
Cirrincione
R.
2018
.
Quantitative X-ray map analyser (Q-XRMA): a new GIS-based statistical approach to mineral image analysis
.
Computers & Geosciences
 ,
115
,
56
65
,
467.
O'Sullivan
G.J.
and
Chew
D.M.
2020
.
The clastic record of a Wilson Cycle: evidence from detrital apatite petrochronology of the Grampian–Taconic fore-arc
.
Earth and Planetary Science Letters
 ,
552
, article
116588
,
468.
O'Sullivan
G.J.
,
Chew
D.
,
Morton
A.
,
Mark
C.
and
Henrichs
I.
2018
.
An integrated apatite geochronology and geochemistry tool for sedimentary provenance analysis
.
Geochemistry, Geophysics, Geosystems
 ,
19
,
1309
1326
,
469.
O'Sullivan
G.
,
Chew
D.
,
Kenny
G.
,
Henrichs
I.
and
Mulligan
D.
2020
.
The trace element composition of apatite and its application to detrital provenance studies
.
Earth-Science Reviews
 ,
201
, article
103044
,
470.
O'Sullivan
G.
,
Hoare
B.
,
Mark
C.
,
Drakou
F.
and
Tomlinson
E.
2023
.
Uranium–lead geochronology applied to pyrope garnet with very low concentrations of uranium
.
Geological Magazine
 ,
160
,
1010
1019
,
471.
Page
F.Z.
,
Kita
N.T.
and
Valley
J.W.
2010
.
Ion microprobe analysis of oxygen isotopes in garnets of complex chemistry
.
Chemical Geology
 ,
270
,
9
19
,
472.
Papapavlou
K.
,
Darling
J.R.
,
Storey
C.D.
,
Lightfoot
P.C.
,
Moser
D.E.
and
Lasalle
S.
2017
.
Dating shear zones with plastically deformed titanite: new insights into the orogenic evolution of the Sudbury impact structure (Ontario, Canada)
.
Precambrian Research
 ,
291
,
220
235
,
473.
Pape
J.
,
Mezger
K.
and
Robyr
M.
2016
.
A systematic evaluation of the Zr-in-rutile thermometer in ultra-high temperature (UHT) rocks
.
Contributions to Mineralogy and Petrology
 ,
171
, article
44
,
474.
Parrish
R.R.
1990
.
U–Pb dating of monazite and its application to geological problems
.
Canadian Journal of Earth Sciences
 ,
27
,
1431
1450
,
475.
Passchier
C.W.
and
Simpson
C.
1986
.
Porphyroclast systems as kinematic indicators
.
Journal of Structural Geology
 ,
8
,
831
843
,
476.
Patchett
P.J.
1983
.
Importance of the Lu–Hf isotopic system in studies of planetary chronology and chemical evolution
.
Geochimica et Cosmochimica Acta
 ,
47
,
81
91
,
477.
Paul
A.N.
,
Spikings
R.A.
,
Chew
D.
and
Daly
J.S.
2019
.
The effect of intra-crystal uranium zonation on apatite U–Pb thermochronology: a combined ID-TIMS and LA-MC-ICP-MS study
.
Geochimica et Cosmochimica Acta
 ,
251
,
15
35
,
478.
Paul
A.N.
,
Spikings
R.A.
and
Gaynor
S.P.
2021
.
U–Pb ID-TIMS reference ages and initial Pb isotope compositions for Durango and Wilberforce apatites
.
Chemical Geology
 ,
586
, article
120604
,
479.
Peixoto
E.
,
Alkmim
F.F.
,
Pedrosa-Soares
A.
,
Lana
C.
and
Chaves
A.O.
2018
.
Metamorphic record of collision and collapse in the Ediacaran–Cambrian Araçuaí orogen, SE-Brazil: insights from P–T pseudosections and monazite dating
.
Journal of Metamorphic Geology
 ,
36
,
147
172
,
480.
Penniston-Dorland
S.C.
,
Kohn
M.J.
and
Piccoli
P.M.
2018
.
A mélange of subduction temperatures: evidence from Zr-in-rutile thermometry for strengthening of the subduction interface
.
Earth and Planetary Science Letters
 ,
482
,
525
535
,
481.
Penniston-Dorland
S.C.
,
Baumgartner
L.P.
,
Dragovic
B.
and
Bouvier
A.-S.
2020
.
Li isotope zoning in garnet from Franciscan eclogite and amphibolite: the role of subduction-related fluids
.
Geochimica et Cosmochimica Acta
 ,
286
,
198
213
,
482.
Pe-Piper
G.
,
Piper
D.J.
and
Triantafyllidis
S.
2014
.
Detrital monazite geochronology, Upper Jurassic–Lower Cretaceous of the Scotian Basin: significance for tracking first-cycle sources
.
Geological Society, London, Special Publications
 ,
386
,
293
311
,
483.
Perchuk
L.
and
Lavrent'Eva
I.
1983
.
Experimental investigation of exchange equilibria in the system cordierite–garnet–biotite
.
Advances in Physical Geochemistry
 ,
3
,
199
239
.
484.
Pereira
I.
and
Storey
C.D.
2023
.
Detrital rutile: records of the deep crust, ores and fluids
.
Lithos
 ,
438
439
, article
107010
,
485.
Pereira
I.
,
Storey
C.
,
Darling
J.
,
Lana
C.
and
Alkmim
A.R.
2019
.
Two billion years of evolution enclosed in hydrothermal rutile: recycling of the São Francisco Craton crust and constraints on gold remobilisation processes
.
Gondwana Research
 ,
68
,
69
92
,
486.
Pereira
I.
et al
In press. A review of detrital heavy mineral contributions to furthering our understanding of continental crust formation and evolution
.
Geological Society, London, Special Publications
 ,
487.
Petrus
J.A.
,
Chew
D.M.
,
Leybourne
M.I.
and
Kamber
B.S.
2017
.
A new approach to laser-ablation inductively-coupled-plasma mass-spectrometry (LA-ICP-MS) using the flexible map interrogation tool ‘Monocle
’.
Chemical Geology
 ,
463
,
76
93
,
488.
Peverelli
V.
,
Berger
A.
,
Mulch
A.
,
Pettke
T.
,
Piccoli
F.
and
Herwegh
M.
2022
.
Epidote U–Pb geochronology and H isotope geochemistry trace pre-orogenic hydration of midcrustal granitoids
.
Geology
 ,
50
,
1073
1077
,
489.
Piccoli
P.M.
and
Candela
P.A.
2002
.
Apatite in igneous systems
.
Reviews in Mineralogy and Geochemistry
 ,
48
,
255
292
,
490.
Piccoli
P.
,
Candela
P.
and
Rivers
M.
2000
.
Interpreting magmatic processes from accessory phases: titanite – a small-scale recorder of large-scale processes
.
Earth and Environmental Science Transactions of the Royal Society of Edinburgh
 ,
91
,
257
267
,
491.
Pidgeon
R.T.
,
Chapman
P.G.
,
Danišík
M.
and
Nemchin
A.A.
2017
.
Dry annealing of metamict zircon: a differential scanning calorimetry study
.
American Mineralogist
 ,
102
,
1066
1072
,
492.
Pidgeon
R.T.
,
Nemchin
A.
,
Roberts
M.
,
Whitehouse
M.J.
and
Bellucci
J.
2019
.
The accumulation of non-formula elements in zircons during weathering: ancient zircons from the Jack Hills, Western Australia
.
Chemical Geology
 ,
530
, article
119310
,
493.
Piechocka
A.M.
,
Gregory
C.J.
,
Zi
J.-W.
,
Sheppard
S.
,
Wingate
M.T.D.
and
Rasmussen
B.
2017
.
Monazite trumps zircon: applying SHRIMP U–Pb geochronology to systematically evaluate emplacement ages of leucocratic, low-temperature granites in a complex Precambrian orogen
.
Contributions to Mineralogy and Petrology
 ,
172
, article
63
,
494.
Plank
T.
,
Cooper
L.B.
and
Manning
C.E.
2009
.
Emerging geothermometers for estimating slab surface temperatures
.
Nature Geoscience
 ,
2
,
611
615
,
495.
Plavsa
D.
,
Reddy
S.
,
Clark
C.
and
Agangi
A.
2018
.
Capricorn Orogen Rutile Study: a Combined Electrons Backscatter Diffraction (EBSD) and Laser Ablation Split Stream (LASS) Analytical Approach
 . Government of Western Australia.
496.
Pochon
A.
,
Beaudoin
G.
,
Branquet
Y.
,
Boulvais
P.
,
Gloaguen
E.
and
Gapais
D.
2017
.
Metal mobility during hydrothermal breakdown of Fe–Ti oxides: insights from Sb–Au mineralizing event (Variscan Armorican Massif, France)
.
Ore Geology Reviews
 ,
91
,
66
99
,
497.
Poitrasson
F.
,
Chenery
S.
and
Shepherd
T.J.
2000
.
Electron microprobe and LA-ICP-MS study of monazite hydrothermal alteration: implications for U–Th–Pb geochronology and nuclear ceramics
.
Geochimica et Cosmochimica Acta
 ,
64
,
3283
3297
,
498.
Pollington
A.D.
and
Baxter
E.F.
2010
.
High resolution Sm–Nd garnet geochronology reveals the uneven pace of tectonometamorphic processes
.
Earth and Planetary Science Letters
 ,
293
,
63
71
,
499.
Pollington
A.D.
and
Baxter
E.F.
2011
.
High precision microsampling and preparation of zoned garnet porphyroblasts for Sm–Nd geochronology
.
Chemical Geology
 ,
281
,
270
282
,
500.
Porter
J.K.
,
McNaughton
N.J.
,
Evans
N.J.
and
McDonald
B.J.
2020
.
Rutile as a pathfinder for metals exploration
.
Ore Geology Reviews
 ,
120
, article
103406
,
501.
Potts
N.J.
,
Barnes
J.J.
,
Tartèse
R.
,
Franchi
I.A.
and
Anand
M.
2018
.
Chlorine isotopic compositions of apatite in Apollo 14 rocks: evidence for widespread vapor-phase metasomatism on the lunar nearside c. 4 billion years ago
.
Geochimica et Cosmochimica Acta
 ,
230
,
46
59
,
502.
Pourteau
A.
,
Scherer
E.E.
,
Schorn
S.
,
Bast
R.
,
Schmidt
A.
and
Ebert
L.
2019
.
Thermal evolution of an ancient subduction interface revealed by Lu–Hf garnet geochronology, Halilbağı Complex (Anatolia)
.
Geoscience Frontiers
 ,
10
,
127
148
,
503.
Prent
A.M.
,
Beinlich
A.
,
Morrissey
L.J.
,
Raimondo
T.
,
Clark
C.
and
Putnis
A.
2019
.
Monazite as a monitor for melt–rock interaction during cooling and exhumation
.
Journal of Metamorphic Geology
 ,
37
,
415
438
,
504.
Prent
A.M.
,
Beinlich
A.
,
Raimondo
T.
,
Kirkland
C.L.
,
Evans
N.J.
and
Putnis
A.
2020
.
Apatite and monazite: an effective duo to unravel superimposed fluid-flow and deformation events in reactivated shear zones
.
Lithos
 ,
376–377
, article
105752
,
505.
Prowatke
S.
and
Klemme
S.
2005
.
Effect of melt composition on the partitioning of trace elements between titanite and silicate melt
.
Geochimica et Cosmochimica Acta
 ,
69
,
695
709
,
506.
Pupin
J.P.
1980
.
Zircon and granite petrology
.
Contributions to Mineralogy and Petrology
 ,
73
,
207
220
,
507.
Pyle
J.M.
and
Spear
F.S.
1999
.
Yttrium zoning in garnet: coupling of major and accessory phases during metamorphic reactions
.
Geological Materials Research
 ,
1
,
1
49
.
508.
Qin
T.
,
Wu
F.
,
Wu
Z.
and
Huang
F.
2016
.
First-principles calculations of equilibrium fractionation of O and Si isotopes in quartz, albite, anorthite, and zircon
.
Contributions to Mineralogy and Petrology
 ,
171
,
1
14
,
509.
Radulescu
I.G.
,
Rubatto
D.
,
Gregory
C.
and
Compagnoni
R.
2009
.
The age of HP metamorphism in the Gran Paradiso Massif, Western Alps: a petrological and geochronological study of ‘silvery micaschists
’.
Lithos
 ,
110
,
95
108
,
510.
Raimondo
T.
,
Clark
C.
,
Hand
M.
,
Cliff
J.
and
Harris
C.
2012
.
High-resolution geochemical record of fluid–rock interaction in a mid-crustal shear zone: a comparative study of major element and oxygen isotope transport in garnet
.
Journal of Metamorphic Geology
 ,
30
,
255
280
,
511.
Raimondo
T.
,
Payne
J.
,
Wade
B.
,
Lanari
P.
,
Clark
C.
and
Hand
M.
2017
.
Trace element mapping by LA-ICP-MS: assessing geochemical mobility in garnet
.
Contributions to Mineralogy and Petrology
 ,
172
, article
17
,
512.
Rasmussen
B.
and
Muhling
J.R.
2007
.
Monazite begets monazite: evidence for dissolution of detrital monazite and reprecipitation of syntectonic monazite during low-grade regional metamorphism
.
Contributions to Mineralogy and Petrology
 ,
154
,
675
689
,
513.
Ravindran
A.
,
Mezger
K.
,
Balakrishnan
S.
,
Kooijman
E.
,
Schmitt
M.
and
Berndt
J.
2020
.
Initial 87Sr/86Sr as a sensitive tracer of Archaean crust–mantle evolution: constraints from igneous and sedimentary rocks in the western Dharwar Craton, India
.
Precambrian Research
 ,
337
, article
105523
,
514.
Reddy
S.M.
,
Saxey
D.W.
,
Rickard
W.D.A.
,
Fougerouse
D.
,
Montalvo
S.D.
,
Verberne
R.
and
van Riessen
A.
2020
.
Atom probe tomography: development and application to the geosciences
.
Geostandards and Geoanalytical Research
 ,
44
,
5
50
,
515.
Regis
D.
,
Rubatto
D.
,
Darling
J.
,
Cenki-Tok
B.
,
Zucali
M.
and
Engi
M.
2014
.
Multiple metamorphic stages within an eclogite-facies terrane (Sesia Zone, Western Alps) revealed by Th–U–Pb petrochronology
.
Journal of Petrology
 ,
55
,
1429
1456
,
516.
Regis
D.
,
Warren
C.J.
,
Mottram
C.M.
and
Roberts
N.M.W.
2016
.
Using monazite and zircon petrochronology to constrain the P–T–t evolution of the middle crust in the Bhutan Himalaya
.
Journal of Metamorphic Geology
 ,
34
,
617
639
,
517.
Renna
M.R.
,
Tribuzio
R.
and
Tiepolo
M.
2007
.
Origin and timing of the post-Variscan gabbro–granite complex of Porto (Western Corsica)
.
Contributions to Mineralogy and Petrology
 ,
154
,
493
517
,
518.
Rezvukhina
O.V.
,
Skublov
S.G.
,
Rezvukhin
D.I.
and
Korsakov
A.V.
2021
.
Rutile in diamondiferous metamorphic rocks: new insights from trace-element composition, mineral/fluid inclusions, and U–Pb ID-TIMS dating
.
Lithos
 ,
394–395
, article
106172
,
519.
Ribeiro
B.V.
,
Finch
M.A.
et al
2022
.
From microanalysis to supercontinents: insights from the Rio Apa Terrane into the Mesoproterozoic SW Amazonian Craton evolution during Rodinia assembly
.
Journal of Metamorphic Geology
 ,
40
,
631
663
,
520.
Romer
R.L.
and
Rötzler
J.
2001
.
P–T–t evolution of ultrahigh-temperature granulites from the Saxon Granulite Massif, Germany. Part II: geochronology
.
Journal of Petrology
 ,
42
,
2015
2032
,
521.
Romer
R.L.
and
Rötzler
J.
2011
.
The role of element distribution for the isotopic dating of metamorphic minerals
.
European Journal of Mineralogy
 ,
23
,
17
33
,
522.
Romer
R.L.
and
Siegesmund
S.
2003
.
Why allanite may swindle about its true age
.
Contributions to Mineralogy and Petrology
 ,
146
,
297
307
,
523.
Romer
R.L.
and
Xiao
Y.
2005
.
Initial Pb–Sr (–Nd) isotopic heterogeneity in a single allanite–epidote crystal: implications of reaction history for the dating of minerals with low parent-to-daughter ratios
.
Contributions to Mineralogy and Petrology
 ,
148
,
662
674
,
524.
Rubatto
D.
2002
.
Zircon trace element geochemistry: partitioning with garnet and the link between U–Pb ages and metamorphism
.
Chemical Geology
 ,
184
,
123
138
,
525.
Rubatto
D.
2017
.
Zircon: the metamorphic mineral
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
261
295
,
526.
Rubatto
D.
and
Angiboust
S.
2015
.
Oxygen isotope record of oceanic and high-pressure metasomatism: a P–T–time–fluid path for the Monviso eclogites (Italy)
.
Contributions to Mineralogy and Petrology
 ,
170
, article
44
,
527.
Rubatto
D.
and
Hermann
J.
2007
.
Experimental zircon/melt and zircon/garnet trace element partitioning and implications for the geochronology of crustal rocks
.
Chemical Geology
 ,
241
,
38
61
,
528.
Rubatto
D.
,
Williams
I.S.
and
Buick
I.S.
2001
.
Zircon and monazite response to prograde metamorphism in the Reynolds Range, central Australia
.
Contributions to Mineralogy and Petrology
 ,
140
,
458
468
,
529.
Rubatto
D.
,
Hermann
J.
and
Buick
I.S.
2006
.
Temperature and bulk composition control on the growth of monazite and zircon during low-pressure anatexis (Mount Stafford, central Australia)
.
Journal of Petrology
 ,
47
,
1973
1996
,
530.
Rubatto
D.
,
Hermann
J.
,
Berger
A.
and
Engi
M.
2009
.
Protracted fluid-induced melting during Barrovian metamorphism in the Central Alps
.
Contributions to Mineralogy and Petrology
 ,
158
,
703
722
,
531.
Rubatto
D.
,
Regis
D.
,
Hermann
J.
,
Boston
K.
,
Engi
M.
,
Beltrando
M.
and
McAlpine
S.R.
2011
.
Yo-yo subduction recorded by accessory minerals in the Italian Western Alps
.
Nature Geoscience
 ,
4
,
338
342
,
532.
Rubatto
D.
,
Chakraborty
S.
and
Dasgupta
S.
2013
.
Timescales of crustal melting in the Higher Himalayan Crystallines (Sikkim, Eastern Himalaya) inferred from trace element-constrained monazite and zircon chronology
.
Contributions to Mineralogy and Petrology
 ,
165
,
349
372
,
533.
Rubatto
D.
,
Burger
M.
et al
2020
.
Identification of growth mechanisms in metamorphic garnet by high-resolution trace element mapping with LA-ICP-TOFMS
.
Contributions to Mineralogy and Petrology
 ,
175
, article
61
,
534.
Rudnick
R.L.
,
Barth
M.
,
Horn
I.
and
McDonough
W.F.
2000
.
Rutile-bearing refractory eclogites: missing link between continents and depleted mantle
.
Science
 ,
287
,
278
281
,
535.
Russell
A.K.
,
Kitajima
K.
,
Strickland
A.
,
Medaris
L.G.
,
Schulze
D.J.
and
Valley
J.W.
2013
.
Eclogite-facies fluid infiltration: constraints from δ18O zoning in garnet
.
Contributions to Mineralogy and Petrology
 ,
165
,
103
116
,
536.
Salama
W.
,
Anand
R.
and
Roberts
M.
2018
.
Cassiterite and rutile as indicator minerals for exploring the VMS system
.
ASEG Extended Abstracts
 ,
2018
,
1
4
,
537.
Salminen
P.E.
,
Hölttä
P.
,
Lahtinen
R.
and
Sayab
M.
2022
.
Monazite record for the Paleoproterozoic Svecofennian orogeny, SE Finland: an over 150-Ma spread of monazite dates
.
Lithos
 ,
416–417
, article
106654
,
538.
Savage
P.S.
,
Armytage
R.M.
,
Georg
R.B.
and
Halliday
A.N.
2014
.
High temperature silicon isotope geochemistry
.
Lithos
 ,
190
,
500
519
,
539.
Sayab
M.
2006
.
Decompression through clockwise P–T path: implications for early N–S shortening orogenesis in the Mesoproterozoic Mt Isa Inlier (NE Australia)
.
Journal of Metamorphic Geology
 ,
24
,
89
105
,
540.
Sayab
M.
,
Suuronen
J.-P.
,
Hölttä
P.
,
Aerden
D.
,
Lahtinen
R.
and
Kallonen
A.P.
2015
.
High-resolution X-ray computed microtomography: a holistic approach to metamorphic fabric analyses
.
Geology
 ,
43
,
55
58
,
541.
Sayab
M.
,
Suuronen
J.-P.
et al
2016
.
Three-dimensional textural and quantitative analyses of orogenic gold at the nanoscale
.
Geology
 ,
44
,
739
742
,
542.
Schaltegger
U.
,
Brack
P.
et al
2009
.
Zircon and titanite recording 1.5 million years of magma accretion, crystallization and initial cooling in a composite pluton (southern Adamello batholith, northern Italy)
.
Earth and Planetary Science Letters
 ,
286
,
208
218
,
543.
Schannor
M.
,
Lana
C.
,
Nicoli
G.
,
Cutts
K.
,
Buick
I.
,
Gerdes
A.
and
Hecht
L.
2021
.
Reconstructing the metamorphic evolution of the Araçuaí orogen (SE Brazil) using in situ U–Pb garnet dating and P–T modelling
.
Journal of Metamorphic Geology
 ,
39
,
1145
1171
,
544.
Schärer
U.
,
Lian-Sheng
Z.
and
Tapponnier
P.
1994
.
Duration of strike-slip movements in large shear zones: the Red River belt, China
.
Earth and Planetary Science Letters
 ,
126
,
379
397
,
545.
Scherer
E.E.
,
Cameron
K.L.
and
Blichert-Toft
J.
2000
.
Lu–Hf garnet geochronology: closure temperature relative to the Sm–Nd system and the effects of trace mineral inclusions
.
Geochimica et Cosmochimica Acta
 ,
64
,
3413
3432
,
546.
Scherer
E.
,
Münker
C.
and
Mezger
K.
2001
.
Calibration of the lutetium–hafnium clock
.
Science
 ,
293
,
683
687
,
547.
Scherer
E.E.
,
Whitehouse
M.J.
and
Munker
C.
2007
.
Zircon as a monitor of crustal growth
.
Elements
 ,
3
,
19
24
,
548.
Schiller
D.
and
Finger
F.
2019
.
Application of Ti-in-zircon thermometry to granite studies: problems and possible solutions
.
Contributions to Mineralogy and Petrology
 ,
174
, article
51
,
549.
Schirra
M.
and
Laurent
O.
2021
.
Petrochronology of hydrothermal rutile in mineralized porphyry Cu systems
.
Chemical Geology
 ,
581
, article
120407
,
550.
Schmidt
A.
,
Weyer
S.
,
John
T.
and
Brey
G.P.
2009
.
HFSE systematics of rutile-bearing eclogites: new insights into subduction zone processes and implications for the Earth's HFSE budget
.
Geochimica et Cosmochimica Acta
 ,
73
,
455
468
,
551.
Schmidt
A.
,
Pourteau
A.
,
Candan
O.
and
Oberhänsli
R.
2015
.
Lu–Hf geochronology on cm-sized garnets using microsampling: new constraints on garnet growth rates and duration of metamorphism during continental collision (Menderes Massif, Turkey)
.
Earth and Planetary Science Letters
 ,
432
,
24
35
,
552.
Schmitz
M.D.
and
Bowring
S.A.
2003
.
Ultrahigh-temperature metamorphism in the lower crust during Neoarchean Ventersdorp rifting and magmatism, Kaapvaal Craton, southern Africa
.
GSA Bulletin
 ,
115
,
533
548
,
553.
Schoene
B.
2014
. U–Th–Pb geochronology. In:
Holland
H.D.
and
Turekian
K.K.
(eds)
Treatise on Geochemistry
 , 2nd edn,
341
378
,
554.
Schoene
B.
and
Bowring
S.A.
2007
.
Determining accurate temperature–time paths from U–Pb thermochronology: an example from the Kaapvaal Craton, southern Africa
.
Geochimica et Cosmochimica Acta
 ,
71
,
165
185
,
555.
Schönig
J.
,
von Eynatten
H.
,
Meinhold
G.
and
Lünsdorf
N.K.
2019
.
Diamond and coesite inclusions in detrital garnet of the Saxonian Erzgebirge, Germany
.
Geology
 ,
47
,
715
718
,
556.
Schulz
B.
2021
.
Monazite microstructures and their interpretation in petrochronology
.
Frontiers in Earth Science
 ,
9
,
557.
Schwandt
C.S.
,
Papike
J.J.
and
Shearer
C.K.
1996
.
Trace element zoning in pelitic garnet of the Black Hills, South Dakota
.
American Mineralogist
 ,
81
,
1195
1207
,
558.
Scibiorski
E.A.
and
Cawood
P.A.
2022
.
Titanite as a petrogenetic indicator
.
Terra Nova
 ,
34
,
177
183
,
559.
Scicchitano
M.R.
,
Spicuzza
M.J.
,
Ellison
E.T.
,
Tuschel
D.
,
Templeton
A.S.
and
Valley
J.W.
2021
.
In situ oxygen isotope determination in serpentine minerals by SIMS: addressing matrix effects and providing new insights on serpentinisation at Hole BA1B (Samail ophiolite, Oman)
.
Geostandards and Geoanalytical Research
 ,
45
,
161
187
,
560.
Sciuba
M.
and
Beaudoin
G.
2021
.
Texture and trace element composition of rutile in orogenic gold deposits
.
Economic Geology
 ,
116
,
1865
1892
,
561.
Scott
D.J.
and
St-Onge
M.R.
1995
.
Constraints on Pb closure temperature in titanite based on rocks from the Ungava orogen, Canada: implications for U–Pb geochronology and P–T–t path determinations
.
Geology
 ,
23
,
1123
1126
,
562.
Scott
K.M.
and
Radford
N.W.
2007
.
Rutile compositions at the Big Bell Au deposit as a guide for exploration
.
Geochemistry: Exploration, Environment, Analysis
 ,
7
,
353
361
,
563.
Seman
S.
,
Stockli
D.F.
and
McLean
N.M.
2017
.
U–Pb geochronology of grossular-andradite garnet
.
Chemical Geology
 ,
460
,
106
116
,
564.
Seydoux-Guillaume
A.-M.
,
Paquette
J.-L.
,
Wiedenbeck
M.
,
Montel
J.-M.
and
Heinrich
W.
2002
.
Experimental resetting of the U–Th–Pb systems in monazite
.
Chemical Geology
 ,
191
,
165
181
,
565.
Seydoux-Guillaume
A.-M.
,
Montel
J.-M.
et al
2012
.
Low-temperature alteration of monazite: fluid mediated coupled dissolution–precipitation, irradiation damage, and disturbance of the U–Pb and Th–Pb chronometers
.
Chemical Geology
 ,
330–331
,
140
158
,
566.
Sha
L.-K.
and
Chappell
B.W.
1999
.
Apatite chemical composition, determined by electron microprobe and laser-ablation inductively coupled plasma mass spectrometry, as a probe into granite petrogenesis
.
Geochimica et Cosmochimica Acta
 ,
63
,
3861
3881
,
567.
Shrestha
S.
,
Larson
K.P.
,
Duesterhoeft
E.
,
Soret
M.
and
Cottle
J.M.
2019
.
Thermodynamic modelling of phosphate minerals and its implications for the development of P–T–t histories: a case study in garnet-monazite bearing metapelites
.
Lithos
 ,
334–335
,
141
160
,
568.
Shrestha
S.
,
Larson
K.P.
,
Martin
A.J.
,
Guilmette
C.
,
Smit
M.A.
and
Cottle
J.M.
2020
.
The Greater Himalayan thrust belt: insight into the assembly of the exhumed Himalayan metamorphic core, Modi Khola Valley, Central Nepal
.
Tectonics
 ,
39
, article
e2020TC006252
,
569.
Simpson
A.
,
Gilbert
S.
et al
2021
.
In-situ Lu–Hf geochronology of garnet, apatite and xenotime by LA ICP MS/MS
.
Chemical Geology
 ,
577
, article
120299
,
570.
Simpson
A.
,
Glorie
S.
,
Hand
M.
,
Spandler
C.
,
Gilbert
S.
and
Cave
B.
2022
.
In-situ Lu–Hf geochronology of calcite
.
GChron
 ,
4
,
353
372
,
571.
Simpson
A.
,
Glorie
S.
,
Hand
M.
,
Spandler
C.
and
Gilbert
S.
2023
.
Garnet Lu-Hf speed dating: A novel method to rapidly resolve polymetamorphic histories
.
Gondwana Research
 ,
572.
Sláma
J.
,
Košler
J.
and
Pedersen
R.
2007
.
Behaviour of zircon in high-grade metamorphic rocks: evidence from Hf isotopes, trace elements and textural studies
.
Contributions to Mineralogy and Petrology
 ,
154
,
335
356
,
573.
Sliwinski
J.T.
and
Stoll
H.M.
2021
.
Combined fluorescence imaging and LA-ICP-MS trace element mapping of stalagmites: microfabric identification and interpretation
.
Chemical Geology
 ,
581
, article
120397
,
574.
Smit
M.A.
,
Scherer
E.E.
,
Bröcker
M.
and
van Roermund
H.L.M.
2010
.
Timing of eclogite facies metamorphism in the southernmost Scandinavian Caledonides by Lu–Hf and Sm–Nd geochronology
.
Contributions to Mineralogy and Petrology
 ,
159
,
521
539
,
575.
Smit
M.A.
,
Scherer
E.E.
and
Mezger
K.
2013
.
Lu–Hf and Sm–Nd garnet geochronology: chronometric closure and implications for dating petrological processes
.
Earth and Planetary Science Letters
 ,
381
,
222
233
,
576.
Smith
H.A.
and
Barreiro
B.
1990
.
Monazite U–Pb dating of staurolite grade metamorphism in pelitic schists
.
Contributions to Mineralogy and Petrology
 ,
105
,
602
615
,
577.
Smith
H.A.
and
Giletti
B.J.
1997
.
Lead diffusion in monazite
.
Geochimica et Cosmochimica Acta
 ,
61
,
1047
1055
,
578.
Smith
M.
,
Storey
C.
,
Jeffries
T.
and
Ryan
C.
2009
.
In situ U–Pb and trace element analysis of accessory minerals in the Kiruna district, Norrbotten, Sweden: new constraints on the timing and origin of mineralization
.
Journal of Petrology
 ,
50
,
2063
2094
,
579.
Smithies
R.H.
,
Lu
Y.
et al
2021
.
Oxygen isotopes trace the origins of Earth's earliest continental crust
.
Nature
 ,
592
,
70
75
,
580.
Smye
A.J.
,
Roberts
N.M.
,
Condon
D.J.
,
Horstwood
M.S.
and
Parrish
R.R.
2014
.
Characterising the U–Th–Pb systematics of allanite by ID and LA-ICPMS: implications for geochronology
.
Geochimica et Cosmochimica Acta
 ,
135
,
1
28
,
581.
Smye
A.J.
,
Marsh
J.
,
Vermeesch
P.
,
Garber
J.
and
Stockli
D.
2018
.
Applications and limitations of U–Pb thermochronology to middle and lower crustal thermal histories
.
Chemical Geology
 ,
494
,
1
18
,
582.
Smythe
D.J.
and
Brenan
J.M.
2016
.
Magmatic oxygen fugacity estimated using zircon–melt partitioning of cerium
.
Earth and Planetary Science Letters
 ,
453
,
260
266
,
583.
Smythe
D.
,
Schulze
D.
and
Brenan
J.
2008
.
Rutile as a kimberlite indicator mineral: minor and trace element geochemistry
.
International Kimberlite Conference: Extended Abstracts
,
Frankfurt, Germany
,
9
.
584.
Söderlund
U.
,
Patchett
P.J.
,
Vervoort
J.D.
and
Isachsen
C.E.
2004
.
The 176Lu decay constant determined by Lu–Hf and U–Pb isotope systematics of Precambrian mafic intrusions
.
Earth and Planetary Science Letters
 ,
219
,
311
324
,
585.
Spandler
C.
,
Hammerli
J.
,
Sha
P.
,
Hilbert-Wolf
H.
,
Hu
Y.
,
Roberts
E.
and
Schmitz
M.
2016
.
MKED1: a new titanite standard for in situ analysis of Sm–Nd isotopes and U–Pb geochronology
.
Chemical Geology
 ,
425
,
110
126
,
586.
Spear
F.S.
2010
.
Monazite–allanite phase relations in metapelites
.
Chemical Geology
 ,
279
,
55
62
,
587.
Spear
F.S.
2017
.
Garnet growth after overstepping
.
Chemical Geology
 ,
466
,
491
499
,
588.
Spear
F.S.
and
Parrish
R.R.
1996
.
Petrology and cooling rates of the Valhalla complex, British Columbia, Canada
.
Journal of Petrology
 ,
37
,
733
765
,
589.
Spear
F.S.
and
Pyle
J.M.
2002
.
Apatite, monazite, and xenotime in metamorphic rocks
.
Reviews in Mineralogy and Geochemistry
 ,
48
,
293
335
,
590.
Spear
F.S.
and
Pyle
J.M.
2010
.
Theoretical modeling of monazite growth in a low-Ca metapelite
.
Chemical Geology
 ,
273
,
111
119
,
591.
Spear
F.S.
and
Wolfe
O.M.
2020
.
Revaluation of ‘equilibrium’ P–T paths from zoned garnet in light of quartz inclusion in garnet (QuiG) barometry
.
Lithos
 ,
372–373
, article
105650
,
592.
Spear
F.S.
,
Selverstone
J.
,
Hickmott
D.
,
Crowley
P.
and
Hodges
K.V.
1984
.
P–T paths from garnet zoning: a new technique for deciphering tectonic processes in crystalline terranes
.
Geology
 ,
12
,
87
90
,
593.
Spencer
C.J.
,
Kirkland
C.L.
and
Taylor
R.J.M.
2016
.
Strategies towards statistically robust interpretations of in situ U–Pb zircon geochronology
.
Geoscience Frontiers
 ,
7
,
581
589
,
594.
Spencer
C.J.
,
Kirkland
C.
,
Roberts
N.
,
Evans
N.
and
Liebmann
J.
2020
.
Strategies towards robust interpretations of in situ zircon Lu–Hf isotope analyses
.
Geoscience Frontiers
 ,
11
,
843
853
,
595.
Spencer
K.J.
,
Hacker
B.R.
et al
2013
.
Campaign-style titanite U–Pb dating by laser-ablation ICP: implications for crustal flow, phase transformations and titanite closure
.
Chemical Geology
 ,
341
,
84
101
,
596.
Stearns
M.A.
,
Hacker
B.R.
,
Ratschbacher
L.
,
Rutte
D.
and
Kylander-Clark
A.R.C.
2015
.
Titanite petrochronology of the Pamir gneiss domes: implications for middle to deep crust exhumation and titanite closure to Pb and Zr diffusion
.
Tectonics
 ,
34
,
784
802
,
597.
Stepanov
A.S.
,
Hermann
J.
,
Rubatto
D.
and
Rapp
R.P.
2012
.
Experimental study of monazite/melt partitioning with implications for the REE, Th and U geochemistry of crustal rocks
.
Chemical Geology
 ,
300
,
200
220
,
598.
Stern
R.A.
and
Rayner
N.M.
2003
. Ages of Several Xenotime Megacrysts by ID-TIMS.
Potential Reference Materials for Ion Microprobe U–Pb Geochronology
 .
Geological Survey of Canada
.
599.
Štípská
P.
,
Hacker
B.R.
,
Racek
M.
,
Holder
R.
,
Kylander-Clark
A.R.C.
,
Schulmann
K.
and
Hasalová
P.
2015
.
Monazite dating of prograde and retrograde P–T–d paths in the Barrovian terrane of the Thaya window, Bohemian Massif
.
Journal of Petrology
 ,
56
,
1007
1035
,
600.
St-Onge
M.R.
1987
.
Zoned poikiloblastic garnets: P–T paths and syn-metamorphic uplift through 30 km of structural depth, Wopmay Orogen, Canada
.
Journal of Petrology
 ,
28
,
1
21
,
601.
Storey
C.D.
,
Jeffries
T.E.
and
Smith
M.
2006
.
Common lead-corrected laser ablation ICP-MS U–Pb systematics and geochronology of titanite
.
Chemical Geology
 ,
227
,
37
52
,
602.
Strzelecki
A.C.
,
Reece
M.
et al
2022
.
Crystal chemistry and thermodynamics of HREE (Er, Yb) mixing in a xenotime solid solution
.
ACS Earth and Space Chemistry
 ,
6
,
1375
1389
,
603.
Stünitz
H.
and
Tullis
J.
2001
.
Weakening and strain localization produced by syn-deformational reaction of plagioclase
.
International Journal of Earth Sciences
 ,
90
,
136
148
,
604.
Su
J.-H.
,
Zhao
X.-F.
,
Li
X.-C.
,
Su
Z.-K.
,
Liu
R.
,
Qin
Z.-J.
and
Chen
M.
2021
.
Fingerprinting REE mineralization and hydrothermal remobilization history of the carbonatite-alkaline complexes, Central China: constraints from in situ elemental and isotopic analyses of phosphate minerals
.
American Mineralogist: Journal of Earth and Planetary Materials
 ,
106
,
1545
1558
,
605.
Sun
J.-F.
,
Yang
J.-H.
,
Wu
F.-Y.
,
Li
X.-H.
,
Yang
Y.-H.
,
Xie
L.-W.
and
Wilde
S.A.
2010
.
Magma mixing controlling the origin of the Early Cretaceous Fangshan granitic pluton, North China Craton: in situ U–Pb age and Sr-, Nd-, Hf- and O-isotope evidence
.
Lithos
 ,
120
,
421
438
,
606.
Suzuki
K.
and
Adachi
M.
1994
.
Middle Precambrian detrital monazite and zircon from the Hida gneiss on Oki-Dogo Island, Japan: their origin and implications for the correlation of basement gneiss of southwest Japan and Korea
.
Tectonophysics
 ,
235
,
277
292
,
607.
Suzuki
K.
and
Kato
T.
2008
.
CHIME dating of monazite, xenotime, zircon and polycrase: protocol, pitfalls and chemical criterion of possibly discordant age data
.
Gondwana Research
 ,
14
,
569
586
,
608.
Suzuki
K.
,
Adachi
M.
and
Kajizuka
I.
1994
.
Electron microprobe observations of Pb diffusion in metamorphosed detrital monazites
.
Earth and Planetary Science Letters
 ,
128
,
391
405
,
609.
Tacchetto
T.
,
Reddy
S.M.
,
Saxey
D.W.
,
Fougerouse
D.
,
Rickard
W.D.A.
and
Clark
C.
2021
.
Disorientation control on trace element segregation in fluid-affected low-angle boundaries in olivine
.
Contributions to Mineralogy and Petrology
 ,
176
, article
59
,
610.
Tacchetto
T.
,
Clark
C.
,
Erickson
T.
,
Reddy
S.M.
,
Bhowany
K.
and
Hand
M.
2022
.
Weakening the lower crust: conditions, reactions and deformation
.
Lithos
 ,
422–423
, article
106738
,
611.
Tamblyn
R.
,
Hand
M.
,
Simpson
A.
,
Gilbert
S.
,
Wade
B.
and
Glorie
S.
2022
.
In situ laser ablation Lu–Hf geochronology of garnet across the Western Gneiss Region: campaign-style dating of metamorphism
.
Journal of the Geological Society, London
 ,
179
, article
jgs2021-094
,
612.
Taylor
R.J.M.
,
Clark
C.
,
Fitzsimons
I.C.W.
,
Santosh
M.
,
Hand
M.
,
Evans
N.
and
McDonald
B.
2014
.
Post-peak, fluid-mediated modification of granulite facies zircon and monazite in the Trivandrum Block, southern India
.
Contributions to Mineralogy and Petrology
 ,
168
, article
1044
,
613.
Taylor
R.J.M.
,
Harley
S.L.
,
Hinton
R.W.
,
Elphick
S.
,
Clark
C.
and
Kelly
N.M.
2015
.
Experimental determination of REE partition coefficients between zircon, garnet and melt: a key to understanding high-T crustal processes
.
Journal of Metamorphic Geology
 ,
33
,
231
248
,
614.
Taylor
R.J.M.
,
Kirkland
C.L.
and
Clark
C.
2016
.
Accessories after the facts: constraining the timing, duration and conditions of high-temperature metamorphic processes
.
Lithos
 ,
264
,
239
257
,
615.
Taylor
R.J.M.
,
Clark
C.
,
Harley
S.L.
,
Kylander-Clark
A.R.C.
,
Hacker
B.R.
and
Kinny
P.D.
2017
.
Interpreting granulite facies events through rare earth element partitioning arrays
.
Journal of Metamorphic Geology
 ,
35
,
759
775
,
616.
Taylor
R.J.M.
,
Johnson
T.E.
,
Clark
C.
and
Harrison
R.J.
2020
.
Persistence of melt-bearing Archean lower crust for >200 my – an example from the Lewisian Complex, northwest Scotland
.
Geology
 ,
48
,
221
225
,
617.
Taylor-Jones
K.
and
Powell
R.
2015
.
Interpreting zirconium-in-rutile thermometric results
.
Journal of Metamorphic Geology
 ,
33
,
115
122
,
618.
Tedeschi
M.
,
Rossi Vieira
P.L.
et al
2023
.
Unravelling the protracted U–Pb zircon geochronological record of high to ultrahigh temperature metamorphic rocks: implications for provenance investigations
.
Geoscience Frontiers
 ,
14
, article
101515
,
619.
Tera
F.
and
Wasserburg
G.J.
1972
.
U–Th–Pb systematics in lunar highland samples from the Luna 20 and Apollo 16 missions
.
Earth and Planetary Science Letters
 ,
17
,
36
51
,
620.
Thomas
J.B.
,
Bodnar
R.J.
,
Shimizu
N.
and
Chesner
C.A.
2003
.
Melt inclusions in zircon
.
Reviews in Mineralogy and Geochemistry
 ,
53
,
63
87
,
621.
Thomas
R.
and
Davidson
P.
2012
.
Water in granite and pegmatite-forming melts
.
Ore Geology Reviews
 ,
46
,
32
46
,
622.
Thompson
J.M.
,
Goemann
K.
,
Belousov
I.
,
Jenkins
K.
,
Kobussen
A.
,
Powell
W.
and
Danyushevsky
L.
2021
.
Assessment of the mineral ilmenite for U–Pb dating by LA-ICP-MS
.
Journal of Analytical Atomic Spectrometry
 ,
36
,
1244
1260
,
623.
Thomson
S.N.
,
Gehrels
G.E.
,
Ruiz
J.
and
Buchwaldt
R.
2012
.
Routine low-damage apatite U–Pb dating using laser ablation–multicollector–ICPMS
.
Geochemistry, Geophysics, Geosystems
 ,
13
,
624.
Tian
S.
,
Inglis
E.C.
et al
2020
.
The zirconium stable isotope compositions of 22 geological reference materials, 4 zircons and 3 standard solutions
.
Chemical Geology
 ,
555
, article
119791
,
625.
Tiepolo
M.
,
Oberti
R.
and
Vannucci
R.
2002
.
Trace-element incorporation in titanite: constraints from experimentally determined solid/liquid partition coefficients
.
Chemical Geology
 ,
191
,
105
119
,
626.
Timms
N.E.
,
Kirkland
C.L.
et al
2020
.
Shocked titanite records Chicxulub hydrothermal alteration and impact age
.
Geochimica et Cosmochimica Acta
 ,
281
,
12
30
,
627.
Tinkham
D.K.
and
Ghent
E.D.
2005
.
Estimating PT conditions of garnet growth with isochemical phase-diagram sections and the problem of effective bulk-composition
.
The Canadian Mineralogist
 ,
43
,
35
50
,
628.
Tolometti
G.D.
,
Erickson
T.M.
,
Osinski
G.R.
,
Cayron
C.
and
Neish
C.D.
2022
.
Hot rocks: constraining the thermal conditions of the Mistastin Lake impact melt deposits using zircon grain microstructures
.
Earth and Planetary Science Letters
 ,
584
, article
117523
,
629.
Tomkins
H.S.
,
Powell
R.
and
Ellis
D.J.
2007
.
The pressure dependence of the zirconium-in-rutile thermometer
.
Journal of Metamorphic Geology
 ,
25
,
703
713
,
630.
Tompkins
H.G.
,
Zieman
L.J.
,
Ibanez-Mejia
M.
and
Tissot
F.L.
2020
.
Zirconium stable isotope analysis of zircon by MC-ICP-MS: methods and application to evaluating intra-crystalline zonation in a zircon megacryst
.
Journal of Analytical Atomic Spectrometry
 ,
35
,
1167
1186
,
631.
Trail
D.
,
Watson
E.B.
and
Tailby
N.D.
2012
.
Ce and Eu anomalies in zircon as proxies for the oxidation state of magmas
.
Geochimica et Cosmochimica Acta
 ,
97
,
70
87
,
632.
Trail
D.
,
Boehnke
P.
,
Savage
P.S.
,
Liu
M.-C.
,
Miller
M.L.
and
Bindeman
I.
2018
.
Origin and significance of Si and O isotope heterogeneities in Phanerozoic, Archean, and Hadean zircon
.
Proceedings of the National Academy of Sciences of the USA
 ,
115
,
10287
10292
,
633.
Trail
D.
,
Savage
P.S.
and
Moynier
F.
2019
.
Experimentally determined Si isotope fractionation between zircon and quartz
.
Geochimica et Cosmochimica Acta
 ,
260
,
257
274
,
634.
Triebold
S.
,
von Eynatten
H.
and
Zack
T.
2012
.
A recipe for the use of rutile in sedimentary provenance analysis
.
Sedimentary Geology
 ,
282
,
268
275
,
635.
Tual
L.
,
Smit
M.A.
,
Cutts
J.
,
Kooijman
E.
,
Kielman-Schmitt
M.
,
Majka
J.
and
Foulds
I.
2022
.
Rapid, paced metamorphism of blueschists (Syros, Greece) from laser-based zoned Lu–Hf garnet chronology and LA-ICPMS trace element mapping
.
Chemical Geology
 ,
607
, article
121003
,
636.
Ushikubo
T.
,
Williford
K.H.
,
Farquhar
J.
,
Johnston
D.T.
,
van Kranendonk
M.J.
and
Valley
J.W.
2014
.
Development of in situ sulfur four-isotope analysis with multiple Faraday cup detectors by SIMS and application to pyrite grains in a Paleoproterozoic glaciogenic sandstone
.
Chemical Geology
 ,
383
,
86
99
,
637.
Valley
J.W.
2003
.
Oxygen isotopes in zircon
.
Reviews in Mineralogy and Geochemistry
 ,
53
,
343
385
,
638.
Valley
J.W.
,
Lackey
J.S.
et al
2005
.
4.4 billion years of crustal maturation: oxygen isotope ratios of magmatic zircon
.
Contributions to Mineralogy and Petrology
 ,
150
,
561
580
,
639.
Valley
J.W.
,
Cavosie
A.J.
et al
2014
.
Hadean age for a post-magma-ocean zircon confirmed by atom-probe tomography
.
Nature Geoscience
 ,
7
,
219
223
,
640.
Vanardois
J.
,
Roger
F.
et al
2022
.
Exhumation of deep continental crust in a transpressive regime: The example of Variscan eclogites from the Aiguilles-Rouges massif (Western Alps)
.
Journal of Metamorphic Geology
 ,
40
,
1087
1120
, .
641.
Vance
D.
and
Mahar
E.
1998
.
Pressure–temperature paths from P–T pseudosections and zoned garnets: potential, limitations and examples from the Zanskar Himalaya, NW India
.
Contributions to Mineralogy and Petrology
 ,
132
,
225
245
,
642.
Vance
D.
,
Strachan
R.A.
and
Jones
K.A.
1998
.
Extensional v. compressional settings for metamorphism: garnet chronometry and pressure–temperature–time histories in the Moine Supergroup, northwest Scotland
.
Geology
 ,
26
,
927
930
,
643.
van Schijndel
V.
,
Stevens
G.
,
Lana
C.
,
Zack
T.
and
Frei
D.
2021
.
De Kraalen and Witrivier greenstone belts, Kaapvaal Craton, South Africa: characterisation of the Palaeo–Mesoarchaean evolution by rutile and zircon U–Pb geochronology combined with Hf isotopes
.
South African Journal of Geology
 ,
124
,
17
36
,
644.
Vavra
G.
1990
.
On the kinematics of zircon growth and its petrogenetic significance: a cathodoluminescence study
.
Contributions to Mineralogy and Petrology
 ,
106
,
90
99
,
645.
Vavra
G.
1993
.
A guide to quantitative morphology of accessory zircon
.
Chemical Geology
 ,
110
,
15
28
,
646.
Vavra
G.
and
Schaltegger
U.
1999
.
Post-granulite facies monazite growth and rejuvenation during Permian to Lower Jurassic thermal and fluid events in the Ivrea Zone (Southern Alps)
.
Contributions to Mineralogy and Petrology
 ,
134
,
405
414
,
647.
Vavra
G.
,
Schmid
R.
and
Gebauer
D.
1999
.
Internal morphology, habit and U–Th–Pb microanalysis of amphibolite-to-granulite facies zircons: geochronology of the Ivrea Zone (Southern Alps)
.
Contributions to Mineralogy and Petrology
 ,
134
,
380
404
,
648.
Verberne
R.
,
Saxey
D.W.
,
Reddy
S.M.
,
Rickard
W.D.A.
,
Fougerouse
D.
and
Clark
C.
2019
.
Analysis of natural rutile (TiO2) by laser-assisted atom probe tomography
.
Microscopy and Microanalysis
 ,
25
,
539
546
,
649.
Verberne
R.
,
Reddy
S.M.
et al
2020
.
The geochemical and geochronological implications of nanoscale trace-element clusters in rutile
.
Geology
 ,
48
,
1126
1130
,
650.
Verberne
R.
,
Reddy
S.M.
et al
2022a
.
Dislocations in minerals: fast-diffusion pathways or trace-element traps?
Earth and Planetary Science Letters
 ,
584
, article
117517
,
651.
Verberne
R.
,
van Schrojenstein Lantman
H.W.
et al
2022b
.
Trace-element heterogeneity in rutile linked to dislocation structures: implications for Zr-in-rutile geothermometry
.
Journal of Metamorphic Geology
 ,
41
,
3
24
,
652.
Vervoort
J.
2013
. Lu–Hf dating: the Lu–Hf isotope system. In:
Rink
W.J.
and
Thompson
J.
(eds)
Encyclopedia of Scientific Dating Methods
 . Springer,
1
20
.
653.
Vho
A.
,
Rubatto
D.
,
Putlitz
B.
and
Bouvier
A.-S.
2020
.
New reference materials and assessment of matrix effects for SIMS measurements of oxygen isotopes in garnet
.
Geostandards and Geoanalytical Research
 ,
44
,
459
471
,
654.
Villaros
A.
,
Stevens
G.
and
Buick
I.S.
2009
.
Tracking S-type granite from source to emplacement: clues from garnet in the Cape Granite Suite
.
Lithos
 ,
112
,
217
235
,
655.
Villaros
A.
,
Buick
I.
and
Stevens
G.
2012
.
Isotopic variations in S-type granites: an inheritance from a heterogeneous source?
Contributions to Mineralogy and Petrology
 ,
163
,
243
257
,
656.
Villaseca
C.
,
Romera
C.M.
,
De la Rosa
J.
and
Barbero
L.
2003
.
Residence and redistribution of REE, Y, Zr, Th and U during granulite-facies metamorphism: behaviour of accessory and major phases in peraluminous granulites of central Spain
.
Chemical Geology
 ,
200
,
293
323
,
657.
Viskupic
K.
and
Hodges
K.V.
2001
.
Monazite–xenotime thermochronometry: methodology and an example from the Nepalese Himalaya
.
Contributions to Mineralogy and Petrology
 ,
141
,
233
247
,
658.
Volante
S.
,
Collins
W.J.
et al
2020a
.
Reassessing zircon–monazite thermometry with thermodynamic modelling: insights from the Georgetown igneous complex, NE Australia
.
Contributions to Mineralogy and Petrology
 ,
175
, article
110
,
659.
Volante
S.
,
Collins
W.J.
,
Pourteau
A.
,
Li
Z.X.
,
Li
J.
and
Nordsvan
A.R.
2020b
.
Structural evolution of a 1.6 Ga orogeny related to the final assembly of the supercontinent Nuna: coupling of episodic and progressive deformation
.
Tectonics
 ,
39
, article
e2020TC006162
,
660.
Volante
S.
,
Pourteau
A.
et al
2020c
.
Multiple P–T–d–t paths reveal the evolution of the final Nuna assembly in northeast Australia
.
Journal of Metamorphic Geology
 ,
38
,
593
627
,
661.
Volante
S.
,
Collins
W.J.
et al
2022
.
Spatio-temporal evolution of Mesoproterozoic magmatism in NE Australia: a hybrid tectonic model for final Nuna assembly
.
Precambrian Research
 ,
372
, article
106602
,
662.
von Blackenburg
F.
1992
.
Combined high-precision chronometry and geochemical tracing using accessory minerals: applied to the Central-Alpine Bergell intrusion (central Europe)
.
Chemical Geology
 ,
100
,
19
40
,
663.
Vry
J.K.
and
Baker
J.A.
2006
.
LA-MC-ICPMS Pb–Pb dating of rutile from slowly cooled granulites: confirmation of the high closure temperature for Pb diffusion in rutile
.
Geochimica et Cosmochimica Acta
 ,
70
,
1807
1820
,
664.
Vry
J.
,
Compston
W.
and
Cartwright
I.
1996
.
SHRIMP II dating of zircons and monazites: reassessing the timing of high-grade metamorphism and fluid flow in the Reynolds Range, northern Arunta Block, Australia
.
Journal of Metamorphic Geology
 ,
14
,
335
350
,
665.
Wafforn
S.
,
Seman
S.
,
Kyle
J.R.
,
Stockli
D.
,
Leys
C.
,
Sonbait
D.
and
Cloos
M.
2018
.
Andradite garnet U–Pb geochronology of the Big Gossan skarn, Ertsberg–Grasberg mining district, Indonesia
.
Economic Geology
 ,
113
,
769
778
,
666.
Walter
B.F.
,
Giebel
R.J.
,
Steele-MacInnis
M.
,
Marks
M.A.
,
Kolb
J.
and
Markl
G.
2021
.
Fluids associated with carbonatitic magmatism: a critical review and implications for carbonatite magma ascent
.
Earth-Science Reviews
 ,
215
, article
103509
,
667.
Walters
J.B.
and
Kohn
M.J.
2017
.
Protracted thrusting followed by late rapid cooling of the Greater Himalayan Sequence, Annapurna Himalaya, Central Nepal: insights from titanite petrochronology
.
Journal of Metamorphic Geology
 ,
35
,
897
917
,
668.
Walters
J.B.
,
Cruz-Uribe
A.M.
,
Song
W.J.
,
Gerbi
C.
and
Biela
K.
2022
.
Strengths and limitations of in situ U–Pb titanite petrochronology in polymetamorphic rocks: an example from western Maine, USA
.
Journal of Metamorphic Geology
 ,
40
,
1043
1066
,
669.
Ward
C.
,
McArthur
J.
and
Walsh
J.
1992
.
Rare earth element behaviour during evolution and alteration of the Dartmoor granite, SW England
.
Journal of Petrology
 ,
33
,
785
815
,
670.
Warren
C.J.
,
Greenwood
L.V.
,
Argles
T.W.
,
Roberts
N.M.W.
,
Parrish
R.R.
and
Harris
N.B.W.
2019
.
Garnet-monazite rare earth element relationships in sub-solidus metapelites: a case study from Bhutan
.
Geological Society, London, Special Publications
 ,
478
,
145
166
,
671.
Watson
E.B.
1979
.
Zircon saturation in felsic liquids: experimental results and applications to trace element geochemistry
.
Contributions to Mineralogy and Petrology
 ,
70
,
407
419
,
672.
Watson
E.B.
1996
.
Dissolution, growth and survival of zircons during crustal fusion: kinetic principals, geological models and implications for isotopic inheritance
.
Earth and Environmental Science Transactions of the Royal Society of Edinburgh
 ,
87
,
43
56
,
673.
Watson
E.B.
and
Harrison
T.M.
1984
.
Accessory minerals and the geochemical evolution of crustal magmatic systems: a summary and prospectus of experimental approaches
.
Physics of the Earth and Planetary Interiors
 ,
35
,
19
30
,
674.
Watson
E.B.
and
Harrison
T.M.
2005
.
Zircon thermometer reveals minimum melting conditions on earliest Earth
.
Science (New York, NY)
 ,
308
,
841
844
,
675.
Watson
E.B.
,
Wark
D.A.
and
Thomas
J.B.
2006
.
Crystallization thermometers for zircon and rutile
.
Contributions to Mineralogy and Petrology
 ,
151
,
413
433
,
676.
Wawrzenitz
N.
,
Krohe
A.
,
Baziotis
I.
,
Mposkos
E.
,
Kylander-Clark
A.R.C.
and
Romer
R.L.
2015
.
LASS U–Th–Pb monazite and rutile geochronology of felsic high-pressure granulites (Rhodope, N Greece): effects of fluid, deformation and metamorphic reactions in local subsystems
.
Lithos
 ,
232
,
266
285
,
677.
Webster
J.D.
and
Piccoli
P.M.
2015
.
Magmatic apatite: a powerful, yet deceptive, mineral
.
Elements
 ,
11
,
177
182
,
678.
Weinberg
R.F.
,
Wolfram
L.C.
,
Nebel
O.
,
Hasalová
P.
,
Závada
P.
,
Kylander-Clark
A.R.C.
and
Becchio
R.
2020
.
Decoupled U–Pb date and chemical zonation of monazite in migmatites: the case for disturbance of isotopic systematics by coupled dissolution–reprecipitation
.
Geochimica et Cosmochimica Acta
 ,
269
,
398
412
,
679.
Wetherill
G.W.
1956
.
Discordant uranium–lead ages, I
.
Eos, Transactions American Geophysical Union
 ,
37
,
320
326
,
680.
Wetherill
G.W.
1963
.
Discordant uranium–lead ages: 2. Discordant ages resulting from diffusion of lead and uranium
.
Journal of Geophysical Research (1896–1977)
 ,
68
,
2957
2965
,
681.
White
R.W.
and
Powell
R.
2002
.
Melt loss and the preservation of granulite facies mineral assemblages
.
Journal of Metamorphic Geology
 ,
20
,
621
632
,
682.
White
R.W.
,
Powell
R.
and
Johnson
T.E.
2014
.
The effect of Mn on mineral stability in metapelites revisited: new a–x relations for manganese-bearing minerals
.
Journal of Metamorphic Geology
 ,
32
,
809
828
,
683.
Whitehouse
M.J.
and
Kemp
A.I.S.
2010
.
On the difficulty of assigning crustal residence, magmatic protolith and metamorphic ages to Lewisian granulites: constraints from combined in situ U–Pb and Lu–Hf isotopes
.
Geological Society, London, Special Publications
 ,
335
,
81
101
,
684.
Whitehouse
M.J.
and
Platt
J.P.
2003
.
Dating high-grade metamorphism – constraints from rare-earth elements in zircon and garnet
.
Contributions to Mineralogy and Petrology
 ,
145
,
61
74
,
685.
Whitehouse
M.J.
,
Kemp
A.I.S.
and
Petersson
A.
2022
.
Persistent mildly supra-chondritic initial Hf in the Lewisian Complex, NW Scotland: implications for Neoarchean crust–mantle differentiation
.
Chemical Geology
 ,
606
, article
121001
,
686.
Williams
I.S.
,
Compston
W.
,
Black
L.
,
Ireland
T.
and
Foster
J.
1984
.
Unsupported radiogenic Pb in zircon: a cause of anomalously high Pb–Pb, U–Pb and Th–Pb ages
.
Contributions to Mineralogy and Petrology
 ,
88
,
322
327
,
687.
Williams
M.A.
,
Kelsey
D.E.
and
Rubatto
D.
2022
.
Thorium zoning in monazite: a case study from the Ivrea–Verbano Zone, NW Italy
.
Journal of Metamorphic Geology
 ,
40
,
1015
1042
,
688.
Williams
M.L.
and
Jercinovic
M.J.
2002
.
Microprobe monazite geochronology: putting absolute time into microstructural analysis
.
Journal of Structural Geology
 ,
24
,
1013
1028
,
689.
Williams
M.L.
,
Jercinovic
M.J.
and
Hetherington
C.J.
2007
.
Microprobe monazite geochronology: understanding geologic processes by integrating composition and chronology
.
Annual Review of Earth and Planetary Sciences
 ,
35
,
137
175
,
690.
Williams
M.L.
,
Jercinovic
M.
,
Harlov
D.
,
Budzyń
B.
and
Hetherington
C.
2011
.
Resetting monazite ages during fluid-related alteration
.
Chemical Geology
 ,
283
,
218
225
,
691.
Williams
M.L.
,
Jercinovic
M.J.
,
Mahan
K.H.
and
Dumond
G.
2017
.
Electron microprobe petrochronology
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
153
182
,
692.
Wing
B.A.
,
Ferry
J.M.
and
Harrison
T.M.
2003
.
Prograde destruction and formation of monazite and allanite during contact and regional metamorphism of pelites: petrology and geochronology
.
Contributions to Mineralogy and Petrology
 ,
145
,
228
250
,
693.
Wu
Y.-B.
,
Zheng
Y.
,
Zhang
S.
,
Zhao
Z.
,
Wu
F.
and
Liu
X.
2007
.
Zircon U–Pb ages and Hf isotope compositions of migmatite from the North Dabie terrane in China: constraints on partial melting
.
Journal of Metamorphic Geology
 ,
25
,
991
1009
,
694.
Wu
Y.-B.
,
Gao
S.
,
Zhang
H.-F.
,
Yang
S.-H.
,
Jiao
W.-F.
,
Liu
Y.-S.
and
Yuan
H.-L.
2008a
.
Timing of UHP metamorphism in the Hong'an area, western Dabie Mountains, China: evidence from zircon U–Pb age, trace element and Hf isotope composition
.
Contributions to Mineralogy and Petrology
 ,
155
,
123
133
,
695.
Wu
Y.-B.
,
Zheng
Y.-F.
,
Gao
S.
,
Jiao
W.-F.
and
Liu
Y.-S.
2008b
.
Zircon U–Pb age and trace element evidence for Paleoproterozoic granulite-facies metamorphism and Archean crustal rocks in the Dabie Orogen
.
Lithos
 ,
101
,
308
322
,
696.
Wudarska
A.
,
Wiedenbeck
M.
et al
2020
.
SIMS- and IRMS-based study of apatite reference materials reveals new analytical challenges for oxygen isotope analysis
. 22nd
EGU General Assembly Conference, 4–8 May 2020, online event, Abstracts
, 18841
.
697.
Yakymchuk
C.
2017
.
Behaviour of apatite during partial melting of metapelites and consequences for prograde suprasolidus monazite growth
.
Lithos
 ,
274
,
412
426
.
698.
Yakymchuk
C.
and
Brown
M.
2014
.
Behaviour of zircon and monazite during crustal melting
.
Journal of the Geological Society
 ,
171
,
465
479
.
699.
Yakymchuk
C.
,
Kirkland
C.L.
and
Clark
C.
2018
.
Th/U ratios in metamorphic zircon
.
Journal of Metamorphic Geology
 ,
36
,
715
737
,
700.
Yang
M.
,
Yang
Y.-H.
et al
2022
.
Natural allanite reference materials for in situ U–Th–Pb and Sm–Nd isotopic measurements by LA-(MC)-ICP-MS
.
Geostandards and Geoanalytical Research
 ,
46
,
169
203
,
701.
Yang
Y.
,
Sun
J.
,
Xie
L.
,
Fan
H.
and
Wu
F.
2008
.
In situ Nd isotopic measurement of natural geological materials by LA-MC-ICPMS
.
Chinese Science Bulletin
 ,
53
,
1062
1070
,
702.
Yang
Z.
,
Wang
B.
,
Cui
H.
,
An
H.
,
Pan
Y.
and
Zhai
J.
2015
.
Synthesis of crystal-controlled TiO2 nanorods by a hydrothermal method: rutile and brookite as highly active photocatalysts
.
The Journal of Physical Chemistry C
 ,
119
,
16905
16912
,
703.
Yurimoto
H.
,
Duke
E.
,
Papike
J.
and
Shearer
C.
1990
.
Are discontinuous chondrite-normalized REE patterns in pegmatitic granite systems the results of monazite fractionation?
Geochimica et Cosmochimica Acta
 ,
54
,
2141
2145
,
704.
Zack
T.
and
Kooijman
E.
2017
.
Petrology and geochronology of rutile
.
Reviews in Mineralogy and Geochemistry
 ,
83
,
443
467
,
705.
Zack
T.
,
Kronz
A.
,
Foley
S.F.
and
Rivers
T.
2002
.
Trace element abundances in rutiles from eclogites and associated garnet mica schists
.
Chemical Geology
 ,
184
,
97
122
,
706.
Zack
T.
,
Moraes
R.
and
Kronz
A.
2004a
.
Temperature dependence of Zr in rutile: empirical calibration of a rutile thermometer
.
Contributions to Mineralogy and Petrology
 ,
148
,
471
488
,
707.
Zack
T.
,
von Eynatten
H.
and
Kronz
A.
2004b
.
Rutile geochemistry and its potential use in quantitative provenance studies
.
Sedimentary Geology
 ,
171
,
37
58
,
708.
Zack
T.
,
Stockli
D.F.
,
Luvizotto
G.L.
,
Barth
M.G.
,
Belousova
E.
,
Wolfe
M.R.
and
Hinton
R.W.
2011
.
In situ U–Pb rutile dating by LA-ICP-MS: 208Pb correction and prospects for geological applications
.
Contributions to Mineralogy and Petrology
 ,
162
,
515
530
,
709.
Zhang
H.-X.
,
Jiang
S.-Y.
,
Yuan
F.
and
Liu
S.-Q.
2022
.
LA-(MC)-ICP-MS U–Th–Pb dating and Nd isotopes of allanite in NYF pegmatite from lesser Qingling orogenic belt, central China
.
Ore Geology Reviews
 ,
145
, article
104893
,
710.
Zhang
L.
,
Wu
J.-L.
,
Zhang
Y.-Q.
,
Yang
Y.-N.
,
He
P.-L.
,
Xia
X.-P.
and
Ren
Z.-Y.
2021
.
Simultaneous determination of Sm–Nd isotopes, trace-element compositions and U–Pb ages of titanite using a laser-ablation split-stream technique with the addition of water vapor
.
Journal of Analytical Atomic Spectrometry
 ,
36
,
2312
2321
,
711.
Zhang
L.-S.
and
Schärer
U.
1996
.
Inherited Pb components in magmatic titanite and their consequence for the interpretation of U–Pb ages
.
Earth and Planetary Science Letters
 ,
138
,
57
65
,
712.
Zhang
W.
,
Wang
Z.
et al
2019
.
Determination of Zr isotopic ratios in zircons using laser-ablation multiple-collector inductively coupled-plasma mass-spectrometry
.
Journal of Analytical Atomic Spectrometry
 ,
34
,
1800
1809
,
713.
Zhong
X.
,
Andersen
N.H.
,
Dabrowski
M.
and
Jamtveit
B.
2019
.
Zircon and quartz inclusions in garnet used for complementary Raman thermobarometry: application to the Holsnøy eclogite, Bergen Arcs, western Norway
.
Contributions to Mineralogy and Petrology
 ,
174
, article
50
,
714.
Zhong
X.
,
Moulas
E.
and
Tajčmanová
L.
2020
.
Post-entrapment modification of residual inclusion pressure and its implications for Raman elastic thermobarometry
.
Solid Earth
 ,
11
,
223
240
,
715.
Zhou
Y.
,
Yao
J.
et al
2016
.
Improving the molecular ion signal intensity for in situ liquid SIMS analysis
.
Journal of the American Society for Mass Spectrometry
 ,
27
,
2006
2013
,
716.
Zhu
X.K.
and
O'Nions
R.K.
1999
.
Monazite chemical composition: some implications for monazite geochronology
.
Contributions to Mineralogy and Petrology
 ,
137
,
351
363
,
717.
Zirner
A.L.
,
Marks
M.A.
,
Wenzel
T.
,
Jacob
D.E.
and
Markl
G.
2015
.
Rare earth elements in apatite as a monitor of magmatic and metasomatic processes: the Ilímaussaq complex, South Greenland
.
Lithos
 ,
228
,
12
22
,
718.
Zucali
M.
,
Corti
L.
,
Roda
M.
,
Ortolano
G.
,
Visalli
R.
and
Zanoni
D.
2021
.
Quantitative X-ray maps analysis of composition and microstructure of Permian high-temperature relicts in acidic rocks from the Sesia–Lanzo Zone eclogitic continental crust, Western Alps
.
Minerals
 ,
11
, article
1421
,

Figures & Tables

Fig. 2.

Representative textures for the key mineral phases described in this contribution. (a) Plane polarized light. (b) Backscattered electron (BSE) and (c) electron microprobe Mn elemental map images showing the early and late stages of fabric development in core and rims, respectively, during garnet growth. (d) Plane polarized light, (e) cathodoluminescence (CL) and (f) laser ablation ICP–MS U (ppm) elemental map images of zircon. (g) Plane polarized light, (h) BSE and (i) quantitative map of Al (a.p.f.u., atom per formula unit) images of titanite. (j) Plane polarized light, (k) BSE and (l) EBSD map images for rutile. (m) Plane and crossed polarized light, (n) BSE and (o) electron microprobe Th and Y elemental map images of monazite. (p) Plane and crossed polarized, (q) BSE and (r) LA–ICP–MS trace element map images of apatite. (s) Plane polarized light, (t) BSE and (u) X-ray elemental map images of allanite. Source: (c) Volante et al. (2020c); (f) Chew et al. (2017); (i) Walters et al. (2022); (l) Moore et al. (2020a, b); (n) Barrote et al. (2022a); (o) Volante et al. (2020c); (r) Henrichs et al. (2019); (u) Corti et al. (2020).

Fig. 2.

Representative textures for the key mineral phases described in this contribution. (a) Plane polarized light. (b) Backscattered electron (BSE) and (c) electron microprobe Mn elemental map images showing the early and late stages of fabric development in core and rims, respectively, during garnet growth. (d) Plane polarized light, (e) cathodoluminescence (CL) and (f) laser ablation ICP–MS U (ppm) elemental map images of zircon. (g) Plane polarized light, (h) BSE and (i) quantitative map of Al (a.p.f.u., atom per formula unit) images of titanite. (j) Plane polarized light, (k) BSE and (l) EBSD map images for rutile. (m) Plane and crossed polarized light, (n) BSE and (o) electron microprobe Th and Y elemental map images of monazite. (p) Plane and crossed polarized, (q) BSE and (r) LA–ICP–MS trace element map images of apatite. (s) Plane polarized light, (t) BSE and (u) X-ray elemental map images of allanite. Source: (c) Volante et al. (2020c); (f) Chew et al. (2017); (i) Walters et al. (2022); (l) Moore et al. (2020a, b); (n) Barrote et al. (2022a); (o) Volante et al. (2020c); (r) Henrichs et al. (2019); (u) Corti et al. (2020).

Fig. 3.

(a) Garnet porphyroblast indicating the kinematics of deformation (arrows) as a sigma-clast and exhibiting inclusion trails of an internal foliation. (b) Polymetamorphic garnet grain from Polnish, Scotland (see Vance et al. 1998; Cutts et al. 2009). The core (Z1) and first rim (Z2) zones preserve the same early foliation (S1) but have cleavage domes and graphite inclusions on their border. A thin, inclusion-rich rim (Z3) which has been poorly preserved is observed. (c) Representative chondrite-normalized REE patterns for different types of garnet (Blereau 2017). (d) Partitioning between zircon and garnet: (1) Buick et al. (2006); (2) Hokada and Harley (2004); (3) GL7 in Rubatto (2002); (4) GP7 in Rubatto (2002); (5) Harley et al. (2001); (6) Kelly and Harley (2005); (7) Whitehouse and Platt (2003); (8) Hermann and Rubatto (2003). (e) Partitioning array plot of zircon/garnet compared to the experimental data of Taylor et al. (2015). The axes of this plot are: x, log(DYb); y, log(Dslope) = log(DYb/DGd) (Blereau 2017). (f) Sm–Nd dating of a large garnet porphyroblast by Pollington and Baxter (2010). The various garnet rims were obtained using a microdrill (the drill lines are marked in black on the garnet). Coloured zones correspond to the coloured isochrons and age points in the plots indicate that garnet grew over c. 8 myr, from 28 to 20 Ma. (g) TW diagram showing how the grain size and the amount of ablated U-rich inclusions affect garnet U–Pb dating. (h) Garnet compositional map of Ca, Fe, Mn and Y (from top left) of sample Verpeneset from Tamblyn et al. (2022). Black and white dots represent the position of core and rim analyses, respectively. (i and j) Lu/Hf isochrons obtained from garnet (i) core and (j) rim. Uncertainties consider overdispersion of the data and isochron ages are calculated with age offset (see Tamblyn et al. 2022 for full details). Source: (a) Photo courtesy of J. Pownall; (c–d) modified after Blereau (2017); (f) adapted from Pollington and Baxter (2010), Baxter et al. (2017); (g) modified after Millonig et al. (2020).

Fig. 3.

(a) Garnet porphyroblast indicating the kinematics of deformation (arrows) as a sigma-clast and exhibiting inclusion trails of an internal foliation. (b) Polymetamorphic garnet grain from Polnish, Scotland (see Vance et al. 1998; Cutts et al. 2009). The core (Z1) and first rim (Z2) zones preserve the same early foliation (S1) but have cleavage domes and graphite inclusions on their border. A thin, inclusion-rich rim (Z3) which has been poorly preserved is observed. (c) Representative chondrite-normalized REE patterns for different types of garnet (Blereau 2017). (d) Partitioning between zircon and garnet: (1) Buick et al. (2006); (2) Hokada and Harley (2004); (3) GL7 in Rubatto (2002); (4) GP7 in Rubatto (2002); (5) Harley et al. (2001); (6) Kelly and Harley (2005); (7) Whitehouse and Platt (2003); (8) Hermann and Rubatto (2003). (e) Partitioning array plot of zircon/garnet compared to the experimental data of Taylor et al. (2015). The axes of this plot are: x, log(DYb); y, log(Dslope) = log(DYb/DGd) (Blereau 2017). (f) Sm–Nd dating of a large garnet porphyroblast by Pollington and Baxter (2010). The various garnet rims were obtained using a microdrill (the drill lines are marked in black on the garnet). Coloured zones correspond to the coloured isochrons and age points in the plots indicate that garnet grew over c. 8 myr, from 28 to 20 Ma. (g) TW diagram showing how the grain size and the amount of ablated U-rich inclusions affect garnet U–Pb dating. (h) Garnet compositional map of Ca, Fe, Mn and Y (from top left) of sample Verpeneset from Tamblyn et al. (2022). Black and white dots represent the position of core and rim analyses, respectively. (i and j) Lu/Hf isochrons obtained from garnet (i) core and (j) rim. Uncertainties consider overdispersion of the data and isochron ages are calculated with age offset (see Tamblyn et al. 2022 for full details). Source: (a) Photo courtesy of J. Pownall; (c–d) modified after Blereau (2017); (f) adapted from Pollington and Baxter (2010), Baxter et al. (2017); (g) modified after Millonig et al. (2020).

Fig. 4.

(a) Cathodoluminescence image of a complexly zoned zircon crystal from the Napier Complex (Antarctica) showing a (recrystallized) bright metamorphic core surrounded by a fine-oscillatory zoned magmatic domain, itself resorbed and overgrown by three distinct metamorphic overgrowths. (b) Igneous zircon crystals exhibiting fine-oscillatory zoning from core to rim except for a thin metamorphic outer overgrowth. (c) Representative chondrite normalized REE patterns for different types of zircon. (1) Rubatto and Hermann (2007); (2) Hoskin and Schaltegger (2003); (3) Rubatto et al. (2013); (4) Taylor et al. (2014); (5) Li et al. (2018). (d) Modelled diffusion modified M-HREE compositions expected from c. 20 µm diameter SHRIMP spot analysis on the edge of a modelled theoretical zircon. C0, original unmodified composition; CEq, equilibrated composition. (e) M-HREE spread generated by incomplete modification of a modelled zircon after different T–t conditions. (f) Concordia diagram illustrating the interpretation of zircon UPb data in terms of ages depending on whether the UPb system evolved in a closed or open system, and when disturbance occurred. (g) Graphical representation of Lu–Hf isotope evolution of distinct geological reservoirs. This graph provides a theoretical framework for the interpretation of zircon Hf isotope signatures. Source: (d) and (e) after Blereau et al. (2022).

Fig. 4.

(a) Cathodoluminescence image of a complexly zoned zircon crystal from the Napier Complex (Antarctica) showing a (recrystallized) bright metamorphic core surrounded by a fine-oscillatory zoned magmatic domain, itself resorbed and overgrown by three distinct metamorphic overgrowths. (b) Igneous zircon crystals exhibiting fine-oscillatory zoning from core to rim except for a thin metamorphic outer overgrowth. (c) Representative chondrite normalized REE patterns for different types of zircon. (1) Rubatto and Hermann (2007); (2) Hoskin and Schaltegger (2003); (3) Rubatto et al. (2013); (4) Taylor et al. (2014); (5) Li et al. (2018). (d) Modelled diffusion modified M-HREE compositions expected from c. 20 µm diameter SHRIMP spot analysis on the edge of a modelled theoretical zircon. C0, original unmodified composition; CEq, equilibrated composition. (e) M-HREE spread generated by incomplete modification of a modelled zircon after different T–t conditions. (f) Concordia diagram illustrating the interpretation of zircon UPb data in terms of ages depending on whether the UPb system evolved in a closed or open system, and when disturbance occurred. (g) Graphical representation of Lu–Hf isotope evolution of distinct geological reservoirs. This graph provides a theoretical framework for the interpretation of zircon Hf isotope signatures. Source: (d) and (e) after Blereau et al. (2022).

Fig. 5.

Titanite. (a) titanite in orthogneisses. (b) Oscillatory zoning in igneous titanite. (c) Patchy zoning in metamorphic titanite. (d) REE patterns in magmatic and metamorphic titanite. (e) Slopes of REE patterns are affected by other mineral phases growing in equilibrium with (i) igneous Precambrian and (ii) metamorphic Caledonian titanite grains. (f) Al v. Fe. (g) Th/U v. Zr-in-titanite. (h) Th/Pb v. Zr-in-titanite. (i) Al/(Al + Fe) v. total REE discrimination diagrams to distinguish between igneous (in red), recrystallized (in purple) and metamorphic (in green) titanite. (j) Schematic Tera–Wasserburg concordia diagram for interpreting U–Pb titanite data. Source: (a), (c), (d) and (e) modified after Garber et al. (2017); (b), (i) and (j) modified after Olierook et al. (2019).

Fig. 5.

Titanite. (a) titanite in orthogneisses. (b) Oscillatory zoning in igneous titanite. (c) Patchy zoning in metamorphic titanite. (d) REE patterns in magmatic and metamorphic titanite. (e) Slopes of REE patterns are affected by other mineral phases growing in equilibrium with (i) igneous Precambrian and (ii) metamorphic Caledonian titanite grains. (f) Al v. Fe. (g) Th/U v. Zr-in-titanite. (h) Th/Pb v. Zr-in-titanite. (i) Al/(Al + Fe) v. total REE discrimination diagrams to distinguish between igneous (in red), recrystallized (in purple) and metamorphic (in green) titanite. (j) Schematic Tera–Wasserburg concordia diagram for interpreting U–Pb titanite data. Source: (a), (c), (d) and (e) modified after Garber et al. (2017); (b), (i) and (j) modified after Olierook et al. (2019).

Fig. 6.

Rutile textures in (a) rutile inclusion in garnet from a HP mafic eclogite from the Sanbagawa belt, SW Japan. (b) HT rutile from an orthopyroxene–cordierite granulite from Madagascar. (c) LT rutile. Left: a BSE image of tabular rutile with a high W core (W-rt) from the Speewah carbonatite, Australia. Right: rutile grain from the Boddington Au–Cu deposit with patchy W zonation and coronitic titanite. (d) Nb–V–Cr ternary discrimination diagram for rutile. Binary plots of rutile (e) U/La v. Zr/Th and (f) Zr/Ba v. Sc/Y trace element ratios from orogenic gold deposits; (g–i) U–Pb data for LA-ICP-MS and SHRIMP analyses for porphyroblastic rutile. Weighted mean age of individual analyses for (g) LA-ICP-MS and (h) SHRIMP data. (i) U–Pb concordia diagram, ellipses representing the 2σ uncertainty; (j–l) TW plots and isochrons/errorchrons combined with Zr-in-rutile thermometry. Source: (a and b) photo courtesy of Pereira; (c) modified after Porter et al. (2020); (d) modified after Hart et al. (2018); (e) and (f) modified after Sciuba and Beaudoin (2021); (g–i) modified after Moore et al. (2020b); (j–l) modified after Bonnet et al. (2022).

Fig. 6.

Rutile textures in (a) rutile inclusion in garnet from a HP mafic eclogite from the Sanbagawa belt, SW Japan. (b) HT rutile from an orthopyroxene–cordierite granulite from Madagascar. (c) LT rutile. Left: a BSE image of tabular rutile with a high W core (W-rt) from the Speewah carbonatite, Australia. Right: rutile grain from the Boddington Au–Cu deposit with patchy W zonation and coronitic titanite. (d) Nb–V–Cr ternary discrimination diagram for rutile. Binary plots of rutile (e) U/La v. Zr/Th and (f) Zr/Ba v. Sc/Y trace element ratios from orogenic gold deposits; (g–i) U–Pb data for LA-ICP-MS and SHRIMP analyses for porphyroblastic rutile. Weighted mean age of individual analyses for (g) LA-ICP-MS and (h) SHRIMP data. (i) U–Pb concordia diagram, ellipses representing the 2σ uncertainty; (j–l) TW plots and isochrons/errorchrons combined with Zr-in-rutile thermometry. Source: (a and b) photo courtesy of Pereira; (c) modified after Porter et al. (2020); (d) modified after Hart et al. (2018); (e) and (f) modified after Sciuba and Beaudoin (2021); (g–i) modified after Moore et al. (2020b); (j–l) modified after Bonnet et al. (2022).

Fig. 7.

(a) Quantitative X-ray Map Analyser images (Q–XRMA; Ortolano et al. 2018) used to distinguish generations of metaigneous allanite grains recording different strain rates. (b) Classification diagram to distinguish magmatic from metamorphic allanite based on their Th/U ratio and the fraction of initial 206Pb (f206) values produced using machine learning (Random Forest algorithm). (c) Chondrite-normalized REE patterns of magmatic (pattern) v. metamorphic (filled) allanite: (1) Renna et al. (2007); (2) Gregory et al. (2012); (3) Zhang et al. (2022); (4) Regis et al. (2014); (5) Boston et al. (2017); (6) Vho et al. (2020); (7) Di Rosa et al. (2020). Use of geochemical data to distinguish magmatic and metamorphic allanite. Despite the degree of deformation of allanite crystals presented in (a) the same sequence of chemical zoning can be observed, with progressively deformed allanite crystals exhibiting variations of (d) Ca and (e) Ce contents in relation to LREE contents. Source: (a) from Corti et al. (2020); (b) modified after Di Rosa et al. (2020); (c) modified after Di Rosa et al. (2020).

Fig. 7.

(a) Quantitative X-ray Map Analyser images (Q–XRMA; Ortolano et al. 2018) used to distinguish generations of metaigneous allanite grains recording different strain rates. (b) Classification diagram to distinguish magmatic from metamorphic allanite based on their Th/U ratio and the fraction of initial 206Pb (f206) values produced using machine learning (Random Forest algorithm). (c) Chondrite-normalized REE patterns of magmatic (pattern) v. metamorphic (filled) allanite: (1) Renna et al. (2007); (2) Gregory et al. (2012); (3) Zhang et al. (2022); (4) Regis et al. (2014); (5) Boston et al. (2017); (6) Vho et al. (2020); (7) Di Rosa et al. (2020). Use of geochemical data to distinguish magmatic and metamorphic allanite. Despite the degree of deformation of allanite crystals presented in (a) the same sequence of chemical zoning can be observed, with progressively deformed allanite crystals exhibiting variations of (d) Ca and (e) Ce contents in relation to LREE contents. Source: (a) from Corti et al. (2020); (b) modified after Di Rosa et al. (2020); (c) modified after Di Rosa et al. (2020).

Fig. 8.

(a) Magmatic textures of monazite crystals from (1) Barrote et al. (2020); (2) Volante et al. (2020a); (3) Piechocka et al. (2017). (b) Metamorphic textures of monazite and xenotime crystals: (4) Volante et al. (2020c); (5) Photo courtesy of Cutts; (6) Cutts et al. (2018); (7) Blereau et al. (2016); (8) Laurent et al. (2018); (9) Manzotti et al. (2018); (10) Barrote et al. (2020). (c) Representative chondrite normalized REE patterns for different types of monazite and xenotime: (1) Manzotti et al. (2018); (2) Rasmussen and Muhling (2007); (3) Rubatto et al. (2013); (4) Taylor et al. (2014); (5) Buick et al. (2010). (d) U–Pb data for monazite grains from sapphirine–cordierite UHT gneisses in Norway. Monazite ages are colour-coded based on Y2O3 content. Source: (d) after Laurent et al. (2018).

Fig. 8.

(a) Magmatic textures of monazite crystals from (1) Barrote et al. (2020); (2) Volante et al. (2020a); (3) Piechocka et al. (2017). (b) Metamorphic textures of monazite and xenotime crystals: (4) Volante et al. (2020c); (5) Photo courtesy of Cutts; (6) Cutts et al. (2018); (7) Blereau et al. (2016); (8) Laurent et al. (2018); (9) Manzotti et al. (2018); (10) Barrote et al. (2020). (c) Representative chondrite normalized REE patterns for different types of monazite and xenotime: (1) Manzotti et al. (2018); (2) Rasmussen and Muhling (2007); (3) Rubatto et al. (2013); (4) Taylor et al. (2014); (5) Buick et al. (2010). (d) U–Pb data for monazite grains from sapphirine–cordierite UHT gneisses in Norway. Monazite ages are colour-coded based on Y2O3 content. Source: (d) after Laurent et al. (2018).

Fig. 9.

(a) BSE images for magmatic apatite. (b) Distinct REE patterns for apatite from different types of granitoids. (c) Discrimination diagram for apatite from various granitoids. BADR, basalt–andesite–dacite–rhyolite series. (d) Chondrite-normalized La/YbN v. YbN diagram of apatite from post-Archean granites and Archean TTGs. (e) Lithological discrimination diagram. Abbreviations: ALK, alkali-rich igneous rocks; IM, mafic I-type granitoids and mafic igneous rocks; LM, low- and medium-grade metamorphic and metasomatic; HM, partial-melts/leucosomes/high-grade metamorphic; S, S-type granitoids and high aluminum saturation index (ASI) ‘felsic’ I-types; UM, ultramafic rocks including carbonatites, lherzolites and pyroxenites. (f) Binary diagram U–Pb v. Lu–Hf ages (Ma) in apatite. (g) 87Sr/86Sr v. age diagram showing how initial apatite Sr isotope signatures can be interpreted in terms of source lithology, derived from time-integrated Rb/Sr ratios, and resulting model age (intersection between grey fans and BSE evolution). Depleted MORB (Mid-ocean ridge basalt) mantle (DMM). Source: (b) and (c) modified after Bruand et al. (2020); (d) modified after Antoine et al. (2020); (e) modified after O'Sullivan et al. (2020); (f) rock lithologies following discrimination diagram from O'Sullivan et al. (2020); (g) modified after Emo et al. (2018).

Fig. 9.

(a) BSE images for magmatic apatite. (b) Distinct REE patterns for apatite from different types of granitoids. (c) Discrimination diagram for apatite from various granitoids. BADR, basalt–andesite–dacite–rhyolite series. (d) Chondrite-normalized La/YbN v. YbN diagram of apatite from post-Archean granites and Archean TTGs. (e) Lithological discrimination diagram. Abbreviations: ALK, alkali-rich igneous rocks; IM, mafic I-type granitoids and mafic igneous rocks; LM, low- and medium-grade metamorphic and metasomatic; HM, partial-melts/leucosomes/high-grade metamorphic; S, S-type granitoids and high aluminum saturation index (ASI) ‘felsic’ I-types; UM, ultramafic rocks including carbonatites, lherzolites and pyroxenites. (f) Binary diagram U–Pb v. Lu–Hf ages (Ma) in apatite. (g) 87Sr/86Sr v. age diagram showing how initial apatite Sr isotope signatures can be interpreted in terms of source lithology, derived from time-integrated Rb/Sr ratios, and resulting model age (intersection between grey fans and BSE evolution). Depleted MORB (Mid-ocean ridge basalt) mantle (DMM). Source: (b) and (c) modified after Bruand et al. (2020); (d) modified after Antoine et al. (2020); (e) modified after O'Sullivan et al. (2020); (f) rock lithologies following discrimination diagram from O'Sullivan et al. (2020); (g) modified after Emo et al. (2018).

Fig. 10.

Synoptic P–T evolutions of a metapelite in different hypothetical scenarios with associated types of zircon, monazite, xenotime, garnet growth/breakdown textures and potential REE patterns. (a) Singular clockwise P–T evolution: (i) inherited zircon; (ii) primary garnet and recrystallization of inherited zircon; (iii) relict zircon core, initial monazite growth; (iv) growth of garnet rim and peritectic garnet; (v) neocrystallized ‘soccer ball’ zircon, additional monazite and recrystallization of zircon; (vi) garnet breakdown; (vii) melt crystallization; (viii) late fluid modification. (b) A residual protolith along a clockwise polymetamorphic P–T path: (i) recrystallization of zircon; (ii) prograde monazite growth in the presence of relict garnet; (iii) at the elevated solidus, minor melting causes additional monazite but no new zircon growth; (iv) garnet breakdown, final monazite growth and complete ‘ghost zoning’ to recrystallized zircon. After a localized hydrous retrograde event, the hydrated areas follow a secondary P–T path; (v) melt modifies zircon and monazite to relict cores stable with relict garnet grains; (vi) new zircon and monazite; (vii) final melt crystallization. (c) Synthetic Tera–Wasserburg plot with interpretation based on the nature of analysed textures; § enlarged inset of younger ages. Source: (a–c) modified after Blereau (2017).

Fig. 10.

Synoptic P–T evolutions of a metapelite in different hypothetical scenarios with associated types of zircon, monazite, xenotime, garnet growth/breakdown textures and potential REE patterns. (a) Singular clockwise P–T evolution: (i) inherited zircon; (ii) primary garnet and recrystallization of inherited zircon; (iii) relict zircon core, initial monazite growth; (iv) growth of garnet rim and peritectic garnet; (v) neocrystallized ‘soccer ball’ zircon, additional monazite and recrystallization of zircon; (vi) garnet breakdown; (vii) melt crystallization; (viii) late fluid modification. (b) A residual protolith along a clockwise polymetamorphic P–T path: (i) recrystallization of zircon; (ii) prograde monazite growth in the presence of relict garnet; (iii) at the elevated solidus, minor melting causes additional monazite but no new zircon growth; (iv) garnet breakdown, final monazite growth and complete ‘ghost zoning’ to recrystallized zircon. After a localized hydrous retrograde event, the hydrated areas follow a secondary P–T path; (v) melt modifies zircon and monazite to relict cores stable with relict garnet grains; (vi) new zircon and monazite; (vii) final melt crystallization. (c) Synthetic Tera–Wasserburg plot with interpretation based on the nature of analysed textures; § enlarged inset of younger ages. Source: (a–c) modified after Blereau (2017).

Fig. 11.

Synoptic P–T evolution diagram including reactions of key petrochronometers such as garnet, allanite, monazite, rutile, apatite, xenotime and titanite. Source: modified after Manzotti et al. (2018) including key mineral reactions from Boston et al. (2017); Fumes et al. (2022); Manzotti et al. (2022).

Fig. 11.

Synoptic P–T evolution diagram including reactions of key petrochronometers such as garnet, allanite, monazite, rutile, apatite, xenotime and titanite. Source: modified after Manzotti et al. (2018) including key mineral reactions from Boston et al. (2017); Fumes et al. (2022); Manzotti et al. (2022).

Fig. 12.

Hypothetical temperature–time paths for two distinct intrusions with one having cooled rapidly and the other one slowly. Note that closure temperatures do not overlap between techniques and minerals for illustrative purposes only. Also, cooling paths are presented as straight lines which may not accurately depict actual cooling paths.

Fig. 12.

Hypothetical temperature–time paths for two distinct intrusions with one having cooled rapidly and the other one slowly. Note that closure temperatures do not overlap between techniques and minerals for illustrative purposes only. Also, cooling paths are presented as straight lines which may not accurately depict actual cooling paths.

Table 1.

Synoptic table of most used tools for zircon investigations

Tool*InterestExample of reference
MorphologyNature of parental magma (alkaline, peraluminous or calc-alkaline)Pupin (1980); Vavra (1993) 
CL and BSE imagesPristineness of crystal lattice and crystallization conditions (igneous, metamorphic or weathered)Vavra (1990); Hanchar and Miller (1993); Corfu et al. (2003); Guitreau et al. (2018) 
Inclusions (mineral and melt)Nature of parental magma and/or fluids (for primary inclusions either as melt or mineral)Thomas et al. (2003); Hopkins et al. (2008); Bell et al. (2015) 
EBSDOrientation and deformation of crystalsTolometti et al. (2022); Cox et al. (2022) 
Trace element concentrationsNature of parental magma, magmatic or metamorphic origin and pristinenessGrimes et al. (2007, 2015); Laurent et al. (2021); Guitreau et al. (2022) 
(Ti)Crystallization temperatureWatson et al. (2006); Fu et al. (2008) 
(K) and (Ca)Primitive v. altered zirconBouvier et al. (2012); McCubbin et al. (2016) 
Th/UZircon crystallization conditions (e.g. magmatic v. metamorphic origin) and weatheringVavra et al. (1999); Kirkland et al. (2015); Yakymchuk et al. (2018); Guitreau and Flahaut (2019) 
Ce anomalyRedox state of parental magmaSmythe and Brenan (2016); Trail et al. (2012) 
U–PbDating (absolute age determination)Wetherill (1956, 1963); Davis et al. (2003 and reference therein); Schoene (2014) 
Lu–HfSource tracing (e.g. mantle or crust)Patchett (1983); Amelin et al. (2000); Gerdes and Zeh (2009); Guitreau et al. (2012) 
OSource tracing (e.g. involvement of clay-rich sediments)Valley (2003); Iizuka et al. (2013) 
Si isotopesSource tracing, parental magma SiO2 content and crystallization historyTrail et al. (2018); Guitreau et al. (2020, 2022)
Zr isotopesMagma crystallization historyIbañez-Mejia and Tissot (2019); Guo et al. (2020); Tompkins et al. (2020) 
Tool*InterestExample of reference
MorphologyNature of parental magma (alkaline, peraluminous or calc-alkaline)Pupin (1980); Vavra (1993) 
CL and BSE imagesPristineness of crystal lattice and crystallization conditions (igneous, metamorphic or weathered)Vavra (1990); Hanchar and Miller (1993); Corfu et al. (2003); Guitreau et al. (2018) 
Inclusions (mineral and melt)Nature of parental magma and/or fluids (for primary inclusions either as melt or mineral)Thomas et al. (2003); Hopkins et al. (2008); Bell et al. (2015) 
EBSDOrientation and deformation of crystalsTolometti et al. (2022); Cox et al. (2022) 
Trace element concentrationsNature of parental magma, magmatic or metamorphic origin and pristinenessGrimes et al. (2007, 2015); Laurent et al. (2021); Guitreau et al. (2022) 
(Ti)Crystallization temperatureWatson et al. (2006); Fu et al. (2008) 
(K) and (Ca)Primitive v. altered zirconBouvier et al. (2012); McCubbin et al. (2016) 
Th/UZircon crystallization conditions (e.g. magmatic v. metamorphic origin) and weatheringVavra et al. (1999); Kirkland et al. (2015); Yakymchuk et al. (2018); Guitreau and Flahaut (2019) 
Ce anomalyRedox state of parental magmaSmythe and Brenan (2016); Trail et al. (2012) 
U–PbDating (absolute age determination)Wetherill (1956, 1963); Davis et al. (2003 and reference therein); Schoene (2014) 
Lu–HfSource tracing (e.g. mantle or crust)Patchett (1983); Amelin et al. (2000); Gerdes and Zeh (2009); Guitreau et al. (2012) 
OSource tracing (e.g. involvement of clay-rich sediments)Valley (2003); Iizuka et al. (2013) 
Si isotopesSource tracing, parental magma SiO2 content and crystallization historyTrail et al. (2018); Guitreau et al. (2020, 2022)
Zr isotopesMagma crystallization historyIbañez-Mejia and Tissot (2019); Guo et al. (2020); Tompkins et al. (2020) 

*BSE, Backscattered electron; CL, cathodoluminescence; EBSD, electron backscatter diffraction.

or Create an Account

Close Modal
Close Modal