Skip to main content
  • Published:

Three generations of monazite in Austroalpine basement rocks to the south of the Tauern Window: evidence for Variscan, Permian and Eo-Alpine metamorphic events

Abstract

Three monazite generations were observed in garnet-bearing micaschists from the Schobergruppe in the basement to the south of the Tauern Window, Eastern Alps. Low-Y monazite of Variscan age (321 ± 14 Ma) and high-Y monazite of Permian age (261 ± 18 Ma) are abundant in the mica-rich rock matrix and in the outer domains of large garnet crystals. Pre-Alpine monazite commonly occurs as polyphase grains with low-Y Variscan cores and high-Y Permian rims. Monazite of Eo-Alpine age (112 ± 22 Ma) is rarer and was observed as small, partly Y-enriched grains (3 wt. % Y2O3) in the rock matrix and within garnet. Based on monazite-xenotime thermometry, Y + HREE values in monazite indicate minimum crystallization conditions of 500 °C during the Variscan and 650 °C for the Permian and Alpine events, respectively. Garnet zoning and thermobarometric calculations with THERMOCALC 3.21 record an amphibolite facies, high-pressure stage of ~600 °C/13–16 kbar, followed by a thermal maximum at 650–700 °C and 6–9 kbar. The Eo-Alpine age for these two events is supported by inclusions of Cretaceous monazite in the garnet domains used for thermobarometric constraints and through the high growth temperatures of Eo-Alpine monazite, which is consistent with that of the thermal maximum (~700 °C). The age and growth conditions of a few Mn-rich garnet cores, sporadically present within Eo-Alpine garnet, are unclear because inclusions of monazite, plagioclase and biotite necessary for thermobarometric- and age constraints are absent. However, based on monazite thermometry, Permian and Variscan metamorphic conditions were high enough for the growth of pre-Alpine garnet. The formation of Variscan garnet and its later resorption, plus Y-release, would also explain the high Y in Permian monazite, which cannot originate from preexisting Variscan monazite only. Monazite of Variscan, Permian and/or Eo-Alpine ages were also observed in other garnet-bearing micaschists from the Schobergruppe. This suggests that the basement of the Schobergruppe was overprinted by three discrete metamorphic events at conditions of at least lower amphibolite facies. While the Variscan event affected all parts of this basement, the younger events are more pronounced in its structurally lower units.

1 Introduction

The Austroalpine basement immediate to the south of the Tauern Window within the Northern-Defereggen-Petzeck Group (the Schobergruppe as a part of it) was overprinted during the Eo-Alpine orogeny at variable metamorphic conditions (e.g., Schuster et al. 2004; Schmid et al. 2004; Schulz et al. 2005). According to some authors (Hoinkes et al. 1999; Linner et al. 2000) and as outlined in the metamorphic map of the Alps (Frey 1999a, b; Neubauer et al. 1999; Hoinkes et al. 1999; Schmid et al. 2004), even Eo-Alpine high-pressure amphibolite facies conditions were reached.

However it is difficult to assess the pre-Alpine metamorphic evolution in this part of the basement, in particular the Permian and Variscan events, since the older metamorphic assemblages are mostly overprinted and common geochronometers largely re-equilibrated. Accessory phases like monazite or zircon are usually refractory even at high-grade metamorphic overprints (e.g. Rubatto et al. 2006) and may therefore better record earlier thermal events than common minerals. However, although highly refractory, monazite shows certain reactivity as well and can crystallize at almost every PT conditions, including very low grades (Read et al. 1987; Evans et al. 2002; Rasmussen et al. 2001; Rasmussen and Muhling 2007, 2009; Wan et al. 2007; Wilby et al. 2007; Biševac et al. 2011). As a consequence, several generations of monazite may be observed within a single sample or even a single grain (e.g., DeWolf et al. 1993; Cocherie et al. 1998; Pyle 2006; Spear et al. 2008; Petrík and Konečný 2009; Martins et al. 2009; Schulz and von Raumer 2011).

Apart from geochronological aspects, the chemical composition of monazite is a valuable petrogenetic indicator. The incorporation of Y and HREE in monazite follows a solvus and can be used as a thermometer, the so called monazite–xenotime miscibility gap thermometer (Gratz and Heinrich 1997; Heinrich et al. 1997; Pyle et al. 2001). In addition, Y-zoning in monazite may be correlated with distinct stages of garnet growth or breakdown (Pyle and Spear 2003), as garnet highly fractionates this element. Y will thus be less available for monazite when garnet is crystallizing simultaneously and the other way around when garnet breaks down (e.g. Zhu and O`Nions 1999a, b; Pyle et al. 2001; Krenn et al. 2009; Martins et al. 2009). Therefore the Y distribution in monazite can be useful in clarifying if garnet growth occurred prior to, at the same time as or after monazite.

In this study, we present data from polymetamorphic Schobergruppe rocks hosting distinct generations of monazite with specific Y contents. Three monazite generations have been identified and corroborated by comparison with other samples from this region. Variable Y contents evidence the existence of pre-Alpine metamorphic assemblages with simultaneous crystallization of garnet (e.g., Gaidies et al. 2008; Pyle and Spear 2003).

2 Geological setting

The studied samples are micaschists from the Austroalpine basement of the Schobergruppe (provinces Carinthia and Tyrol in Austria; Fig. 1a), which in turn belongs to the Northern-Defereggen-Petzeck Group (Fig. 1b). This part of the Austroalpine basement nappe is located to the south of the central to eastern Tauern Window (Schulz 1993a, b). The samples were taken close to eclogitic amphibolites (Fig. 1b, d) described by Schulz (1993a) and Schulz et al. (2005). Other micaschist samples from the Schobergruppe or from adjacent parts of the Austroalpine basement (Fig. 1b, e) have been considered for age control. Some of them were located in areas free of Alpine overprint, according to available Carboniferous K–Ar and Rb–Sr mica ages (Fig. 1c).

Fig. 1
figure 1

Geological map showing the Schobergruppe (SG) within the tectonic framework of the Eastern Alps (a) and in more detail with sample location (b) (modified after Schulz et al. (2005, 2008). (c) Distribution of K–Ar and Rb–Sr mica ages to the south of the Tauern Window (compiled from Borsi et al. (1978); Hoinkes et al. (1999); Schuster et al. (2001)). Cross section through the Schobergruppe and Prijakt eclogitic amphibolites (d) and through the Austroalpine basement to the south of Hopfgarten (e). Sample locations and Th–U–Pb chemical ages of monazite are also shown

As in many crystalline complexes in the Eastern Alps, the Northern-Defereggen-Petzeck Group was overprinted by Alpine metamorphism during Cretaceous times (e.g. Schuster et al. 2004; Schulz et al. 2008). This metamorphic overprint reached greenschist to high-pressure amphibolite facies conditions (Exner 1962; Oxburgh et al. 1966; Troll and Hölzl 1974; Troll et al. 1976, 1980; Schulz 1993a; Linner et al. 2000). In high-grade zones, relicts of the pre-existing Devonian-Carboniferous Variscan metamorphism are often blurred.

So far, Variscan amphibolite facies metamorphism has been reported mainly from the basement south and southeast to the Schobergruppe (Schulz 1990, 1993b; Schuster et al. 2001; Schulz et al. 2005, 2008; Steidl et al. 2009, 2010a, b), where mica cooling ages do not record significant Alpine overprint (Fig. 1c). From a tectonic point of view, the area to the south of the so-called Defereggen-Antholz-Vals shear zone (DAV) belongs to an upper part of the Austroalpine basement, while the Schobergruppe represents a structurally lower part.

Variscan ages in metapelitic basement rocks south to the Tauern Window are generally rare. Schulz et al. (2005, 2008) for instance report Variscan monazite from the Schobergruppe and Steidl et al. (2010a, b) Variscan monazite from the Michlbach Complex, in the eastern Defereggen Alps, south to the DAV. Geochronological data to the east of the Schobergruppe range from 90 to 310 Ma (Hoke 1990; Schuster et al. 2001) and are interpreted as a mixture of Variscan and Eo-Alpine ages. Similarly, Permian ages in metapelitic rocks were also interpreted as mixed ages (Carboniferous to Cretaceous) or were ascribed to a slow Variscan cooling history (e.g. Borsi et al. 1978). So far, traces of a Permian event in the Schobergruppe seem to be restricted to the emplacement of pegmatite bodies and related HT/LP overprint nearby. Although Permian pegmatites are very common in the basement of the Schobergruppe (Bücksteeg 1999), it is not clear if this area was also affected by a Permian HT/LP event as is the case in other crystalline complexes south and southeast to the Schobergruppe (Schuster and Stüwe 2008).

3 Samples

Monazite and garnet were studied in detail in three garnet-bearing micaschist samples (520, 527 and Alk 8a) located in the vicinity of eclogitic amphibolites. The rocks are strongly foliated with modal abundances of approximately 5–10 vol. % garnet, 20–25 % quartz, 15–20 % plagioclase, 15–20 % muscovite, 10–15 % biotite, 3–5 % staurolite, 1–5 % chlorite and <3 % kyanite. Accessory phase are apatite, zircon, monazite and xenotime. The prevailing Ti-phase is ilmenite (0.5–2 vol. %).

Garnet is variable in size and shape, ranging from small (<300 μm) isometric, partly, rounded grains up to very large (several mm) eu- to subhedral crystals (Fig. 2). Under the microscope, garnet appears relatively homogeneous and not polyphase with little marginal breakdown and alteration. Inclusions of ilmenite, quartz, biotite, muscovite and plagioclase are common everywhere in the garnet. Garnet was also observed around staurolite (Fig. 2) or occurs intergrown with staurolite and/or mica-rich batches with remnants of staurolite (Fig. 2). Staurolite forms up to 5 mm large, eu- to subhedral crystals, sometimes twinned, sometimes elongated parallel to the foliation. In some places, staurolite occurs together with kyanite (see below). Kyanite is scarce and occurs as small grains or clusters (partly fibroblastic) associated with mica and/or staurolite.

Fig. 2
figure 2

Microphotographs showing garnet crystals associated with staurolite from sample 520 (a) and 527 (b)

Monazite from other garnet-bearing micaschists in this area was studied by Schulz et al. (2005, 2008). The corresponding samples were taken in the vicinity and below the eclogitic amphibolites to the south of the Oligocene Defereggen-Antholz-Vals shear zone (Fig. 1d, e). The samples were situated within and above a pegmatite-rich zone and within the monotonous metapsammopelitic Defereggen Group. The age of metamorphism remained unclear and the peak PT conditions were considered pre-Alpine in age (Schulz et al. 2005, 2008). This is why we have revisited this area and resampled metapelites with zoned monazite as well as monazite inclusions in garnet, in order to establish correlations between monazite- and garnet growth stages.

4 Monazite

4.1 Chemical Th–U–Pb monazite dates

Backscattered electron imaging (BSE), phase identification with an energy-dispersive system (EDS) as well as dating and chemical analyses using the wavelength-dispersive system (WDS) were carried out with a JEOL-JX8600 electron microprobe following the procedures described in Krenn et al. (2008) and Krenn and Finger (2004). Single monazite dates and weighted average ages (Table 1) were calculated utilizing the method of Montel et al. (1996) and isoplot 2.1 (Ludwig 2001) considering the analytical 2 sigma errors on Th, U and Pb measurements. The analytical Pb-errors range from 0.010 to 0.016 for a Pb dwell time of 160–400 s.

Table 1 Th, U, Pb concentration, Th* values, ages and 2 sigma errors of monazites

The statistical distribution of measured monazite ages shows three peaks corresponding to Carboniferous (Variscan), Permian and Cretaceous (Eo-Alpine) times, respectively (Fig. 3). Data cluster around 300–330, 250–280 and 70–135 Ma, respectively, whereas a few dates between 220 and 250 Ma and ~300 Ma are interpreted as a mixture of ages due to the polyphase nature of monazite (see below). The weighted average dates per sample are listed in Table 1. In the Th* versus total-Pb diagram after Suzuki et al. (1991), pre-Alpine and Alpine monazite analyses arrange themselves along three trendlines (Fig. 4), which provide isochron ages of 321 ± 14 Ma, 261 ± 18 Ma and 112 ± 22 Ma, respectively. The Variscan trendline is characterized by a slope of 0.01429x (±0.0006x), an interception value of ~0 and a MSWD value of 0.4; the Permian regression defines a slope of 0.0117x (±0.0008x), an intercept value of ±0.0007 and a MSWD value of 0.3 and the Alpine trendline has a slope of 0.0055x (±0.0013x), an intercept value of 0.001 and a MSWD value of 0.24.

Fig. 3
figure 3

Frequency diagram of monazite ages in samples 520, 527 and Alk 8

Fig. 4
figure 4

Th* vs. Pb diagram after Suzuki et al. (1991) showing the trendlines defined through Alpine, Permian and Variscan monazite. Age bars on the right site are isochrones forced through 0

Variscan, Permian and/or Eo-Alpine (Cretaceous) monazite were also observed in other samples from the Schobergruppe and from the Defereggen Group, south to the DAV (Schulz et al. 2005, 2008). Sample HPr10 shows Permian and abundant Cretaceous monazite ages (Fig. 1b, d; Table 2), samples Alk2, Sti14 and P24 preferentially Permian ages and samples Alk 8, 839, 431b and 750b Variscan (Carboniferous) monazite ages (Fig. 1d, e; Table 2). It is important to note that it was beyond the scope of these earlier studies to analyze all monazite grains in a sample. Therefore, it is likely that other monazite generations would be found as well if studied in more detail and the absence of a monazite age in these samples should not be over- or misinterpreted.

Table 2 Th-U–Pb CHIME model ages of monazite from other garnet-bearing mica schists from the Schobergruppe and adjacent Austroalpine basement, as reported in Schulz et al. (2005)

4.2 Morphology and chemistry of monazite

In all three samples investigated in this work, Variscan and Permian monazite form morphologically indistinguishable, relatively large eu- to subhedral grains (ca. 20–150 μm) (Fig. 5a, b). Some crystals show straight grain boundaries, others are irregular and embayed (Fig. 5b, c). Monazite grains were observed in the matrix as well as inclusions in staurolite and in the outermost domains of large garnet crystals (Fig. 5a), but not in the cores of the latter.

Fig. 5
figure 5

Backscattered electron images showing monazite from samples 520, 527 and Alk 8. Polyphase monazite (Variscan core, Permian rim) enclosed in the outer domain of garnet (a, b). Polyphase matrix monazite with a Variscan core (labeled as analysis 1), Variscan rim (analysis 2) and an outermost Permian rim (analysis 3) (c). Polyphase matrix monazite with a Variscan core and Permian rim (d, e). Alpine monazite enclosed in garnet. (h) Alpine monazite in the staurolite-kyanite bearing matrix (f, g). Chemistry and ages of monazite are listed in Tables 1, 2, 3, and garnet compositions are shown in Figs. 10 and 11

Approximately 1/3–1/2 of all monazite grains display a characteristic zoning in BSE images with bright cores (enriched in thorium; up to 16 wt.  % ThO2) and darker rims with lower Th contents (Table 3). Dating of these rim zones revealed that most of them are Permian in age, although both Variscan and Permian rims were locally observed (Fig. 5c). In Fig. 5c, a bright Th-rich Variscan core is surrounded by an optically darker Variscan rim, which in turn is followed by an outermost Permian rim (Fig. 5c; Table 3). Variscan domains (cores and rims) are low in Y (<1 wt.  % Y2O3), while Y2O3 values in Permian rims are systematically (0.5–1.5 wt.  %) higher than in the corresponding Variscan core (Table 3). In average, Variscan monazite shows higher Th- and lower Y contents than Permian monazite (3–16 wt.  % ThO2 and 0.1–1wt.  % Y2O3 vs. 1–8 wt.  % ThO2 and 0.2–2.5 wt.  % Y2O3; Figs. 6, 7). In addition, Variscan monazite provides a slight positive covariation between Th and Y, while Th and Y in Permian monazite are negatively correlated or not correlated at all (Fig. 7).

Table 3 Selected microprobe analyses of monazites
Fig. 6
figure 6

Th versus Y trends of Variscan, Permian and Alpine monazite

Fig. 7
figure 7

Chemical composition of Variscan and Permian monazite

Eo-Alpine monazite occurs in small and unzoned grains (<10 μm), which, in some places, arrange themselves in clusters (Fig. 5f, g). Monazite in sample Alk8 was only observed in the matrix, while monazite in samples 520 and 527 was found in the matrix as well as enclosed in small, optically homogeneous and widely unaltered garnet crystals (Fig. 5f, g). Eo-Alpine monazite is not texturally associated with allanite (as is in sample HPr10 from Schulz et al. 2005, 2008) or with pre-Alpine monazite, but was found together with or nearby xenotime (Fig. 5h). Eo-Alpine monazite yields ca. 1.5–6 wt.  % ThO2 and shows a considerable Y-variation from ca. 1 wt. % Y2O3 (as is the case for monazite in garnet) up to 2.9 wt. % Y2O3 (Fig. 6; Table 3). The highest Y-values were observed in a matrix at the boundary of staurolite-kyanite-muscovite aggregates and in the close vicinity to xenotime crystals (Fig. 5). Xenotime is common in all samples and occurs as small grains or in clusters in the matrix and enclosed in garnet (see below).

4.3 Monazite-xenotime thermometry

Minimum growth temperatures of monazite were obtained utilizing monazite-xenotime miscibility gap thermometers. Figure 8 shows the experimentally calibrated 2 kbar thermometer curve of Gratz and Heinrich (1997), the empirical thermometer of Heinrich et al. (1997), which was obtained on low-pressure rocks and the empirical thermometer of Pyle et al. (2001) calibrated on high pressure rocks. Also shown is the experimentally 2 kbar curve for huttonite-bearing monazite with ca. 10 % huttonite (Seydoux-Guillaume et al. 2002). This thermometer curve runs sub-parallel to the 2 kbar curve of Gratz and Heinrich (1997) at ca. 30–50 °C lower temperatures. The Gratz and Heinrich (1997) thermometer was calibrated in the binary CePO4–YPO4 system disregarding the stronger Y-uptake of huttonite-bearing monazite.

Fig. 8
figure 8

Xenotime versus temperature diagram showing the published monazite-xenotime miscibility gap thermometer curves and the minimum formation temperatures of Variscan, Permian and Alpine monazite under study

The highest xenotime value observed in Permian monazite (~8 mol %) intersects the low-pressure curves of Heinrich et al. (1997); Gratz and Heinrich (1997) and Seydoux-Guillaume et al. (2002) between ca. 650 and 700 °C (Fig. 8). The highest xenotime value observed in Eo-Alpine monazite (ca. 10 mol %) intersects the thermometer curve of Pyle et al. (2001) at temperatures of ca. 650 °C, those calibrated for low-pressure rocks at ~700 °C. Maximum xenotime contents observed in Variscan monazite (5 mol %) implies growth temperatures between 450 °C (Pyle et al. 2001) and 550 °C (Heinrich et al. 1997). Table 4 lists the (minimum) formation temperatures calculated utilizing the thermometer-functions given by Gratz and Heinrich (1997); Heinrich et al. (1997) and Pyle et al. (2001).

Table 4 Xenotime values of monazite and corresponding monazite-xenotime miscibility gap temperatures

5 Major phase composition and geothermobarometry

Most garnet crystals, including those with inclusions of Eo-Alpine and Permian monazite (Fig. 5a, f) and those surrounding staurolite (Fig. 9b), show a simple core to rim zoning with a rimward decrease of MnO and CaO (1–0.1 wt. %; 3–0.1 wt. %) and increase of MgO (2.5–4 wt. %; Fig. 10). However, a few of the larger garnet crystals also yield a Mn-enriched core (plateau) with 3–4 wt. % MnO and 1–2 wt. % CaO and MgO (Fig. 11), followed by a rim domain, which is characterized by an abrupt decrease of MnO and increase of CaO and MgO (up to ca. 3–4 wt. %). The composition of the latter rims is similar to the garnet crystals shown in Fig. 10. A narrow retrogressive rim (max. 100–200 μm) is common in all investigated garnet grains.

Fig. 9
figure 9

Backscattered electron images of garnet from samples 520, 527 and Alk 8. Euhedral garnet with a simple zoning and a narrow alteration rim (a). Garnet around staurolite shown in Fig. 2b (b). Large garnet with diffuse zoning and inclusion of Permian monazite in its outer domains (c). Large garnet with diffuse zoning and inclusion of xenotime. Chemical profiles are shown in Figs. 10 and 11 (d)

Fig. 10
figure 10

Chemical composition of low Mn-garnet, shown in Figs. 5a, f and 9a, b

Fig. 11
figure 11

Chemical composition of high Mn-garnet grains, shown in the Fig. 9c and d

Biotite is chemically homogeneous with a Mg/(Mg + Fe) of 0.4–0.5 and Ti contents from 1.5 to 2 wt % TiO2, whether in the matrix or enclosed in garnet. Muscovite usually shows a weak diffuse zoning with slightly varying Na- and Ti values (1.3–2.5 wt. % Na2O and 0.6–1.5 wt. % TiO2). Plagioclase composition ranges from An2 in the core to An12 at the rim.

An independent set of reactions obtained with the THERMOCALC software 3.21 (Holland and Powell 1998) yields PT estimates of ca. 600–700 °C and 6–9 kbar for the assemblage biotite, muscovite, plagioclase-core and Mg-enriched garnet domains, and 550–600 °C and 14–16 kbar for the assemblage biotite, muscovite, plagioclase-rim and grossular-enriched garnet domains. We are aware that pressure estimates suffer from low An-contents of the plagioclase (e.g. Todd 1998). However similar PT estimates of 650–700 °C/8–10 kbar and 550–650 °C/12–14 kbar were obtained by Schulz et al. (2005) for other low- and high-Ca metapelites from the Schobergruppe. These authors used the garnet-biotite thermometer of Bhattacharya et al. (1992) in combination with the garnet-muscovite-biotite-plagioclase barometer of Holland and Powell (1990) with updated activity models from Ganguly et al. (1996) and Powell and Holland (1993). The PT estimates obtained from metapelites overlap with those obtained from eclogites and eclogitic amphibolites (Schulz 1993a; Schulz et al. 2008), which are intimately associated with the studied metapelites (see Fig. 1b and discussion below).

6 Discussion

6.1 Formation of pre-Alpine monazite

The occurrence of pre-Alpine monazite in the Schobergruppe samples is a clear evidence that parts of this area were pervasively metamorphosed during Variscan and Permian times. We conclude from the monazite-xenotime miscibility gap thermometry that the Variscan and Permian events reached temperatures of at least 500 and 650 °C, respectively (Fig. 8; Table 4). However, many Variscan monazite grains show xenotime contents of <5 mol % possibly because they crystallized at lower temperatures along the prograde Variscan metamorphic path or because they failed to attain maximum Y contents. Growth of monazite at temperatures much below 500 °C is not consistent with its high Th contents (up to 16 wt. % ThO2) and with its large, euhedral crystals. Low-T monazite usually has much lower Th contents and/or its crystals are smaller and sub- to anhedral (e.g., Read et al. 1987; Rasmussen et al. 2001; Evans et al. 2002; Kryza et al. 2004; Wilby et al. 2007; Krenn et al. 2008; Del Rio et al. 2009; Biševac et al. 2011; Čopjaková et al. 2011). In addition monazite often occurs in greater abundances at upper greenschist to lower amphibolite facies conditions in many metapelitic rocks (Kingsbury et al. 1993; Lanzirotti and Hanson 1996; Wing et al. 2003).

It is likely that Variscan monazite with lower Y contents grew at a stage when Y was trapped in other phases like garnet. Garnet has indeed a high affinity for Y (Pyle and Spear 1999; Pyle et al. 2001) and this element will be less available for monazite if garnet growth occurs before monazite; alternatively, Y will be released and made available when garnet breaks down. Thus low Y contents in some Variscan monazite grains could be explained by its uptake in garnet porphyroblasts, which crystallized at Variscan times prior to monazite. Subsequent garnet resorption coupled to Y release, could be responsible for the much higher Y contents observed in Permian monazite (Fig. 7), which cannot come from the Variscan low-Y monazite generation alone. It is also likely that xenotime crystallized during this period of garnet resorption (e.g. Pyle et al. 2001). Possibly some of the xenotime grains observed within garnet formed together with Permian high-Y monazite.

The replacement of Variscan, low-Y monazite by Permian, high-Y monazite might be explained in terms of dissolution and reprecipitation processes caused through the high temperatures and higher Y activity during the Permian event. Retrogression of the samples during the Permian coupled with liberation of Y could be a plausible explanation for the onset of abundant Permian monazite. Variscan low-Y monazite grains probably recrystallized in order to adjust their Y contents to the higher temperatures during the Permian event. This is a recrystallization mechanism that has been reported in other studies as well (e.g. Janots et al. 2008). A high strain rate (Berger et al. 2006) and a high degree of retrogression/alteration (Lanzirotti and Hanson 1996; Poitrasson et al. 1996, 2000; Krenn and Finger 2007; Budzyń et al. 2011) are also favorable for monazite recrystallization. Indentions and protrusions observed at the contacts between distinct monazite age zones (Fig. 5) suggest that pre-existing Variscan monazites were marginally replaced through a dissolution and reprecipitation mechanism (e.g. Putnis 2002, 2009). On the other hand, the euhedral shape of Variscan monazite cores also indicate that some of them were probably overgrown by neighboring monazite substance. This could be an explanation for the lower Th-contents in some Permian monazite rims compared to the corresponding Variscan monazite cores.

6.2 Formation of Eo-Alpine monazite

Although monazite can survive high-grade metamorphism and even migmatite-grade conditions (e.g. Zhu and O`Nions 1999a, b; Martins et al. 2009), it is surprising that so many pre-Alpine monazite grains survived the Eo-Alpine amphibolite facies event. It is assumed that pre-Alpine monazites remained unaffected because their Y contents (at least in Permian monazite) were high and the degree of retrogression of the samples during the Eo-Alpine event was low. Krenn and Finger (2007) showed that the pre-existing Variscan monazite generation in polymetamorphic rocks from Crete remained largely unchanged in quartz-rich layers but barely survived in mica-rich layers of the same rock. The results showed that a lack of monazite age generation should not be misinterpreted as indicating that the degree of a metamorphic event was low or that a metamorphic event did not occur at all. Krenn and Finger (2007) also showed that the younger monazite generation likely formed in allanite-bearing domains, while allanite-absent domains hardly yielded young monazite. Indeed, in sample HPr10 from the Schobergruppe, where Eo-Alpine monazite is abundant (Schulz et al. 2005), allanite is common as well. It served as a direct precursor to Eo-Alpine monazite, as reported in other studies (e.g. Wing et al. 2003; Janots et al. 2006, 2008). No allanite formed in samples 520, 527 and Alk8, probably because of a low bulk Ca content, which is more favorable to the growth of monazite relative to that of allanite according to Janots et al. (2007) and Spear (2010). Allanite is often the prevailing REE-bearing phase in metapelites up to lower amphibolite facies conditions, where it reacts to monazite (e.g., Wing et al. 2003). The allanite to monazite transition has a slight positive slope in PT space (Janots et al. 2007; Spear 2010) and moves towards lower temperatures with decreasing CaO whole-rock content or with increasing Al2O3 content. This results in a decrease or complete disappearance of the allanite stability field.

6.3 Growth of garnet

We identified three growth stages of garnet (M1–M3). M1 is observed sporadically in the cores of a few garnet crystals and is characterized by high Mn and low Mg values. M2 shows high Ca and intermediate Mg values and can be related to a high-pressure amphibolite facies stage at 550–600 °C/13–16 kbar. Garnet growth zone M3 prevails in many garnet crystals and is characterized by decreasing Ca and increasing Mg at low Mn contents resulting from a thermal maximum at 650–700 °C/6–9 kbar. The following assemblages (including quartz and muscovite) were observed together with the garnet growth zones M1–M3: M1 − garnet + chlorite; M2 − garnet + chlorite + biotite + plagioclase + monazite + xenotime ± staurolite ± kyanite; M3 − garnet + chlorite + biotite + plagioclase + staurolite + kyanite + monazite + xenotime. M2 and M3 stages were also documented in other micaschists from the Schobergruppe (Schulz et al. 2005, 2008) and match the thermobarometric results obtained from eclogitic amphibolites (Schulz 1993a; Schulz et al. 2008). Eclogitic amphibolites (Schulz 1993a; Schulz et al. 2008) contain a high-pressure assemblage of garnet plus clinopyroxene, which can be related to the M2 stage observed in metapelites and a subsequent high-temperature assemblage with Ca-amphibole plus zoisite, which can be ascribed to the M3 stage in metapelites. Eclogitic amphibolites are situated structurally below and above the micaschists with parallel planar-linear structures indicative of a common deformation history. PT estimates for M2 and M3 are also in agreement with those obtained for other basement rocks from the Schobergruppe (Hoinkes et al. 1999; Linner et al. 2000).

However, while Hoinkes et al. (1999) ascribed the M2 and M3 stages to a Cretaceous event, Schulz (1993a) and Schulz et al. (2005, 2008) proposed a Variscan age. Results from this study, in particular the observation of Eo-Alpine monazite inclusions within M2 and M3 garnet domains, suggest that the M2 and M3 stages are Eo-Alpine rather than pre-Alpine. It is unlikely that the Eo-Alpine monazite grains shown in Fig. 5e crystallized after they were enclosed in garnet. There are no textural evidences such as pseudomorphs, recrystallization patches or overgrowths, which would argue for a later formation of monazite. Also, there are no cracks visible along which REEs could have entered garnet. In addition, REEs are relatively immobile during regional metamorphic processes and migration of REEs is expected rather at low metamorphic grade associated with diagenetic-, hydrothermal- and metasomatic processes (see Čopjaková et al. 2011 and references therein). Moreover, the high xenotime content observed in some Eo-Alpine monazite is also consistent with crystallization during the M3 thermal peak.

6.4 Pre-Alpine development of garnet

The age and growth conditions of M1-related garnet are not well known because they lack inclusions of biotite, plagioclase and monazite. Although we do not know if M1-related garnet domains represent pre-Alpine remnants, this is likely for several reasons. According to the monazite thermometry, samples 520, 527, Alk 8a were metamorphosed at temperatures of 500 and 650 °C during the Variscan and Permian events, respectively. This is high enough for the growth of garnet. In addition, growth of Variscan and Permian garnet has been documented from other metapelitic samples from the basement south of the Tauern window (Schuster et al. 2001; Steidl et al. 2010a, b). Steidl et al. (2010a, b) proposed PT conditions of 3.8–5.8 kbar and up to 650 °C for the Permian event and assumed amphibolite facies conditions for the Variscan event. Similarly our results would argue for a high-temperature Permian event and a Variscan event at lower amphibolite facies conditions.

According to the structural position of some of the studied metapelites (Fig. 1b, d), Permian recrystallization in these rocks may be related to the intrusion of Permian pegmatites. However, Permian monazite in other metapelites, which occur at a greater distance to pegmatites (Fig. 1), would rather argue for a discrete Permian HT/LP event. This is supposed by Schuster and Stüwe (2008) for the metamorphic complexes south and southeast of the Schobergruppe (Jenig and Strieden Complexes). Moreover, the large number of pegmatites across the entire basement of the Schobergruppe (Bücksteeg 1999) is also consistent with a regional Permian HT/LP event (Schuster and Stüwe 2008), which would have been accompanied by partial melting of the lower crust.

Permian metamorphism is widespread in the Eastern Alps (Habler and Thöni 2001; Schuster et al. 2001) and is particularly pronounced in the crystalline complexes of the so-called Wölz-Koralpe nappe system (Schuster et al. 2004). Remnants of Variscan mineral assemblages are rarely observed in rocks from the Wölz-Koralpe nappe system and have been reported for instance from the Strieden and Jenig Complexes (Schuster et al. 2001, 2004), the Michelbach Complex of the Defereggen Mountains (Steidl et al. 2010a, b) or from the Rappold Complex (Gaidies et al. 2008). From these units, the Rappold Complex shows a similarly high-grade metamorphic Eo-Alpine overprint (Gaidies et al. 2008) like the studied samples from the Schobergruppe. Other crystalline complexes of the Wölz-Koralpe nappe system (e.g., Wölz Unit, Saualpe-Koralpe Complexes, Pohorjeh Mountains, Schneeberg Complex) usually contain assemblages indicative for a Permian and/or a Eo-Alpine imprint (e.g., Abart and Martinelli 1991; Schuster and Thöni 1996; Thöni and Miller 1996; Bernhard and Hoinkes 1999; Faryad and Chakraborty 2005; Gaidies et al. 2006; 2008; Schuster et al. 2001).

7 Conclusions

Petrographic and textural evidence from different generations of monazite crystals indicate a Cretaceous Eo-Alpine age for the high-pressure amphibolite facies metamorphic events M2 and M3 recorded in the Schobergruppe. In contrast, relict monazite crystals record pre-Alpine metamorphic events of amphibolite grade, whatever the intensity of the subsequent Alpine overprint. Monazite records Carboniferous, Permian and Cretaceous age groups in agreement with preexisting data in the Schobergruppe, like sample HPr10, which documents a distinct Cretaceous monazite crystallization event in the lower parts of the Northern-Defereggen-Petzeck Group. The finding of Permian monazite ages in samples with both multi and single monazite age populations implies a metamorphism event clearly distinct from those of Variscan and Alpine ages. This study also illustrates the resistance of monazite at very high-grade metamorphic conditions and its potential use as a relict in unraveling the complex history of polymetamorphic rocks.

References

  • Abart, R., & Martinelli, W. (1991). Variszische und alpidische Entwicklungsgeschichte des Wölzer Kristallins (Steiermark, Österreich). Mitteilungen der Gesellschaft der Geologie- und Bergbaustudenten in Österreich, 37, 1–14.

    Google Scholar 

  • Berger, A., Herwegh, M., & Gnos, E. (2006). Deformation of monazite in an amphibolite-facies shear zone. Geochimica et Cosmochimica Acta, 70, A47–A47.

    Article  Google Scholar 

  • Bernhard, F., & Hoinkes, G. (1999). Polyphase micaschists of the central Wölzer Tauern, Styria, Austria. Berichte der Deutschen Mineralogischen Gesellschaft, Beiheft zum European Journal of Mineralogy, 11, 32.

    Google Scholar 

  • Bhattacharya, A., Mohanty, L., Maji, A., Sen, S. K., & Raith, M. (1992). Non-ideal mixing in the phlogopite-annite binary: constraints from experimental data on Fe-Mg partitioning and a reformulation of the garnet-biotite geothermometer. Contributions to Mineralogy and Petrology, 111, 87–93.

    Article  Google Scholar 

  • Biševac, V., Krenn, E., Balen, D., Finger, F., & Balogh, K. (2011). Petrographic, geochemical and geo-chronological investigations on granitic pebbles from Permotriassic metasediments of the Tisia terrane (eastern Papuk, Croatia). Mineralogy and Petrology, 102, 163–180.

    Article  Google Scholar 

  • Borsi, S., Del Moro, A., Sassi, F. P., Zanferrari, A., Zirpoli, G. (1978). New geopetrologic and radiome-tric data on the Alpine history of the Austridic continental margin south of the Tauern Window. Memorie dell’ Istituto Geologico dell` Università di Padova, 32, 1–17.

  • Bücksteeg, A. (1999). Zur Geologie des Kristallins der Schobergruppe (Osttirol/Österreich). Aachener Geowissenschaftliche Beiträge, 33, 1–206.

    Google Scholar 

  • Budzyń, B., Harlov, D. E., Williams, M. L., & Jercinovic, M. J. (2011). Experimental determination of stability relations between monazite, fluorapatite, allanite, and REE-epidote as a function of pressure, temperature, and fluid composition. American Mineralogist, 96, 1547–1567.

    Article  Google Scholar 

  • Cocherie, A., Legendre, O., Peucat, J., & Koumelan, A. (1998). Geochronology of polygenetic monazites constrained by in situ microprobe Th-U-total lead determination: implications for lead behavior in monazite. Geochimica et Cosmochimica Acta, 62, 2475–2497.

    Article  Google Scholar 

  • Čopjaková, R., Novák, M., & Franců, E. (2011). Formation of authigenic monazite-(Ce) to monazite-(Nd) from Upper Carboniferous graywackes of the Drahany Upland: roles of the chemical composition of host rock and burial temperature. Lithos, 127, 373–385.

    Article  Google Scholar 

  • Del Rio, P., Barbero, P., Mata, P., & Fanning, C. M. (2009). Timing of diagenesis and very low-grade metamorphism in the eastern sector of the Sierra de Cameros (Iberian Range, Spain): a U-Pb SHRIMP study on monazite. Terra Nova, 21, 438–445.

    Article  Google Scholar 

  • DeWolf, C. P., Belshaw, N. S., & O′Nions, R. K. (1993). A metamorphic history from micron-scale 207Pb/206Pb chronometry of Archean monazite. Earth and Planetary Science Letters, 120, 207–220.

    Article  Google Scholar 

  • Evans, J. A., Jalasiewicz, J. A., Fletcher, I. R., Rasmussen, B., & Pearce, N. J. G. (2002). Dating diagenetic monazite in mudrocks: constraining the oil window? Special Publication of the Geological Society, 159, 619–622.

    Article  Google Scholar 

  • Exner, Ch. (1962). Die Perm-Trias-Mulde des Gödnachgrabens an der Störungslinie von Zwischenbergen (Kreuzeckgruppe, östlich Lienz). Verhandlungen der Geologischen Bundesanstalt, 1962, 24–27.

    Google Scholar 

  • Faryad, S. W., & Chakraborty, S. (2005). Duration of Eo-Alpine metamorphic events obtained from multicomponent diffusion modeling of garnet: a case study from the Eastern Alps. Contributions to Mineralogy and Petrology, 150, 306–318.

    Article  Google Scholar 

  • Frey, M., Desmons, J., Neubauer, F. (1999). Metamorphic maps of the Alps. Published by the editors and as enclosure to the Schweizerische Mineralalogische und Petrographische Mitteilungen, 79/1.

  • Frey, M., Desmons, J., & Neubauer, F. (1999b). The new metamorphic map of the Alps: introduction. Schweizerische Mineralalogische und Petrographische Mitteilungen, 79, 1–4.

    Google Scholar 

  • Gaidies, F., Abart, R., de Capitani, C., Schuster, R., Connolly, J. A. D., & Reusser, E. (2006). Characterization of polymetamorphism in the Austroalpine basement east of the Tauern Window using garnet isopleth thermobarometry. Journal of Metamorphic Geology, 24, 451–475.

    Article  Google Scholar 

  • Gaidies, F., Krenn, E., de Capitani, C., & Abart, R. (2008). Coupling forward modelling of garnet growth with monazite geochronology: an application to the Rappold Complex (Austroalpine crystalline basement). Journal of Metamorphic Petrology, 26, 775–793.

    Article  Google Scholar 

  • Ganguly, J., Cheng, W., & Tirone, M. (1996). Thermodynamics of aluminosilicate garnet solid solution: new experimental data, an optimized model, and thermometric applications. Contributions to Mineralogy and Petrology, 123, 137–151.

    Article  Google Scholar 

  • Gratz, R., & Heinrich, W. (1997). Monazite-xenotime thermobarometry: experimental calibration of the miscibility gap in the binary system CePO4 -YPO4. American Mineralogist, 82, 772–780.

    Google Scholar 

  • Habler, G., & Thöni, M. (2001). Preservation of Permo-Triassic low-pressure assemblages in the Cretaceous high-pressure metamorphic Saualpe crystalline basement (Eastern Alps, Austria). Journal of Metamorphic Geology, 19, 679–697.

    Article  Google Scholar 

  • Heinrich, W., Andrehs, G., & Franz, G. (1997). Monazite-xenotime miscibility gap thermometry. I. An empirical calibration. Journal of Metamorphic Geology, 15, 3–16.

    Article  Google Scholar 

  • Hoinkes, G., Koller, F., Rantitsch, G., Dachs, E., Höck, V., Neubauer, F., et al. (1999). Alpine metamorphism of the Eastern Alps. Schweizerische Mineralalogische und Petrographische Mitteilungen, 79, 155–181.

    Google Scholar 

  • Hoke, L. (1990). The Altkristallin of the Kreuzeck Mountains, SE Tauern Window, Eastern Alps - basement crust in a convergent plate boundary zone. Jahrbuch der Geologischen Bundesanstalt, 133, 5–87.

    Google Scholar 

  • Holland, T. J. B., & 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.

    Article  Google Scholar 

  • Holland, T. J. B., & Powell, R. (1998). An internally consistent thermodynamic data set for phases of petrological interest. Journal of Metamorphic Geology, 16, 309–343.

    Article  Google Scholar 

  • Janots, E., Brunet, F., Goffé, B., Poinssot, C., Burchard, M., & 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.

    Article  Google Scholar 

  • Janots, E., Engi, M., Berger, A., Allaz, J., Schwarz, J.-O., & 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.

    Article  Google Scholar 

  • Janots, E., Negro, F., Brunet, F., Goffé, B., Engi, M., & Bouybaouene, M. L. (2006). Evolution of the REE mineralogy in HP-LT metapelites of the Sebtide complex, Rif, Marocco: monazite stability and geochronology. Lithos, 87, 214–234.

    Article  Google Scholar 

  • Kingsbury, J. A., Miller, C. F., Wooden, J. L., & Harrison, T. M. (1993). Monazite paragenesis and U-Pb systematics in rocks of the eastern Mojave Desert, California, U.S.A., implications for thermochronometry. Chemical Geology, 110, 147–167.

    Article  Google Scholar 

  • Krenn, E., & Finger, F. (2004). Metamorphic formation of Sr-apatite and Sr-bearing monazite in a high pressure rock from the Bohemian Massif. American Mineralogist, 89, 1323–1329.

    Google Scholar 

  • Krenn, E., & Finger, F. (2007). Formation of monazite and rhabdophane at the expense of allanite during Alpine low temperature retrogression of metapelitic basement rocks from Crete, Greece: microprobe data and geochronological implications. Lithos, 95, 130–147.

    Article  Google Scholar 

  • Krenn, E., Janák, M., Fritz, F., Broska, I., & Konečný, P. (2009). Two types of metamorphic monazite with contrasting La/Nd, Th and Y signature in a (ultra) high pressure metapelite from the Pohorje Mountains, Slovenia: indications for a pressure-depended REE exchange between apatite and monazite? American Mineralogist, 94, 801–815.

    Article  Google Scholar 

  • Krenn, E., Ustaszewski, K., & Finger, F. (2008). Detrital and newly formed metamorphic monazite in amphibolite-facies metapelites from the Motajica Massif, Bosnia. Chemical Geology, 254, 164–174.

    Article  Google Scholar 

  • Kryza, R., Zalasiewicz, J. A., Charnley, N., Milodowski, A. E., Kostylew, J., & Tyszka, R. (2004). In situ growth of monazite in anchizonal to epizonal mudrocks: first record from the Variscan accretionary prism of the Kaczawa Mountains, West Sudetes, SW Poland. Geologia Sudetica, 36, 39–51.

    Google Scholar 

  • Lanzirotti, A., & Hanson, G. N. (1996). Geochronology and geochemistry of multiple generations of monazite from Wepawaug Schist, Connecticut, USA: implications for monazite stability in metamorphic rocks. Contributions to Mineralogy and Petrology, 125, 332–340.

    Article  Google Scholar 

  • Linner, M., Thöni, M., & Richter, W. (2000). Exhumation history of Eo-Alpine high-pressure rocks in the Austroalpine Schober basement, Eastern Alps. Terra Nostra, 1, 69.

    Google Scholar 

  • Ludwig, K. (2001). Users Manual for Isoplot/Ex (rev. 2.49): A Geochronological Toolkit for Microsoft Excel. (p. 55) Berkeley Geochronology Center, Special Publication, 1a.

  • Martins, L., Vlach, S. R. F., & Janasi, V. D. (2009). Reaction microtextures of monazite: correlation between chemical and age domains in the Nazare Paulista migmatite, SE Brazil. Chemical Geology, 261, 271–285.

    Article  Google Scholar 

  • Montel, J.-M., Foret, S., Veschambre, M., Nicollet, C., & Provost, A. (1996). Electron microprobe dating of monazite. Chemical Geology, 131, 37–51.

    Article  Google Scholar 

  • Neubauer, F., Hoinkes, G., Sassi, F. P., Handler, R., Höck, V., Koller, F., et al. (1999). Pre-Alpine metamorphism in the Eastern Alps. Schweizerische Mineralalogische und Petrographische Mitteilungen, 79, 41–62.

    Google Scholar 

  • Oxburgh, E.R., Lambert, R.S., Baadsgard, H., Simons, J.G. (1966). Potassium-Argon age studies across the southeast margin of the Tauern window in the Eastern Alps. Verhandlungen der Geologischen Bundesanstalt, 17–33.

  • Petrík, I., & Konečný, P. (2009). Metasomatic replacement of inherited metamorphic monazite in a biotite-garnet granite from the Nízke Tatry Mountains, Western Carpathians, Slovakia: chemical dating and evidence for disequilibrium melting. American Mineralogist, 94, 957–974.

    Article  Google Scholar 

  • Poitrasson, F., Chenery, S., & Bland, D. J. (1996). Contrasted monazite hydrothermal alteration mechanisms and their geochemical implications. Earth and Planetary Science Letters, 145, 79–96.

    Article  Google Scholar 

  • Poitrasson, F., Chenery, S., & 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.

    Article  Google Scholar 

  • Powell, R., & Holland, T. J. B. (1993). On the formulation of simple mixing models for complex phases. American Mineralogist, 78, 1174–1180.

    Google Scholar 

  • Putnis, A. (2002). Mineral replacement reactions: from macroscopic observations to microscopic mechanisms. Mineralogical Magazine, 66, 689–708.

    Article  Google Scholar 

  • Putnis, A. (2009). Mineral replacement reactions. In: K.D. Putirka, F.J. Tepley (Eds.), Minerals, inclusions and volcanic processes. Reviews in Mineralogy and Geochemistry, 70 (pp. 87–124).

  • Pyle, J. M. (2006). Temperature-time paths from phosphate accessory phase paragenesis in the Honey Brook Upland and associated cover sequence, SE Pennsylvania, USA. Lithos, 88, 201–232.

    Article  Google Scholar 

  • Pyle, J. M., & Spear, F. S. (1999). Yttrium zoning in garnet: coupling of major and accessory phases during metamorphic reactions. Geological Materials Research, 1, 1–49.

    Google Scholar 

  • Pyle, J. M., & Spear, F. S. (2003). Four generations of accessory-phase growth in low-pressure migmatites from SW New Hampshire. American Mineralogist, 88, 338–351.

    Google Scholar 

  • Pyle, J. M., Spear, F. S., Rudnick, R., & Mcdonough, W. (2001). Monazite-xenotime-garnet equilibrium in metapelites and a new monazite-garnet thermometer. Journal of Petrology, 42, 2083–2107.

    Article  Google Scholar 

  • Rasmussen, B., Fletcher, I. R., & McNaughton, N. J. (2001). Dating low grade metamorphic events by SHRIMP U-Pb analysis of monazite in shales. Geol, 29, 963–966.

    Article  Google Scholar 

  • Rasmussen, B., & Muhling, J. R. (2007). Monazite begets: evidence for dissolution of detrital monazite and reprecipitation of syntectonic monazite during low-grade regional metamorphism. Contributions to Mineralogy and Petrology, 154, 675–689.

    Article  Google Scholar 

  • Rasmussen, B., & Muhling, J. R. (2009). Reactions destroying detrital monazite in greenschist-facies sandstones from the Witwatersrand basin, South Africa. Chemical Geology, 264, 311–327.

    Article  Google Scholar 

  • Read, D., Cooper, D. C., & McArthur, J. M. (1987). The composition and distribution of nodular monazite in the Lower Palaeozoic rocks of Great Britain. Mineralogical Magazine, 51, 271–280.

    Article  Google Scholar 

  • Rubatto, D., Hermann, J., & 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.

    Article  Google Scholar 

  • Schmid, S. M., Fügenschuh, B., Kissling, E., & Schuster, R. (2004). Tectonic map and overall architecture of the Alpine orogen. Eclogae Geologicae Helvetiae, 97, 93–117.

    Article  Google Scholar 

  • Schulz, B. (1990). Prograde-retrograde P-T-t-de-formation path of Austroalpine micaschists du-ring Variscan continental collision (Eastern Alps). Journal of Metamorphic Geology, 8, 629–643.

    Article  Google Scholar 

  • Schulz, B. (1993a). Mineral chemistry, geother-mo-barometry and pre-Alpine high-pressure meta-morphism of eclogitic amphibolites and mica schists from the Schobergruppe, Austroalpine basement, Eastern Alps. Mineralogical Magazi-ne, 57, 189–202.

    Article  Google Scholar 

  • Schulz, B. (1993b). P-T-deformation paths of Variscan metamorphism in the Austroalpine basement: controls on geothermobarometry from microstructures in progressively deformed metapelites. Schweizerische Mineralogische und Petrographische Mitteilungen, 73, 257–274.

    Google Scholar 

  • Schulz, B., Finger, F., Krenn, E. (2005). Auflösung variskischer, permischer und alpidischer Ereignisse im polymetamorphen ostalpinen Kristallin südlich der Tauern mit EMS-Datierung von Monazit. Arbeitstagung der Geologischen Bundesanstalt Österreich, 141153.

  • Schulz, B., Steenken, A., Siegesmund, S. (2008). Geodynamics of an Alpine terrane—the Austroalpine basement to the south of the Tauern Window as a part of the Adriatic Plate. In: S. Siegesmund, B. Fügenschuh., & N. Froitzheim (Eds.), Tectonic aspects of the Alpine-Dinaride-Carpathian system, 298, (pp. 5–43). Geological Society of london special publications.

  • Schulz, B., & von Raumer, J. F. (2011). Detection of a pre-Variscan metamorphic event by EMP monazite dating and thermobarometry of garnet metapelites in the Alpine External Aiguilles Rouges Massif. Swiss Journal of Geosciences, 104, 67–79.

    Article  Google Scholar 

  • Schuster, R., Koller, F., Höck, V., Hoinkes, G., & Bousquet, R. (2004). Explanatory notes to the map: metamorphic structure of the Alps—Metamorphic evolution of the Eastern Alps. Mitteilungen der Österreichischen Mineralogischen Gesellschaft, 149, 175–199.

    Google Scholar 

  • Schuster, R., Scharbert, S., Abart, R., & Frank, W. (2001). Permo-Triassic extension and related HT/LP metamorphism in the Austroalpine—Southalpine realm. Mitteilungen der Gesellschaft der Geologie und Bergbaustudenten Österreichs, 45, 111–141.

    Google Scholar 

  • Schuster, R., & Stüwe, K. (2008). Permian metamorphic events in the Alps. Geol, 36, 603–606.

    Article  Google Scholar 

  • Schuster, R., & Thöni, M. (1996). Permian garnet: indications for a regional Permian metamorphism in the southern part of the Austroalpine basement units. Mitteilungen der Österreichischen Geologischen Gesellschaft, 141, 219–221.

    Google Scholar 

  • Seydoux-Guillaume, A.-M., Wirth, R., Heinrich, W., & Montel, J. M. (2002). Experimental determination of Thorium partitioning between monazite and xenotime using analytical electron microscopy and X-ray diffraction Rietveld analysis. European Journal of Mineralogy, 14, 869–878.

    Article  Google Scholar 

  • Spear, F. S. (2010). Monazite–allanite phase relations in metapelites. Chemical Geology, 279, 55–62.

    Article  Google Scholar 

  • Spear, F. S., Cheney, J. T., Pyle, J. M., Harrison, T. M., & Layne, G. (2008). Monazite geochronology in central New England: evidence for a fundamental terrain boundary. Journal of Metamorphic Geology, 26, 317–329.

    Article  Google Scholar 

  • Steidl, M., Tropper, P., Linner, M., & Schuster, R. (2009). Petrology of metapelites from the Michelbach Complex (Defreggen Complex, Eastern Tyrol). Mitteilungen der Österreichischen Geologischen Gesellschaft, 155, 149.

    Google Scholar 

  • Steidl, M., Tropper, P., Linner, M., & Schuster, R. (2010a). Unravelling the polymetamorphic (Variscan vs. Permian) history of the Michlbach Complex (Deffregger Alps, East Tyrol) by using REE phosphates (monazite, xenotime). Journal of Alpine Geology, 52, 234–235.

    Google Scholar 

  • Steidl, M., Tropper, P., Linner, M., & Schuster, R. (2010b). Petrology of the polymetamorphic metapelites from the Michlbach Complex (Defreggen Complex, East Tyrol). Journal of Alpine Geology, 52, 235.

    Google Scholar 

  • Suzuki, K., Adachi, M., & Tanaka, T. (1991). Middle Precambrian provenance of Jurassic sandstone in the Mino Terrane, central Japan: th-U-total Pb evidence from an electron microprobe monazite study. Sedimentary Geology, 75, 141–147.

    Article  Google Scholar 

  • Thöni, M., & Miller, Ch. (1996). Garnet Sm-Nd data from the Saualpe and the Koralpe (Eastern Alps, Austria): chronological and P-T constraints on the thermal and tectonic history. Journal of Metamorphic Geology, 14, 453–466.

    Article  Google Scholar 

  • Todd, C. S. (1998). Limits on the precision of geobarometry at low grossular and anorthite content. American Mineralogist, 83, 1161–1167.

    Google Scholar 

  • Troll, G., Baumgartner, S., & Daiminger, W. (1980). Zur Geologie der südwestlichen Schobergruppe (Osttirol, Österreich). Mitteilungen der Gesellschaft der Geologie- und Bergbaustudenten Österreichs, 26, 277–295.

    Google Scholar 

  • Troll, G., Forst, R., Söllner, F., Brack, W., Kohler, H., & Müller-Sohnius, D. (1976). Über Bau, Alter und Metamorphose des Altkristallins der Schobergruppe, Osttirol. Geologische Rundschau, 65, 483–511.

    Article  Google Scholar 

  • Troll, G., & Hölzl, E. (1974). Zum Gesteinsaufbau des Altkristallins der zentralen Schobergruppe, Osttirol. Jahrbuch der Geologischen Bundesanstalt, 117, 1–16.

    Google Scholar 

  • Wan, Y. S., Song, T. R., Liu, D. Y., Yang, T. N., Yin, X. Y., Chen, Z. Y., et al. (2007). Mesozoic monazite in Neoproterozoic metasediments: evidence for low-grade metamorphism of Sinian Sediments during Triassic Continental Collision, Liaodong Peninsula, NE China. Geochemical Journal, 41, 47–55.

    Article  Google Scholar 

  • Wilby, P. R., Page, A. A., Zalasiewicz, J. A., Milodowski, A. E., Williams, M., & Evans, J. A. (2007). Syntectonic monazite in low-grade mudrocks: a potential geochronometer for cleavage formation? Journal of the Geological Society, 164, 53–56.

    Article  Google Scholar 

  • Wing, B., Ferry, J., & Harrison, T. (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.

    Article  Google Scholar 

  • Zhu, X. K., & O`Nions, R. K. (1999a). Zonation of monazite in metamorphic rocks and its implications for high temperature thermochronology: a case study from the Lewisian Terrain. Earth and Planetary Science Letters, 171, 209–220.

    Article  Google Scholar 

  • Zhu, X. K., & O`Nions, R. K. (1999b). Monazite chemical composition: some implications for monazite geochronology. Contributions to Mineralogy and Petrology, 137, 351–363.

    Article  Google Scholar 

Download references

Acknowledgments

This work was supported by the Austrian Science Foundation through projects P13070 and P22408 (to F.F.) and by the Deutsche Forschungsgemeinschaft (Project SCHU-676-9). M. Göbbels and N. Langhof are thanked for their assistance during electron-microprobe sessions and François Bussy for handling the manuscript. The work strongly benefited from thorough reviews of Alfons Berger and Peter Tropper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Erwin Krenn.

Additional information

Editorial handling: François Bussy

Rights and permissions

Reprints and permissions

About this article

Cite this article

Krenn, E., Schulz, B. & Finger, F. 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 J Geosci 105, 343–360 (2012). https://doi.org/10.1007/s00015-012-0104-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00015-012-0104-6

Keywords