You are on page 1of 18

Available online at www.sciencedirect.

com

Lithos 102 (2008) 598 – 615


www.elsevier.com/locate/lithos

Tectonically controlled fluid flow and water-assisted melting in the


middle crust: An example from the Central Alps
Alfons Berger ⁎, Thomas Burri 1 , Peter Alt-Epping, Martin Engi
Institute of Geological Sciences, University of Bern, Baltzerstrasse 1 & 3, 3012 Bern, Switzerland
Received 19 August 2006; accepted 25 July 2007
Available online 5 September 2007

Abstract

Melting triggered by influx of a free aqueous fluid in the continental crust has commonly been inferred, but the source of water
in such contexts remains a matter of debate. We focus on the Tertiary migmatites in the Southern Steep Belt of the Central Alps
(Switzerland) to discuss the petrology, structures and geodynamic setting of water-assisted melting. These migmatites comprise
various structural types (e.g. metatexites, diatexites, melt in shear zones), which reflect variable leucosome fractions. The melting
event itself as well as the variable melt fractions are related to the amount of aqueous fluids. At a given P and T, melt-fractions in
rocks of minimum melt composition correlate with the amount of infiltrated aqueous fluids. In more granodioritic systems the
water distributes between melt and newly crystallizing hydrous phases such as amphibole, such that the melt fraction correlates
with the contents of H2O, Al, and Ca in the system. Phase-equilibrium modelling indicates that the stabilization of amphibole leads
to slightly lower melt fractions than in a granitic system at the same P, T and bulk water content. Phase-equilibrium models further
indicate that in the Alpine migmatite belt: (1) several wt.% water (fluid:rock ratio of ∼ 1:30) are necessary to produce the inferred
melt fraction; (2) the activity of H2O in the fluid is high; and (3) spatially associated metapelites are unlikely as a source for the
required aqueous fluids.
We present a tectonic scenario for the southern margin of the Central Alps, to which these migmatites are confined, and we
propose that water was produced from dehydration reactions in metapelites in the Southern Alps. We model fluid production rates
at the time of melting and demonstrate that the resulting fluid flow pattern is mainly controlled by the differences in permeability
between the fluid source region and melting region. The proposed model requires strong gradients in temperature and permeability
for the two tectonic blocks. This is consistent with the scenario involving indenter tectonics at the boundary between the Central
and the Southern Alps in Oligocene times.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Water-assisted melting; Amphibole-bearing migmatites; Fluid flow; Modelling; Alps

1. Introduction

The formation of major proportions of silicate melt in


the continental (middle) crust strongly depends on the
⁎ Corresponding author. Tel.: +41 31 6318493. availability of volatile components. Wherever major
E-mail address: berger@geo.unibe.ch (A. Berger). amounts of melt occur, for example in migmatite
1
Now at: Kellerhals & Häfeli AG, Kapellenstr. 22, 3011 Bern. terrains, substantial amounts of such volatiles, notably
0024-4937/$ - see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.lithos.2007.07.027
A. Berger et al. / Lithos 102 (2008) 598–615 599

Fig. 1. Schematic–tectonic sketch of the Central Alps. The location of Alpine in situ melting is shown. Map is simplified after Burri et al. (2005).

H2O, are implied, and their source is an obvious question. where the source of water for partial melting has been
Potential source rocks are those containing one or more related to the infiltration of fluids from a crystallizing
hydrate phases, which at suprasolidus temperatures may pluton (e.g., Yardley and Barber, 1991; Berger and
break down to produce a partial melt.2 In the last decades Rosenberg, 2003; Johnson et al., 2003, 2004). However,
research on migmatites has focused on migmatites that water-assisted melting also occurs in large-scale migma-
involve such hydrate-breakdown melting (e.g., Waters tite terrains, which are characterized by masses of gran-
and Whales, 1984; Montel et al., 1992; Braun et al., 1996; itoid gneiss (e.g., Brown, 1979; Burri et al., 2005). Water-
Brown and Dallmeyer, 1996; Kalt et al., 1999; White assisted melting is an important process in metagranitoids,
et al., 2003; Harris et al., 2005). However, there is which have bulk-rock compositions close to the minimum
evidence for melt generation without dehydration reac- melt composition (Sawyer, 1998).
tions. The low porosity (≪0.01 vol.%) of metamorphic Fundamental questions related to these large-scale
rocks in the middle crust allows only for very small melt migmatite terrains are:
fractions (b1%) from fluids stored in situ. The origin of
migmatites, which have high leucosome fractions but did (1) In which geodynamic scenarios is fluid assisted
not undergo hydrate-breakdown reactions, has to be melting possible? Such a scenario is difficult to
related to water-assisted partial melting. This process has imagine, because either fluids will direct generate
been considered as a possible cause for migmatite melts, or fluid is produced below the solidus and
formation (Mogk, 1992; Butler et al., 1997; Viruete, will never hit the melting field.
1999; Prince et al., 2001; Garlick and Gromet, 2004; (2) What is the petrology during fluid assisted melting?
White et al., 2005). One possible location of water- Recent experimental studies and models have
assisted melting is in contact metamorphic aureoles, improved our understanding of the process of
hydrate-breakdown melting, but the phase relations
2
The term dehydration melting has been widely used for this
during water-assisted melting as addressed by Castro
process, but it is misleading, as no fluid is produced; hence we use the et al. (2000) and Gardien et al. (2000) are less well
term hydrate-breakdown melting, as introduced by Brown (2004). known.
600 A. Berger et al. / Lithos 102 (2008) 598–615

This paper reports on phase relations, structures and are widespread. In the area of Bellinzona a zone of
the fluid evolution in migmatites of the Central Alps muscovite-breakdown melting has been mapped,
(Switzerland, N-Italy). This migmatite belt is well suited whereas biotite dehydration melting is negligible
for this kind of investigation, because it has zones of (Burri et al., 2005). On average the SSB shows a high
mixed lithologies and much is known about its spatial leucosome fraction but only little evidence of hydrate-
dimensions and internal characteristics (Burri et al., breakdown melting. The timing of the partial melting is
2005). constrained by monazite ages in migmatites and ages
In order to have some new insights, we compare of aplites, which scatter between 30 and 25 Ma (e.g.
detailed observations in the Alpine migmatites with Gebauer, 1996; Köppel, 1993).
different modelling approaches. The first part of the Stromatic migmatites are common and may show
paper describes the petrography and petrology of transitions into orthogneiss-types, which had been
migmatites in the Central Alps, with special emphasis termed “injections gneisses” (Gutzwiller, 1912; Wenk,
on amphibole-bearing migmatites. The next section 1975). Within the migmatites that were derived from
summarizes the geodynamic and tectonic situation of granitoids, large variations in leucosome volumes are
the Alps during water-assisted melting. This tectonic observed. In other words, the distribution of partial
scenario is used to construct a simple fluid flow model, melts was spatially variable, even where partial melting
based on which some numerical results are presented. occurred in a homogeneous and compositional suitable
Finally, we discuss implications of our observations and protolith (Fig. 2). Virtually leucosome free gneisses
of these model results with respect to the evolution of (Fig. 2a, left) locally grade into metatexitic, stromatic
the Alpine migmatite belt.

2. Methods

Whole-rock major elements compositions were


determined using conventional XRF methods on glass
pellets. Major element equilibration was calculated
using TWQ (Berman, 1991). We use phase diagrams
and phase variation diagrams computed with the
program THERIAK and DOMINO (de Capitani and
Brown, 1987), in combination with the Holland and
Powell (1998) database and updates for melt bearing
systems (Holland and Powell, 2001; White et al., 2001).
Leucosome volumes were estimated in 2D using field
photographs and NIH-Image. Modelling of fluid flow
was carried out using the high-temperature flow model
Hydrotherm (Hayba and Ingebritsen, 1994).

3. Field relationships and petrology

3.1. Geological setting and field observations

In the Alps, classic basement thrust sheets can be


distinguished from heterogeneous tectonic mélange
units (e.g. Adula-, Cima Lunga-, Someo- and Orse-
lina-units) that contain eclogite relics (Fig. 1; Heinrich,
1982; Trommsdorff, 1990; Engi et al., 2001). The
Southern Steep Belt (SSB) of the Central Alps bears Fig. 2. Anatectic granodioritic gneiss. (a) The leucosome fraction in
the most convincing evidence of Alpine anatexis the right half of the image is estimated 0.19–0.22 using image analysis.
(Wenk, 1970; Burri et al., 2005). The zone of anatexis Note homogeneous distribution of the melt. Only a small melt fraction
is observed in the dominantly gneissic part to the left. (Pocket knife for
is spatially associated with aplites and small granites of scale) (b) Stromatic metatexite in the right part of the picture grades
continental origin. Pegmatite and aplite dykes, meter to into an almost diatexite structure (same locality as panel a). Width of
decameter-sized granitic stocks, and in situ migmatites photograph ca. 80 cm.
A. Berger et al. / Lithos 102 (2008) 598–615 601

Fig. 3. Field- and microphotographs of different migmatites. (a) Fold with leucosome parallel to fold axial plane (b) Melt-filled shear zones in
migmatitic biotite-gneiss (Ponte Brolla), (c) Microphotograph of muscovite-breakdown melting in the central part of the migmatite belt. Hydrate-
breakdown melting dominates in metapelitic types. (d) Amphibole-bearing migmatite of initially granodioritic composition (see also Fig. 5).

migmatites (Fig. 2a, right) and even into meter scale, case, the leucosome volume fractions determined from
almost diatexitic migmatites (Fig. 2b). Here the gneissic area measurements in field photographs of ten localities,
fabric of the protolith becomes diffuse or even vary between 0.05–0.39, with an average value of 0.2.
disintegrated. Beside the stromatic and strongly variable In addition, some outcrops have been found without
migmatites, leucosome also occur in some structural leucosome at all. Changes of the leucosome fraction
positions, as shear zones and fold axial planes (Fig. 3a often occur in metagranite and these changes must be
and b). correlated with spatially variable fluid contents. Like the
In several cases the heterogeneity of observed coexisting diatexites and metatexites, this is evidence
structures can be related to variable melt fractions. for the presence of fluids and the dependence of melt
Melt volume fractions generated through water-assisted fractions on fluid volumes.
melting are difficult to estimate. In case of reaction-
controlled incongruent melting the solid products can be 3.2. Petrography and petrology of the migmatites
used to estimate the melt fraction (e.g. Nyman et al.,
1995; Spear et al., 1999; Milord et al., 2001). In the case In addition to the variable leucosome volumes, the
of water-assisted melting, however, it is not possible to migmatites also differ in terms of their petrography.
estimate melt fraction from solid products. Whereas the According to the petrographic composition we distin-
volume of leucosome may give a first order estimate of guish the following groups:
the melt fraction, this approach poses two problems:
(1) it is not clear at what spatial and temporal scale the 1. Migmatites formed through white mica-breakdown
total volume of leucosome represented the actual melt melting (Fig. 3c). These rocks are often pelitic or
fraction at any one time; (2) the accuracy and statistical semipelitic in composition and sillimanite occurs as
significance of these estimates are strongly scale an incongruent melting product of the melting
dependent (e.g. Berger and Kalt, 1999). For the present reaction. However, in a few cases reactions of the
602 A. Berger et al. / Lithos 102 (2008) 598–615

as the cause for amphibole formation in anatectic


granites. In experiments by Naney (1983) and Gardien
et al. (2000), amphibole was stabilized only above the
water-saturated solidus upon addition of 3–5 wt.% of
H2O. This is clearly in excess of the amount of fluid that
could be stored in the pore space at amphibolite facies
conditions. This suggests that amphibole-bearing leuco-
somes should form in granitic to granodioritic rocks
only following fluid infiltration.
Studies carried out on amphibole-bearing leuco-
somes in the Central Alps (Fig. 3) indicate that a
reaction similar to (1) was indeed responsible for the
formation of partial melts and amphibole. Textural
evidence supporting this reaction is provided by epidote
grains, which are euhedral if included in biotite, but is
Fig. 4. Photographs of anatexis in amphibolites. (a) Hand sample from xenomorphic in shape within the quartz-feldspar matrix
the border of leucosome and mesosome. Note the poikilitic amphiboles (Fig. 5a). A finer-grained second generation of epidote,
of different grain size in the leucosome. (b) Microphotograph of the typically as aggregates, commonly rims the partially
same sample as in (a). Note the new large K-feldpsar crystals including
plagioclase and quartz. These are part of the crystallizing melt. Scale of
resorbed older epidote grains or occurs as precipitates in
the horizontal length of the image is ∼250 μm. fractures and along grain boundaries (Fig. 5a). We
interpret these textures as due to partial resorption of
epidote by incongruent melting (reaction (1)), followed
type muscovite + quartz + feldspar + biotite1 ⇒ garnet by a partial retrograde reaction during crystallization of
+biotite2 + melt do occur, which formed migmatites the melt to form the smaller grained epidote fraction.
with garnet, but without sillimanite. Other textural evidence involves quartz-biotite-sym-
2. Metabasic rocks with evidence of localized melting plectites, which are enclosed in amphibole (Fig. 5b).
and small volumes of leucosomes (Fig. 4). K-feldspar The symplectite is interpreted to have been formed due
occurs in these leucosomes and as thin films between to the breakdown of biotite according to reaction (1).
amphibole and plagioclase. These may have crystal- Inclusion of the symplectite during coeval growth of
lized from small melt fractions located along grain amphibole may have led to its partial preservation and
boundaries (Fig. 4b). Several amphibolites contain may have inhibited a complete reaction. Retrograde
different sized leucosomes, which have a granitic reaction of amphibole to form biotite, epidote and quartz
composition and magmatic texture. These observa- during crystallization of the melt has been observed
tions indicate, that these leucosomes are crystallized frequently (Fig. 5c). Biotite symplectites are not found
melt. Because of the granitic composition of these at such sites. Amphibole replacement occurs dominantly
leucosomes the amphibolites are not likely to be the along amphibole grain boundaries, or along its cleavage.
source for these melts. The replacement of amphibole along cleavage planes in
3. Migmatites in metagranitoids rocks. Fig. 5 may be promoted if retrograde reaction is volume
4. Amphibole-bearing migmatites related to water- increasing. Secondary minerals grew along fractures in
assisted melting (Fig. 3d). These are commonly amphibole and plagioclase, but not in quartz (Fig. 5c).
associated with partially molten meta-granitoid This indicates that retrograde replacement of amphibole
rocks. The origin of such migmatites has been occurred during crystallization of the melt, at super-
explored in an experimental study by Gardien et al. solidus conditions. The back-reaction is incomplete,
(2000) and will be discussed below. indicating that melt had escaped from the leucosome.
In metapelitic slivers within the SSB, muscovite-
For migmatites that developed under relatively low breakdown melting occurred only in small regions
temperature conditions, Mogk (1992) proposed the (Burri et al., 2005) and these migmatites show a large
melting reaction: variety of assemblages. Some samples contain relics of
white mica together with sillimanite (Fig. 3c); others
show only biotite and garnet or a paragenesis of garnet-
Bio þ Plagð1Þ þ Qtz 
þ Epi þ H2 O
sillimanite (Berger et al., 2005). Most of these
¼ Hbl þ Plagð2Þ þ MeltðKfs  richÞ ð1Þ migmatites have low leucosome volume and in most
A. Berger et al. / Lithos 102 (2008) 598–615 603

examples a stromatic structure. A particularly informa- Table 1


tive example is located in an inhomogeneous trail of Chemical bulk-rock compositions used in this contribution
meta-sedimentary and meta-mafic rocks inside the meta- Sample FON1 a Ar 9902-aver. b Granite c
SiO2 44.13 65.91 70.77
TiO2 2.11
Al2O3 27.36 16.84 14.83
FeO 4.39
Fe2O3 13.28 2.44
MnO 0.30
MgO 2.70 2.11 1.19
CaO 4.58 3.92 1.34
Na2O 2.96 3.88 3.33
K2O 1.24 2.95 6.1
LOI 0.68
total 99.33 100 100
a
FON1 is a metapelitic residue.
b
Sample exists of mesosomes (∼ 70 vol.%) and leucosomes
(∼ 30 vol.%); an average composition used for modelling in Fig. 8.
c
Composition taken from Naney (1983) and used in Fig. 11c.

granitoid migmatites. In the order of decreasing


abundance, the residual rock consists of plagioclase,
garnet, kyanite, biotite, rutile/ilmenite, staurolite, and
white mica. K-feldspar has not been observed, quartz
seems to be confined to a few veinlets of uncertain
significance. This bulk-rock composition is that of a
residue (high Al, but low K-contents and no quartz;
FON1 in Table 1). Melt extraction would here change
the bulk composition by increasing the Ca/Na- and Ca/
K- and decreasing the Al/Si-ratios in the residue (Burri,
2005). As discussed below, the main equilibration of the
solid phases occurred at pressures of 0.7–0.75 GPa and
temperatures of 670 °C. These conditions are outside
any hydrate-breakdown melting field. This is supported
by the presence of stable white mica and biotite in this
sample. However, the bulk composition of this sample
can be explained only with substantial melt extraction.
This is strong evidence that fluid infiltration occurred
into the meta-sedimentary rocks and melting without
dehydration reactions.

Fig. 5. Textures of amphibole-bearing migmatites. (a) Epidote phase


relations in amphibole-bearing leucosome. Bright low relief phases
are feldspar and quartz, dark phase is biotite, bright high relief phase
is epidote. Numbers: 1 denotes primary epidote included in biotite
grains. Note the idiomorphic shapes of these grains. 2 denotes primary
epidote partially resorbed along the upper grain boundary during
partial melting. 3 denotes a secondary, finer-grained epidote fraction,
probably formed during crystallisation of the melt. (b) Quartz-biotite
symplectite partially included in amphibole and interpreted as a result
of a mixed mode melting reaction leading to the formation of
amphibole. (c) Potassium X-ray-maps showing partial back-reaction
of igneous amphibole in granitic leucosome. Amphibole is replaced
by epidote, biotite and quartz. Part of the biotite is replaced by
secondary chlorite. Replacement of amphibole mainly occurs along its
rim or along the mineral cleavage.
604 A. Berger et al. / Lithos 102 (2008) 598–615

3.3. P–T conditions and melt fractions

Estimating P–T conditions of partial melting in meta-


granitoid rocks is difficult, because the equilibrated
assemblage at the stage of partial melting is not preserved
(melt + solid phases). In addition, assemblages in granites
are not very useful for thermobarometry. Some informa-
tion on P and T may be obtained from amphibole-bearing
migmatites, in which amphibole thermobarometry can be
applied (Holland and Blundy, 1994; Anderson and Smith,
1995). Burri et al. (2005) used amphibole-plagioclase
thermobarometry to infer equilibration conditions when
melt was present. Calculated equilibrium pressure and
temperature at the time of partial melting close to the
solidus are 0.7–0.8 GPa and 690°–740 °C (Fig. 6). As
discussed in the previous section, metapelitic migmatites
show equilibration of the solid minerals (sample Fon1).
Therefore, equilibration condition can be estimated by Fig. 7. Multi-equilibrium calculation for sample Fon1 (composition in
using compositions of the solid phases in this system. This Table 1) using the software TWQ. Dashed reactions are dominated by
results in pressures of 0.7–0.8 GPa and temperatures of staurolite or biotite endmembers siderophyllite and eastonite. Ellipse
670 °C (Fig. 7; using TWQ; Berman, 1991). These indicates inferred P–T conditions for this sample.
conditions reflect the equilibration of the solid, residue
material at intermediate stages of melting (Burri, 2005). et al., 2005). These samples were taken in the area, where
Some of the metasedimentary migmatites with low the highest P–T conditions have been inferred and where
leucosome volumes contain garnet. The compositions of muscovite-breakdown melting had occurred. The textures
garnet cores suggest conditions near 0.7 GPa and 700– of muscovite dehydration reactions support P–T condi-
750 °C, those of the rim near 0.6 GPa and 650 °C (Berger tions close to the white mica out reaction (Figs. 3c and 6).
The few examples with garnet can be used to estimate P–T
during partial melting are limited to a P–T interval
between 0.8–0.6 GPa and 740° and 650 °C (Fig. 6). The
variations of the mineral composition within a single
sample indicate changes of pressures (∼0.1 GPa) and
temperatures (∼90 °C) during partial melting along the
retrograde path.
In granite that is close to a minimum melt composition,
a nearly linear relation exists between water content and
melt fraction at given P and T below the water saturation
of the melt (Johannes and Holtz, 1996). Assuming a
pressure and temperature of 0.8 GPa and 700 °C, a melt
fraction of ∼0.25 would require some 2.5 wt.% of water
(Johannes and Holtz, 1996; White et al., 2001).
More significant are effects related to a change in bulk
composition from a granitic to a more granodioritic
composition. Phase relations in granodioritic systems
include more phases (e.g. Naney, 1983), but melt fractions
Fig. 6. Summary of P–T conditions estimated in the SSB (data are are still related to the water content. The relationship
from Burri et al., 2005; Berger et al., 2005 and this study). In some between Ca- and water-content can be illustrated by
samples a P–T history inside the melting field is indicated (see line mineral equilibrium models using a rock-composition of
between data-points). The white mica out conditions are indicated. In an amphibole-bearing, natural migmatite (see Fig. 8;
addition, the water-saturated solidus of the haplogranitic system is
shown (Johannes and Holtz, 1996). Most data indicate partial melting
sample Ar9902; Table 1). This rock contains leucosomes
in the stability field of white mica and melt. This is consistent with (mainly quartz + two feldspars), mesosomes (quartz, two
observations in the migmatites of the SSB. feldspars, biotite and amphibole) as well as some bands
A. Berger et al. / Lithos 102 (2008) 598–615 605

without amphibole (quartz, two feldspars and biotite). (de Capitani and Brown, 1987) show that the fraction of
Table 1 shows the average bulk composition from an XRF melt is mainly a function of the amount of water, which
analysis of the leuco- and mesosomes in this rock. The has to be related to the influx of fluids (Gardien et al.,
results of the mineral equilibria model using DOMINO 2000). This fluid has to be rich in H2O, because at 700 °C
aH2O b 0.9 would prevent the generation of melt (Fig. 8a;
Johannes and Holtz, 1996). Thus the presence of
amphibole adds further constraints on the (P)–T–X
condition of these migmatites (Fig. 8b). In summary, the
results from mineral-equilibrium models agree well with
the following observations: (1) the distribution of
leucosomes is inhomogeneous, indicating influx of
water as the cause of melting; (2) small changes in bulk
composition affect the stable assemblage, i.e. amphibole-
bearing versus amphibole absent bands in these rocks.
The melt fraction increases with decreasing bulk Ca-
contents, because more water is stored in amphiboles in
the case of higher Ca-content (Fig. 8b). This is consistent
with the observed different volumes of leucosomes in
different rocks.

4. Geodynamic setting during water-assisted melting


in the Alps

Water-assisted melting requires two things: an


efficient source of water and a force that drives fluids
into the rock at the P and T conditions of interest. Fluids
can only be released and mobilized at conditions below
the solidus, because fluids would otherwise be used up
in situ to produce melt. If the permeability distribution is
homogeneous and fluid flow is solely driven by density
differences due to temperature variations, a vertical flow
component would be promoted. Thus an additional
driving force is required that promotes up-temperature
flow of fluid into granitoid rocks. There is evidence that
fluid flow was related to the deformation and perme-
ability changes inside the Southern Steep Belt (e.g.
Berger et al., 1996; Burri et al., 2005). This means that
the process of partial melting has to be viewed in the
context of the tectonic movements in the Southern Steep
Belt.

4.1. Structural history


Fig. 8. Binary T–X diagrams using DOMINO (de Capitani and Brown,
1987) for composition of an amphibole-bearing migmatite (sample
The Tertiary migmatite belt of the Central Alps is
Ar9902). Light gray shading indicates the presence of free fluid,
intermediate shading a melt and dark grey shading coexistence of fluid spatially related to the Southern Steep Belt (e.g. Burri
and melt. Composition of the main composition is Si(22.61), Al(6.81), et al., 2005). The tectonic scenario at the time of partial
Ca(1.44), Fe(1.26), Mg(1.08), K(1.29), Na(2.58), H(8), O(65.26) melting (Oligocene/Miocene) is characterized by trans-
(a) Water activity versus temperature diagram (b) Binary T–X diagram pressional backthrusting of the Southern Steep Belt
for variable Ca-content (main composition as given above and minus 1
against the colder Southern Alps (Figs. 1 and 9). Tectonic
CaO for the lower composition). Melt proportions of the system are
indicated by the numbers inside the boxes. Note the variable melt models indicate a steep contact between the cold Southern
volumes at a given P and T in relation to the Ca-content and the related Alps and the uplifted, hot Southern Steep Belt during the
paragenesis. Paragenesis (3) fits the observed one. Oligocene (e.g., Schmid et al., 1989, 1996; Pfiffner et al.,
606 A. Berger et al. / Lithos 102 (2008) 598–615

Fig. 9. Tectonic sketch of the Alps at Oligocene times during partial melting (simplified after Schmid et al., 1996). The position of the fluid modelling
box is indicated.

2000). In the north, the Southern Steep Belt is part of the with the Insubric Line (Fig. 9). In the northern limb of this
folded units of the adjacent nappe pile, whereas in the anticline some relics of Mesozoic cover are preserved.
center and the south it is characterized by a highly variable More of these sedimentary rocks are expected to remain
mix of rock types. The Southern Steep Belt was strongly deeper in the crust along the northern limb (e.g.,
deformed under amphibolite facies conditions, with many Schönborn, 1997). At the time of the indenter tectonics,
units grading into the ductile to brittle mylonites of the these meta-sedimentary rocks came into contact with the
Insubric Line, which are an expression of localized defor- Southern Steep Belt, as they are situated today. During
mation (e.g., Schmid et al., 1989; Handy et al., 2005). later transpressional backthrusting this crustal segment
The structures related to backthrusting, strike-slip was uplifted and exposed by localized deformation along
movements, orogen-parallel normal-faulting and folding the Insubric Line. Therefore, the level of erosion is deeper
indicate that these processes occurred more or less north of the Insubric Line then in the Southern Alps.
simultaneously. Localized deformation at lower tempera- Therefore, the migmatites of the SSB, are today at the
tures offers the possibility to look at the lower section of present erosional surface, whereas the Tertiary amphib-
this transpressional zone, which is exposed north of the olite facies rocks of the Southern Alps remain deeply
low-grade mylonites. The melting event produced small buried.
dykes, which crosscut the steep structures and were folded
at a post-nappe stage. Therefore, partial melting must 4.2. Some remarks on the thermal structure
have occurred during transpressional deformation (Burri
et al., 2005). The width of the in situ melting zone is Fluid production and melting were partly controlled
approximately 3–7 km. At the time of partial melting, the by heating of individual units along their P–T–t paths.
Southern Steep Belt was already in a steep position A likely cause for water-assisted melting would be the
(Berger et al., 1996). development of inverted isotherms. In such a situation,
The transpressional deformation is related to the metamorphic dehydration reactions occurring outside
deformation of the Adria-indenter. The deformation of the melting field would produce fluids, which ascend
the indenter involved thin-skinned thrusting and large- into the melting field (upwards flow from a colder to
scale folding to produce a large anticline at the contact a hotter area). Such a situation is commonly realized
A. Berger et al. / Lithos 102 (2008) 598–615 607

Fig. 10. Compilation of some published thermal models for the Alps. (a) Thermal model “MELONPIT” from Roselle et al. (2002). Thermal structure
redrawn from their Fig. 10d after 70 Ma of elapsed model-time. The Insubric Line is included in the model. (b) Thermal structure of the “lower
lithosphere” model of Goffé et al. (2003). The figure is redrawn from their Fig. 6 using the 30 mm/year convergence rate. (c) Isotherms of the “Hydr”
model from Stöckhert and Gerya (2005) redrawn from their Fig. 5. The elapsed model-time is 29 Ma. The Insubric Line is drawn in the Figure, but not
taken into account in the model. (d) Thermal structure of model 10 from Burg and Gerya (2005) after 25 Ma of elapsed model-time and shortening of
500 km (their Fig. 4d). The Insubric Line is not included in this model. All models include a multiple layer model, which is not redrawn in this figure,
but the mantle is indicated in grey. The area of interest for partial melting is marked by a circle. Note, that the absolute temperatures are not sufficient
for melting at the pressures of interest. However, the relative thermal structure can be compared. In all models still overturned isotherms are computed
at great depth, below the mantle wedge of the Southern Alps. These overturned isotherms are not relevant for melting in the Central Alps.
608 A. Berger et al. / Lithos 102 (2008) 598–615

during subduction. The length and time scale of such are addressed at the end of this section. A major
overturned isotherms depend on numerous parameters possibility for a fluid source is metamorphic dehydra-
(e.g., subduction rate, age of the crust, radioactive heating, tion reactions at conditions cooler than where the first
shear heating; see Roselle et al., 2002; Gerya et al., 2002; melts are produced (see also section on thermal history).
Stöckhert and Gerya, 2005; Burg and Gerya, 2005; The amount of dehydration fluid produced depends on
Fig. 10). However, water-assisted melting in the Alps the rock-type and the P–T path (e.g. Thompson, 2001;
occurred during backthrusting, long after the main stage Burri, 2005). We have used mineral-equilibrium mod-
of subduction was over. Subduction of the lower crustal elling to quantify fluid volumes that could be produced
parts has nevertheless been assumed to be ongoing during under different conditions. The P–T path of our model is
backthrusting (Figs. 9 and 10), and this needs to be constrained by the estimated P–T conditions during
considered in combination with the major temperature melting in the migmatites. The composition of the rocks
difference between the Central Alps and the Southern acting as fluid sources is poorly known.
Alps. Such a thermal setup is also indicated by the One possibility is that fluids were derived from
proposed indenter tectonics (see last section). Available metapelites, which constitutes the most effective source
thermal modelling gives insight into the thermal structure, of hydrous fluids. Accounting for our thermal model
during the temporal sequence from subduction to setup, the pressure of the dehydration reactions has to be
collision, backthrusting, and erosion (Fig. 10). set equal or slightly higher, whereas temperatures must
Thermal modelling indicates that spatially condensed be lower than within the migmatites. Assuming
isotherms developed in the upper and middle crust and pressures between 0.8–0.75 GPa and temperatures
that overturned isotherms may have persisted in the below 640 °C and using a ΔT of 50° (i.e. a temperature
lower crust and the mantle over prolonged periods (e.g., path from 590–640 °C) and a pressure of ∼ 0.75 GPa,
Roselle et al., 2002; Goffé et al., 2003; Burg and Gerya, some 1.3 vol.% water could be produced from an
2005). If so, this would have led to a situation, where average metapelite. This value increases with decreasing
at similar depth a hotter orogenic part was adjacent Al-content of the metapelites, because the Al-content
to a colder upper plate in the zone of backthrusting. correlates with the modal content of micas. One volume
However, these contemporaneous movements, i.e. unit (e.g. 1 km3) of such a metapelite is able to produce a
backthrusting in the upper crust and subduction of melt fraction of ∼ 0.12 in the same volume of a near
lower crust (Fig. 10) would have led to a fairly uncertain minimum melt granite at 690 °C and 0.8 GPa (see also
thermal structure in the lowest part of the crust. Table 2). To account for the average melt fraction
Regardless of this uncertainty, however, potentially documented in the migmatite belt of the central Alps, the
overturned isotherms would result at a depth of 75– metapelite volume required would be 1.6 times the
100 km, whereas the depth of melt production has been migmatite volume. All of these calculations assume
found to be ∼20–25 km (see also Fig. 10). Therefore, water-saturated equilibrium. In addition to continuous
we surmise that this thermal structure is not directly equilibrium, we can also infer overstepping of any
relevant to this problem, as any potential fluid released reaction by kinetic reasons. Any overstepping of
in the subducted crust in the area of overturned isotherms dehydration reaction survives hydrate phases to higher
would related to the mantle wedge or the lower crust, temperatures. In this scenario, even smaller rock-
rather than the zone of the migmatites investigated in this volumes would produce the necessary amount of fluids
study. In the area of interest, the thermal structure (cf. Buick et al., 2004). Metapelites constitute an
developed in the middle crust is likely to have been
layered normally. This means at the time of melting,
the rocks below the migmatites were hotter then the Table 2
Inferred flux of water during water-assisted melting
migmatites itself. Therefore, rocks with a similar solidus
as the migmatites and located below these could not Migmatite Bergell Metapelite in the
belt pluton Southern Alps
have been the fluid source for the migmatites.
Measured or required 6.96e + 08 1.91e+07 1e + 09
area (m2)
5. Fluid sources and modelling of fluid flow pathways
Percentage of produced −100 10.6 89.4
(or consumed) water
5.1. Fluid sources and fluid production Percentage of consumed 2.1 4 1.3
or released H2O (wt. %)
Several potential fluid sources exist. We first discuss Using a two dimensional map view through the problem (inferring a
the most promising ones, and those of minor importance simple geometry in the third dimension).
A. Berger et al. / Lithos 102 (2008) 598–615 609

increasingly potent fluid source, the higher the temper-


ature difference between source region and melting
region.
A second possible source of aqueous fluids is
hydrated metabasic rocks, which notably lose much of
their chlorite at the transition from the greenschist to the
amphibolite facies (see also Elmer et al., 2006). The
amount of water that can be released due to the
breakdown of chlorite is high, but the temperature at
which this occurs is relatively low. If one considers
metabasic rocks to be the source of the fluids, the
temperature difference between source and melting
region has to be higher than is likely to have been the
case in the present scenario.
In addition to the infiltration of external water, fluid
recycling inside the crystallizing migmatites has to be
accounted for. This process has been explored in terms
of crystallization of granites by Johannes and Holtz
(1996). We modeled a crystallizing haplogranitic melt
from PT conditions of 710 °C and 0.8 GPa to 600 °C and
0.7 GPa (Fig. 11). During crystallization of this melt, a
fluid phase will be released at ∼ 640 °C (position B in
Fig. 11). The remaining crystal mush has a melt volume
of 78 vol.%. If the released fluid flows into a
leucogranite, it would produce ∼ 20 vol.% melt again.
This melt batch will always be water undersaturated,
and water from a second melt batch would only be
released at the solidus. This process only occurs close to
the solidus (e.g. release of water at 640 °C with a solidus
at 630 °C). Fig. 11c shows that the evolution of a rock of
granitic composition (granite in Table 1) is similar to a Fig. 11. Mineral-equilibrium models for ascending and crystallizing
rock of haplogranitic composition. Therefore, water melt using THERIAK. (a) sketch of the crystallization evolution of
recycling may be an additional factor in the total melt ascending and crystallizing aplites in a migmatite. Position A, B and C
production and may lower the volume of external fluxes are the situation as modelled in (b) and (c). (b) evolution of modal
contents of minerals, fluids and melt along a crystallization path of a
necessary to produce a specific volume of melt.
water-saturated haplogranitic melt. (c) similar as (b) but with a
However, water recycling plays a role only at tempera- complete granitic composition (see Table 1 for composition).
tures close to the solidus.
There are two other processes that may account for
part of the in situ production of fluids. One possibility is source (transforming mainly from a garnet-omphacite to
the direct release of fluids from rocks inside the SSB. In plagioclase-amphibole assemblages). Metabasic rocks,
other words, rocks with higher solidus release water into which underwent heating, could release modest amounts
their immediate surroundings. This would be an of fluids in the amphibolite facies (below the solidus of
effective water source, because the SSB comprises in amphibolites). This fluid source is related in minor part
large part a lithospheric mélange containing different to sliding reactions including the amphiboles and, under
rock types. However, metapelites have a similar solidus certain circumstances, to the breakdown of epidote
temperature as the surrounding metagranitoids and thus (Schmidt and Poli, 1998; Elmer et al., 2006). However,
can be ruled out, because any fluid from prograde the volume percentage of fluids is low and the volume
dehydration reactions of such metapelites would lead to fraction of basic rocks in the SSB is also minor.
in situ melting, rather than releasing mobile hydrous Therefore, basic rocks are no effective fluid source at
fluids into their surroundings. Similarly, metabasic the conditions of interest. The second minor possibility
rocks undergoing decompression from the eclogite for fluids inside the SSB is the release of fluids from
facies may serve as a fluid sink, but not as a fluid crystallizing magmas, such as the neighboring Bergell
610 A. Berger et al. / Lithos 102 (2008) 598–615

as a result of dehydration reactions. We did not assign


hydrological properties to the contact between the
Central and the Southern Alps. Pressure and temperature
were fixed along the top boundary at 0.65 GPa and
570 °C, respectively. Thus fluid was allowed to leave or
enter the domain through the top. The side boundaries
were impermeable and adiabatic. The temperature at the
bottom boundary was fixed at 750 °C and 720 °C at the
bottom of the left and right model block, respectively.
This accounts for the colder Southern Alps in nature,
which produce aqueous fluids but no melt. At the
bottom of the domain a conductive low-permeability
layer was set to prevent fluid exchange through the
bottom boundary.
Fig. 12. Model design, rock units and boundary condition of the fluid The initial condition was a uniform temperature of
flow model between the Southern Alps and the Central Alps. Compare 570 °C throughout the domain and a hydrostatic
set up with the section of the Alps in Oligocene times as presented in pressure distribution. Each simulation consisted of two
Fig. 9.
parts. First we carried out an initialization run to
calculate a steady state temperature and pressure
pluton (Berger et al., 1996). The measured exposure area distribution without fluid production. Then we switched
of the Bergell pluton today is 20 km2 (Table 2); assuming on the fluid production in the lower meta-sediments and
an initial water content of 4 wt.% for this magma, a let the system run to steady state (i.e. 1 Ma). We simulate
maximum of 10 wt.% of the required aqueous fluids a base-case scenario (see parameters in Table 3) and
could have been released from this pluton (Table 2). So, compared the results with scenarios that used different
the Bergell pluton is too small to account for the total permeabilities.
influx of fluids, though it may have played a supporting The initialization of the base-case scenario (i.e.
role. steady state flow and temperature fields without fluid
production) shows that fluids circulate in an isolated
5.2. Fluid flow rates and pathways convection cell within the more permeable lower unit
(Fig. 13a). It should be noted that even though this unit
We designed a flow model based on the tectonic has the highest permeability, a permeability of 10− 16 m2
scenario at the time of the melting event as described is quite low, and it is remarkable that the fluid convects.
above. We used the high-temperature flow code Hy- The fact that convection occurs has to be related to the
drotherm (Hayba and Ingebritsen, 1994) to explore fluid fluid properties (density and viscosity) at the given P
flow patterns and calculate flow rates as a function of the and T conditions and, as will be shown below, the
permeability and temperature distribution in the rock. somewhat higher geotherm in the left model block. As
The model domain represents an 8.0 km by 6.2 km soon as fluid production in the right unit starts, the fluids
cross-section (Fig. 12). The right model block consists infiltrate the surrounding rocks, preferentially those with
of two hydrologically different units: (1) a low- the next highest permeability (Fig. 13b). Fluid flow is
permeability top and (2) a higher permeability bottom
unit (representing the tectonic situation of the Central Table 3
Alps and the Southern Alps). The left model block has Thermal and hydrological parameters for the base-case (Case A)
scenario
an intermediate permeability. We use data from the
literature to constrain permeability values (Fig. 12 and Left Upper part of the Lower part of the
block right block right block
Table 3). In general, it can be assumed that deforming
rocks have higher permeability than other rocks (see Permeability 1e − 19 1e − 22 1e −16
(x, y, z) (m2)
Dipple and Ferry, 1992; Ferry, 1994; Oliver, 1996;
Thermal 3.0 3.0 3.0
Hanson, 1997 and references therein). The model does conductivity
not include possible permeability changes during (J/m s K)
metamorphic dehydration or due to hydro-fracturing Porosity 0.025 0.005 0.05
during fluid infiltration (for discussion see Miller et al., Fluid production 0.0 0.0 5e −4
(kg/year m3)
2003). We assumed that the bottom unit produced water
A. Berger et al. / Lithos 102 (2008) 598–615 611

Fig. 13. Steady state flow pattern, flow velocities. Contour values reflect the magnitude of the flow velocity in m/a. (a) Scenario without fluid
production. (b) Base case scenario (see text for explanation and Table 3 for input parameters). (c) Simulation with a uniform geothermal gradient of
25 °C/km. (d) The upper meta-sediments have a permeability two orders of magnitude higher than in the base-case scenario (see panel b). (e) The left
model block (Southern Steep Belt) has a permeability two orders of magnitude higher than in the base-case scenario.

now driven by the pressure increase in the source rocks indicated by relatively high flow rates along the bottom
of the right model block due to fluid production in boundary and at the contact with the left model block
addition to the buoyancy component that causes the (Fig. 13b). Fluid flow from the source rocks into the
fluid to convect in the case of no fluid production neighboring left model block is enhanced by the low-
(Fig. 13). The sum of these processes controls the flow permeability upper part of the right model block, which
pattern. The fluid flow pattern in the source rocks inhibits vertical flow. Fluid infiltration into the left block
preserves a buoyancy-driven component, which is extends over several kilometers (Fig. 13b). Flow
612 A. Berger et al. / Lithos 102 (2008) 598–615

Table 4 model block), this reduction amounts to some 40%. This


Variations in input parameters for the different scenarios of the suggests that the permeability of the sealing layer does
hydrological modelling (Fig. 13)
not have to be much lower than that of the left model
Permeability Permeability Permeability Other block to increase the infiltration rate into the left block,
(m2) of the left (m2) of the (m2) of the changes
model block upper right lower right
significantly. If we increase the permeability of the left
block block model block by two orders of magnitude and leave the
permeability of the upper part of the right model block
Case B 1e − 19 1e − 22 1e − 16 Temperature
change (Case D; Table 4), the highest infiltration rates occur
Case C 1e − 19 1e − 20 1e − 16 no above the bottom boundary. Here, flow is directed
Case D 1e − 17 1e − 22 1e − 16 no towards the left model block (Fig. 13e) as a result of the
Note, that case A is the base-case scenario. buoyancy component, as discussed above. As the
permeability contrast between the left and right model
block is reduced, lower fluid pressure gradients are
velocities on the order of 0.2–0.3 m/year occur at the required to move fluids across the contact. The high
contact between the right and the left model block, infiltration rate at the bottom reduces the infiltration at
whereas along most of the contact fluids infiltrate at the top of the fluid source region. Because infiltration
velocities on the order of 0.1–0.2 m/year. now goes deeper, a larger region of the northern model
We then explored forced versus buoyancy-driven block is affected by fluid infiltration (Fig. 13e).
flow and the significance of rock-dependent geotherms
in more detail by running a simulation with a uniform 5.3. Implications of the fluid flow models for the Alps
geothermal gradient (Tables 3 and 4). The results from
the initialization run show that fluids no longer convect The fluid flow models are build up to be compared
in the lower part of the source region of the model. This with a cross-section through the border between the
emphasizes the significance of the higher geotherm in Southern and the Central Alps. The model represents an
the left model block for buoyancy-driven convection to area in the middle crust including a major boundary in
occur on the southern side. There is no flow and purely the center of the model. This boundary should represent
conductive heat dissipation throughout the domain. the change in temperature and rock properties across the
With fluid production turned on, the flow pattern and Insubric Line (compare also Figs. 12 and 9). The right
mass flux distribution in the lower part of the right block model block should compare with the Southern Alps,
is consistent with the lack of a buoyancy component. which are subdivided into a lower and upper part. The
Flow velocities within the lower part of the right model lower part contains the source region, where the
block increase steadily towards the contact between the metapelitic rocks of the Southern Alps are dehydrating,
upper and lower right and left model block (Fig. 13c), and the upper part should represent the low-grade
and do not show the high values along the bottom basement and sedimentary rocks, which serve as a
boundary and the contact between the lower right block hydrological seal. The left model block represents the
and the left model block visible in Fig. 13b. However, Southern Steep Belt of the Central Alps. The main
the distribution of mass fluxes inside the left model problem is the transfer from used average permeabilities
block is the same, regardless of the geothermal gradient. into the geological situation. Permeability is defined for
This means that the distribution of mass fluxes becomes porous media flow, but in crystalline rock masses, also
more uniform across the contact, and we should see fracturing may occur (see discussion above). However,
similar fluid/rock ratios on both sides of the contact. it has been discussed, that permeability can be used in
To assess the importance of the upper part of the right fractured medium using an average over large volumes.
model block as a hydrological seal, we then increased In this context, the modeled fluid volumes as well as the
the permeability of the upper sediments by two orders of fluid pathways can be only used to look on large-scale
magnitude (Case C; Table 4). The results show that average values. In nature, several parameters refine this
fluxes from the lower part of the right block into the average picture. However, using the different models
overlying units are now higher, and infiltration from the presented above, it is evident that mainly the perme-
lower part into the left model block is slightly reduced ability will control the fluid pathways and fluid flow
compared to the base-case and tends to be more uniform rates in such a geodynamic setting. We tested the role of
along the contact (Fig. 13d). The highest mass flux is the assumed temperature difference between the SSB
about 16% lower than in the base-case scenario. If we and the Southern Alps, which influences also the fluid
decrease the permeability to 10− 19 m2, (same as the left pathways, but the general flow pattern and mass fluxes
A. Berger et al. / Lithos 102 (2008) 598–615 613

are similar (Fig. 13c). The role of a hydrological seal by between and within the different rocks (see Fig. 13). The
upper parts of the anticline in the Southern Alps is tested stable flow pattern is less dependent on absolute
in model case C. The absolute values of the permeability permeability values, but rather on the relative permeabil-
are unknown, but even at only slight permeability ity between the lithological units in the Southern Alps
differences, the fluid will infiltrate into the SSB. The and the migmatite belt in the Central Alps. An increase in
role of the upper hydrological seal will only change the fluid pressure in the lower units of the Southern Alps due
mass fluxes. The opposite effect can be simulated by to fluid production leads to infiltration of water into the
increasing the permeability in the migmatite belt southern parts of Central Alps (Fig. 13).
(Case D). The detailed permeability distribution also Following partial melting in the Southern Steep Belt,
influences the detailed flow pathways (Fig. 13e). the melts underwent cooling and crystallization. The
The scenarios show that the permeability distribution melts release their solved water at the solidus in a narrow
of the three model units controls the distribution of fluid temperature interval, producing high fluid volumes at a
infiltration into the simulated migmatite belt. Large nearly constant temperature and in a short time. Some
permeability contrasts tend to focus fluid flow and exceptions are the observed pegmatites, which are late in
infiltration, promoting localized higher water/rock ratios the structural evolution (Wenk, 1970; Burri et al., 2005).
within the Southern Steep Belt, whereas low-perme- These may be local melting related to the released water
ability contrasts promote more uniform infiltration at and melt enriched on incompatible elements, which have
lower rates. A buoyancy component, if present, may a significant lower solidus (e.g. London, 1992). The high
thus have had a strong effect on the distribution of fluid production rates would have favored hydrofractur-
infiltrated fluids and on water/rock ratios in the Southern ing in the overlying solid metamorphic rocks (e.g. Miller
Steep Belt. et al., 2003). Hydrofracturing is an effective way to
transport fluids away in a short time interval, making
6. Summary and discussion fluids less available at conditions close to the solidus.
This may explain that substantial retrograde reactions
The combination of petrological and structural near the solidus are rarely observed in the migmatites of
observations indicates that water-assisted melting is an the Central Alps. Additional evidence for hydrofracturing
important feature in the Tertiary migmatite belt of the is given by hydrothermal veins developed in the amphibo-
Central Alps. Metapelitic rocks incorporated in the SSB lite facies in direct vicinity to the migmatite belt (e.g.
are not a realistic fluid source, because, at the conditions Klein, 1976).
of partial melting, the released aqueous fluid would lead
to in situ partial melting and no fluid would be released Acknowledgements
into the directly adjacent or overlying rocks. A major
fluid source directly below the migmatite belt can be We thank M. Brown and F. Bussy for constructive
also excluded, because of the steep solidus of most and careful reviews. Schweizerischer Nationalfonds has
relevant systems in P–T space. supported our research over several years (2000-
The situation at the time of melting involved indenter 055306.98, 20-63593.00, 20020-101826, and 200020-
tectonics of the Southern Alps into the Central Alps (Fig. 9, 109637). Mike Brown also introduce the term hydrate-
Schmid et al., 1989, 1996; Pfiffner et al., 2000). The most breakdown melting. He is further acknowledged for this
likely scenario is that the fluids were primarily generated important improvement of terminology.
through dehydration of metapelites in the Southern Alps,
where temperatures were lower than in the Central Alps. References
With the help of petrological data, we can estimate the
average water contents infiltrated into these migmatites. Anderson, J.L., Smith, D.R., 1995. The effects of temperature and fO2
Experimental data from a haplogranitic system and on the Al-in-hornblende barometer. American Mineralogist 80,
549–559.
thermodynamic modelling indicate that water contents Berger, A., Kalt, A., 1999. Structures and melt fractions as indicators of
were on the order of 1–3 wt.% (or ∼5 vol.%). These the rheology of cordierite bearing migmatites of the Bayerische Wald
estimated values depend on the accuracy of the inferred (Vasican belt, Germany). Journal of Petrology 40, 1629–1639.
leucosome (melt) volumes and bulk composition of the Berger, A., Rosenberg, C., Schmid, S.M., 1996. Ascent, emplacement
protolith (and additional uncertainties from water recycling and exhumation of the Bergell Pluton within the Southern Steep
Belt of the Central Alps. Schweizerische Mineralogische und
inside the migmatites near the solidus). Petrographische Mitteilungen 76, 357–382.
Fluid circulation models show that pathways of the Berger, A., Rosenberg, C.L., 2003. Preservation of chemical residue-
mobilized fluid depend on the permeability distribution melt equilibria in natural anatexite: the effects of deformation and
614 A. Berger et al. / Lithos 102 (2008) 598–615

rapid cooling. Contributions to Mineralogy and Petrology 144, Garlick, S.R., Gromet, L.P., 2004. Diffusion creep and partial melting in
416–427. high temperature mylonitic gneisses, Hope Valley shear zone, New
Berger, A., Scherrer, N.C., Bussy, F., 2005. Equilibration and England Appalachians, USA. Journal of Metamorphic Geology 22,
disequilibration between monazite and garnet: indication from 45–62.
phase-composition and quantitative texture analysis. Journal of Gebauer, D., 1996. A P–T–t Path for a high pressure ultramafic rock-
Metamorphic Geology 23, 865–880. associations and their felsic country-rocks based on SHRIMP-
Berman, R.G., 1991. Thermobarometry using multi-equilibrium Dating of magmatic and metamorphic Zircon domains. Example:
calculations: a new technique with petrological applications. Alpe Arami (Central Swiss Alps). In: Hart, A., Basu, S.R. (Eds.),
Canadian Mineralogist 29, 833–855. Earth Processes: Reading the Isotope Code, vol. 95, pp. 307–328.
Braun, I., Raith, M., Ravindra Kumar, G.R., 1996. Dehydration Gerya, T.V., Stöckert, B., Perchuk, A.L., 2002. Exhumation of high-
melting phenomena in Leptynitic gneisses and the generation of pressure metamorphic rocks in a subduction channel: a numerical
Leucogranites: a case study from the Kerala Khondalite Belt, simulation. Tectonics 21, 1056.
Southern India. Journal of Petrology 37, 1285–1305. Goffé, B., Bousquet, R., Henry, P., Le Pichon, X., 2003. Effect of
Brown, M., 1979. The petrogenesis of the Saint Malo migmatite belt, the chemical composition of the crust on the metamorphic evolution
Armorica massif, France. Neues Jahrbuch für Mineralogie 135, of orogenic wedges. Journal of Metamorphic Geology 21, 123–141.
48–74. Gutzwiller, E., 1912. Injektionsgneisse aus dem Kanton Tessin.
Brown, M., 2004. The mechanism of melt extraction from lower Unpubl. PhD, Universität Zürich, 64 pp.
continental crust of orogens. Transactions of the Royal Society of Handy, M.R., Babist, J., Rosenberg, C.L., Wagner, R., Konrad, M.,
Edinburgh. Earth Sciences 95, 35–48. 2005. Decoupling and its relation to strain partitioning in
Brown, M., Dallmeyer, R.D., 1996. Rapid Variscan exhumation and continental lithosphere — insight from the Periadriatic fault
the role of magma in core complex formation; southern Brittany system (European Alps). In: Brun, J.P., Cobbold, P.R., Gapais, D.
metamorphic belt, France. Journal of Metamorphic Geology 14, (Eds.), Deformation Mechanisms, Rheology and Tectonics.
361–379. Geological Society, London, Special Publications, pp. 249–276.
Buick, I.S., Stevens, G., Gibson, R.L., 2004. The role of water Hanson, R.B., 1997. Hydrodynamics of regional metamorphism due to
retention in the anatexis of metapelites in the Bushveld Complex continental collision. Economic Geology 92, 880–891.
aureole, South Africa; an experimental study. Journal of Petrology Harris, N.B.W., Caddick, M., Kosler, J., Goswami, S., Vance, D., Tindle,
45, 1777–1797. A.G., 2005. The pressure-temperature-time path of migmatites from
Burg, J.-P., Gerya, T.V., 2005. The role of viscous heating in Barrovian the Sikkim Himalaya. Journal of Metamorphic Geology 22, 249–264.
metamorphism of collisional orogens: thermomechanical models Hayba, D.O., Ingebritsen, S.E., 1994. The Computer Model Hydro-
and application to the Lepontine Dome in the Central Alps. Journal therm, A Three Dimensional Finite Difference Model to Simulate
of Metamorphic Geology 23, 75–95. Ground-Water Flow and Heat Transport in the Temperature Range
Burri, T., 2005. From high-pressure to migmatisation: on orogenic of 0 to 1200 °C. U.S. Geological Survey.
evolution to the Southern Lepontine (Central Alps of Switzerland/ Heinrich, C.A., 1982. Kyanite-eclogite to amphibolite facies evolution
Italy). PhD Thesis, Universität Bern, Bern, 150 pp. of hydrous mafic and pelitic rocks, Adula nappe, Central Alps.
Burri, T., Berger, A., Engi, M., 2005. Tertiary migmatites in the Central Contributions to Mineralogy and Petrology 81, 30–38.
Alps: regional distribution, field relations, conditions of formation Holland, T., Blundy, J., 1994. Non ideal interactions in calcic
and tectonic implications. Schweizerische Mineralogische Petro- amphiboles and their bearing on amphibole-plagioclase thermom-
graphische Mitteilungen 85, 215–232. etry. Contributions to Mineralogy and Petrology 116, 433–447.
Butler, R.W., Harris, N.B.W., Whittington, A.G., 1997. Interactions Holland, T.J.B., Powell, R., 1998. An internally consistent thermody-
between deformation, magmatism and hydrothermal activity namic data set for phases of petrological interest. Journal of
during active crustal thickening: a field example from Nanga Metamorphic Geology 16, 309–343.
Parbat, Pakistan Himalayas. Mineralogical Magazine 61, 37–52. Holland, T.J.B., Powell, R., 2001. Calculation of phase relations
Castro, A., Corretgé, L.G., El-Biad, M., El-Hmidi, H., Fernandez, C., involving haplogranitic melts using an internally consistent
Patiño Douce, A.E., 2000. Experimental constraints on hercynian thermodynamic data set. Journal of Petrology 42, 673.
anatexis in the Iberian Massif, Spain. Journal of Petrology 41, Johannes, W., Holtz, F., 1996. Petrogenesis and Experimental
1471–1488. Petrology of Granitic Rocks. Springer. 335 pp.
de Capitani, C., Brown, T.H., 1987. The computation of chemical Johnson, T.E., Gibson, R.L., Brown, M., Buick, I.S., Cartwright, I.,
equilibrium in complex systems containing non-ideal solutions. 2003. Partial melting of metapelitic rocks beneath the Bushveld
Geochimica et Cosmochimica Acta 51, 2639–2652. Complex, South Africa. Journal of Petrology 44, 789–813.
Dipple, G.M., Ferry, J.M., 1992. Metasomatism and fluid flow in Johnson, T.E., Brown, M., Gibson, R., Wing, B.A., 2004. Spinel-
ductile fault zones. Contributions to Mineralogy and Petrology cordierite symplectites replacing andalusite; evidence for melt-
112, 149–164. assisted diapirism in the Bushveld Complex, South Africa. Journal
Elmer, F.L., White, R.W., Powell, R., 2006. Devolatilization of of Metamorphic Geology 22, 529–545.
metabasic rocks during greenschist-amphibolite facies metamor- Kalt, A., Berger, A., Blümel, P., 1999. Metamorphic evolution of
phism. Journal of Metamorphic Geology 24, 497–513. cordierte bearing migmatites of the Bayerische Wald (Varsican
Engi, M., Berger, A., Roselle, G., 2001. The role of the tectonic belt, Germany). Journal of Petrology 40, 601–627.
accretion channel in collisional orogeny. Geology 29, 1143–1146. Klein, H.-H., 1976. Alumosilikatführende Knauern im Lepontin.
Ferry, J.M., 1994. A historical review of metamorphic fluid flow. Schweizerische Mineralogische und Petrographische Mitteilungen
Journal of Geophysical Research 99, 15487–15498. 56, 435–456.
Gardien, V., Thompson, A.B., Ulmer, P., 2000. Melting of biotite+ Köppel, V., 1993. The Lepontine area, a geochronological summary.
plagioclase+quartz gneisses: the role of H2O in the stability of In: von Raumer, J.F., Neubauer, F. (Eds.), Pre-Mesozoic Geology
amphibole. Journal of Petrology 41, 651–666. in the Alps, pp. 345–348.
A. Berger et al. / Lithos 102 (2008) 598–615 615

London, D., 1992. The application of experimental petrology to the Schmid, S.M., Pfiffner, A., Froitzheim, N., Schönborn, G., Kissling,
genesis and crystallization of granitic pegmatites. Canadian N., 1996. Geophysical-geological transect and tectonic evolution
Mineralogist 30, 499–540. of the Swiss-Italian Alps. Tectonics 15, 1036–1064.
Miller, S.A., van der Zeeb, W., Olgaard, D.L., Connolly, J.A.D., 2003. Schmidt, M.W., Poli, S., 1998. Experimentally based water budgets for
A fluid-pressure feedback model of dehydration reactions: dehydrating slabs and consequences for arc magma generation.
experiments, modelling, and application to subduction zones. Earth and Planetary Science Letters 163, 361–379.
Tectonophysics 370, 241–251. Schönborn, G., 1997. Alpine tectonics and kinematic models of the
Milord, I., Sawyer, E.W., Brown, M., 2001. Formation of diatexite Central Southern Alps. Memoire degli Istituti Geologia e Miner-
migmatite and granite magma during anatexis of semi-pelitic alogia dell' Universita di Padova 44, 229–393.
metasedimentary rocks: an example from St. Malo, France. Journal Spear, F.S., Kohn, M.J., Cheney, J.T., 1999. P–T paths from anatectic
of Petrology 42, 487–505. pelites. Contributions to Mineralogy and Petrology 134, 17–32.
Mogk, D.W., 1992. Ductile shearing and migmatization at mid crustal Stöckhert, B., Gerya, T.V., 2005. Pre-collisional high pressure
levels in an Archean high grade gneiss belt, northern Gallatin range, metamorphism and nappe tectonics at active continental margins:
Montana, USA. Journal of Metamorphic Geology 10, 427–438. a numerical simulation. Terra Nova 17, 102–110.
Montel, J.M., Marignac, C., Barbey, P., Pichavant, M., 1992. Thermo- Thompson, A.B., 2001. Clockwise P–T paths for crustal melting and
barometry and granite genesis: the Hercynian low-P, high T Velay H2O recycling in granite source regions and migmatite terrains.
anatectic dome (French massif Central). Journal of Metamorphic Lithos 56, 33–45.
Geology 10, 1–15. Trommsdorff, V., 1990. Metamorphism and tectonics in the Central
Naney, M.T., 1983. Phase equilibria of rock-forming ferromagnesian Alps: the Alpine lithospheric melange of Cima Lunga and Adula.
silicates in granitic systems. American Journal of Science 283, Memorie della Società Geologica Italiana 45, 39–49.
993–1033. Viruete, J.E., 1999. Hornblende-bearing leucosome development
Nyman, M.W., Pattison, D.R.M., Ghent, E.D., 1995. Melt extrac- during syn-orogenic crustal extension in the Tormes Gneiss
tion during formation of sillimanite and K-feldspar migmatites, Dome, NW Iberian Massif, Spain. Lithos 46, 751–772.
west of Revelstoke, British Columbia. Journal of Petrology 36, Waters, D.J., Whales, C.J., 1984. Dehydration melting and the granulite
351–372. transition in metapelites from southern Namaqualand, S.-Africa.
Oliver, H.H.S., 1996. Review and classification of structural controls on Contributions to Mineralogy and Petrology 88, 269–275.
fluid flow during regional metamorphism. Journal of Metamorphic Wenk, E., 1970. Zur Regionalmetamorphose und Ultrametamorphose
Geology 14, 477–492. im Lepontin. Fortschritte Mineralogie 47, 34–51.
Pfiffner, O.A., Ellis, S., Beaumont, C., 2000. Collision tectonics in the Wenk, E., 1975. Zur alpinen Metamorphose. Schweizerische Miner-
Swiss Alps: insight from geodynamic modeling. Tectonics 19, alogische und Petrographische Mitteilungen 55, 116–125.
1065–1094. White, R.W., Powell, R., Holland, T.J.B., 2001. Calculation of partial
Prince, C., Harris, N., Vance, D., 2001. Fluid-enhanced melting during melting equilibria in the system Na2O–CaO–K2O–FeO–MgO–
prograde metamorphism. Journal of the Geological Society of Al2O3–SiO2–H2O (NCKFMASH). Journal of Metamorphic
London 158, 233–241. Geology 19, 139–153.
Roselle, G.T., Thüring, M., Engi, M., 2002. MELONPIT: a finite element White, R.W., Powell, R., Clarke Geoffrey, L., 2003. Prograde
code for simulating tectonic mass movement and heat flow within metamorphic assemblage evolution during partial melting of
subduction zones. American Journal of Science 302, 381–409. metasedimentary rocks at low pressures: migmatites from Mt
Sawyer, E.W., 1998. Formation and evolution of granite magmas Stafford, Central Australia. Journal of Petrology 44, 1937–1960.
during crustal reworking: the significance of diatexites. Journal of White, R.W., Pomroy, N.E., Powell, R., 2005. An in situ metatexite
Petrology 39, 1147–1167. diatexite transition in upper amphibolite facies rocks from Broken
Schmid, S.M., Aebli, H.R., Heller, F., Zingg, A., 1989. The role of the Hill, Australia. Journal of Metamorphic Geology 23, 579–602.
Periadriatic line in the tectonic evolution of the Alps. In: Dietrich, Yardley, B.W., Barber, J.P., 1991. Melting reactions in the Connemara
D., Coward, M.P. (Eds.), Alpine Tectonics. Geological Society of Schists: the role of water infiltration in the formation of amphibolite
London Special Publication, London, vol. 45, pp. 153–171. facies migmatites. American Mineralogist 76, 848–856.

You might also like