Anda di halaman 1dari 24

Mineralogy and Petrology (2001) 72: 185207

Hornblende thermobarometry of granitoids


from the Central Odenwald (Germany) and
their implications for the geotectonic
development of the Odenwald
E. Stein1 and C. Dietl2
1

Institut fur Mineralogie, TU Darmstadt, Federal Republic of Germany


Geologisch-Palaontologisches Institut, Universitat Heidelberg, Federal Republic of
Germany

With 6 Figures
Received July 14, 1999;
revised version accepted October 6, 2000
Summary
The three major units of the Bergstrasser Odenwald (Frankenstein Complex,
Flasergranitoid Zone and southern Bergstrasser Odenwald) are, according to literature,
separated by two major shear zones. The aim of the present paper is to evaluate the
importance of these sutures by comparing new hornblende geothermobarometry data
from ve plutons of the Flasergranitoid Zone with published P-T data from the entire
Bergstrasser Odenwald. Furthermore radiometric, geochemical and structural data from
the literature were also used for this purpose. Temperatures were calculated with the
amphibole-plagioclase thermometer and range from 600 to 800  C. Determinations of
the intrusion depth, using the Al-in-hornblende barometer show that most plutons
intruded at pressures ranging from about 4 to 6 kbar (13 to 20 km). These combined data
do not allow to postulate a major suture zone between the Flasergranitoid Zone and the
southern Bergstrasser Odenwald, while comparison of similar data from the Flasergranitoid Zone and the Frankenstein Complex verify the importance of this shear zone.
Moreover, our P-T data show that the high temperature low pressure metamorphism in
the Bergstrasser Odenwald can also be interpreted as contact metamorphism and not
necessarily as regional metamorphism.
Zusammenfassung
Hornblende-Thermobarometrie an Granitoiden des Mittleren Odenwaldes (Deutschland) und ihre Implikation fur die geotektonische Entwicklung des Odenwaldes

186

E. Stein and C. Dietl

Die drei Haupteinheiten des Bergstrasser Odenwaldes (Frankenstein-Komplex, Flasergranitoid-Zone und sudlicher Bergstrasser Odenwald) werden nach der Literatur durch
zwei bedeutende Scherzonen voneinander getrennt. Ziel der vorliegenden Arbeit ist es,
die wirkliche Bedeutung dieser beiden Suturen herauszuarbeiten. Dazu wurden eigene,
neue Hornblende-Geothermobarometrie-Daten, die an funf Plutonen der Flasergranitoidzone ermittelt wurden, mit bereits publizierten P-T-Daten aus dem gesamten
Bergstrasser Odenwald verglichen. Zudem wurden radiometrische, geochemische und
strukturgeologische Datensatze aus der Literatur fur diesen Zweck benutzt. Kristallisationstemperaturen wurden mit Hilfe des Amphibol-Plagioklas-Thermometers errechnet und liegen zwischen 600 und 800  C. Die Bestimmung der Intrusionstiefe mit dem
Al-in-Hornblende-Barometer ergab fur die meisten Plutone Drucke im Bereich von 4
6 kbar (1320 km). Diese, sowie radiometrische, geochemische und strukturgeologische
Daten aus der Flasergranitoid-zone und dem sudlichen Bergstrasser Odenwald geben
keinen Hinweis auf eine wichtige Suturzone zwischen diesen beiden geotektonischen
Einheiten, wohingegen der Vergleich ahnlicher Daten aus der Flasergranitoid-Zone und
dem Frankenstein-Komplex die Bedeutung der Scherzone zwischen diesen beiden
Einheiten hervorhebt. Unsere P-T-Daten zeigen auerdem, da die HochtemperaturNiederdruck-Metamorphose im Bergstrasser Odenwald nicht notwendigerweise eine
Regionalmetamorphose sein mu, sondern ebenso gut als Kontaktmetamorphose
interpretiert werden kann.

Regional setting of the Odenwald


Introduction
The Crystalline Odenwald, part of a magmatic arc along the northern margin of the
Saxothuringian zone, is the largest exposure of crystalline rocks within the so-called
Mid-German Crystalline Rise. To the west it is bound by the Rhine valley, to the
north by the Saar-Selke Trough and to the south and east it is covered by Mesozoic
sediments (compare Fig. 3, Stein, this volume). The Crystalline Odenwald can be
divided geographically and geologically into two parts: the smaller Bollstein Gneiss
dome to the east and the Bergstrasser Odenwald to the west. Both parts have been
interpreted as magmatic arcs, the rst of pre- to early Variscan (Altenberger and
Besch, 1993), the latter of mid- to late Variscan age (Henes-Klaiber, 1992; Kreher,
1994). They are separated by a major Variscan shear zone, the so-called Otzberg
Zone (Hess and Schmidt, 1989). The Bergstrasser Odenwald, which is the subject of
the present paper, consists of ca. 90% calc-alkaline magmatic rocks, and ca. 10%
metasediments forming narrow and distinct, NE-SW trending belts, which separate
the igneous complexes. Willner et al. (1991) distinguished three units in the
Bergstrasser Odenwald: the Frankenstein Complex in the north (unit 1), the central
Flasergranitoid Zone (unit 2) and the southern Bergstrasser Odenwald (unit 3),
which, according to several authors (e.g. Henes-Klaiber, 1992; Krohe, 1994; Altherr
et al., 1999) are also separated by two important shear zones. The aim of this study
was to evaluate the geotectonic importance of the postulated suture zones by
applying hornblende geothermobarometry to rocks from the Flasergranitoid Zone
and comparing these data with published P-T data from the entire Bergstrasser
Odenwald. Moreover, we included published radiometric, geochemical and
structural data for this purpose.

Hornblende thermobarometry of granitoids from the Central Odenwald

187

Fig. 1. Geologic map of the Flasergranitoid Zone with the sampled plutons marked

The magmatic rocks of the Bergstrasser Odenwald


Most of the igneous rocks of the Frankenstein Complex and the adjacent
Bergstrasser Odenwald show I-type signatures with a typical subduction-related
geochemistry, i.e. they are characterized by negative anomalies of Nb, Ta and Ti
(Henes-Klaiber, 1992). In particular, radiometric data point to a different intrusion
and cooling history for the northern and southern parts. 207 Pb/206 Pb-dating on single

188

E. Stein and C. Dietl

zircons (362  9 Ma) as well as 40 Ar/39 Ar-dating on hornblende (363  7 Ma) and
plagioclase (359  3 Ma) of gabbros from the Frankenstein Complex were interpreted as intrusion ages with a very rapid subsequent cooling history (Kirsch et al.,
1986). Low Sr-initials (0.703) indicate a mantle-derived gabbroic magma (Kirsch
et al., 1986). Intermediate to felsic plutonic rocks of the central and southern
Bergstrasser Odenwald have a different intrusion history: the oldest data were
obtained from a granodiorite in the northern part (K/Ar hornblende: 340 Ma; K/Ar
biotite: 330-327 Ma). Similar rocks from the central and southern Bergstrasser
Odenwald provided very homogeneous results which are, however, ca. 510 Ma
younger (K/Ar hornblende: 330-335 Ma; K/Ar biotite: 323-325 Ma; Kreuzer and
Harre, 1975). According to geochemical data, these felsic rocks were derived from
metaluminous crustal protoliths (Altherr et al., 1999).
The metamorphic rocks of the Bergstrasser Odenwald
The magmatic rocks of the study area are seperated by four narrow (less than 1 km
wide) zones of metamorphic rocks, consisting of gneisses, micaschists, graphitebearing quartzites, marbles, calc-silicate rocks and amphibolites. Characteristic
mineral assemblages with sillimanite, andalusite and cordierite in metapelitic rocks
and wollastonite in metacarbonates indicate a high temperature low pressure
amphibolite facies metamorphism (ca. 34 kbar, 600650  C; Okrusch, 1995). This
implies a geothermal gradient of about 50  C/km. Differences in the metamorphic
history between the Frankenstein Complex, characterized by anticlockwise P-T
paths, and clockwise paths in the rest of the Bergstrasser Odenwald were rstly
described by Willner et al. (1991).
235 207
U/ Pb- and 238 U/206 Pb-dating of zircon from metasedimentary rocks of the
central (336-337 Ma) and southern Bergstrasser Odenwald (342 Ma, 332 Ma) were
linked with the thermal peak of the regional metamorphism (Todt et al., 1995). The
subsequent cooling history has been derived from K/Ar - and 40 Ar/39 Ar data of
hornblende (343-335 Ma ; 334 Ma) and biotite (328-317 Ma ; 330 Ma) by Kreuzer
and Harre (1975) and Rittmann (1984).
The geology of the Flasergranitoid Zone
A zone of special interest is the so-called Flasergranitoid Zone in the central
Bergstrasser Odenwald, which is characterized by an intimate association of
gabbros, diorites, granodiorites and granites. In the northern part of the Flasergranitoid Zone predominantly felsic granites alternating with metasediments are
exposed, whereas in the south basic diorites with a few gabbros make up ca. 60% of
the rocks (Stein, 2000). Moreover, biotite diorites are common in the north, while
hornblende diorites are restricted to the Hauptdioritzug in the south. Therefore, it
can be stated that the basicity of the magmatic rocks within the Flasergranitoid Zone
decreases continuously from south to north. This trend is reverse to the general trend
in the Bergstrasser Odenwald (Stein, 2000).
Most of the magmatic rocks show a pronounced planar fabric, the origin of
which is still debated. It is either tectonic, due to syntectonic intrusion in a transtensional regime (Krohe, 1994), or magmatic, due to the successive emplacement of

Hornblende thermobarometry of granitoids from the Central Odenwald

189

different plutons, called nested diapirs (Stein, 2000). Another obvious trend is
documented for the fabric development. To the south, where basic igneous rocks are
intimately associated with intermediate and felsic intrusives as well as metamorphic
rocks, the fabrics are clearly magmatic. Magmatic layering, magmatic foliations and
magmatic lineations are common. These fabrics are only locally overprinted by
solid-state deformations (Stein, 2000). In the northern part, however, magmatic
fabrics within granitic lithologies are strongly obliterated by solid-state fabrics,
these are concentrated in discrete ductile shear zones, which are up to several metres
wide. These shear zones are restricted to a 1.5 km wide area along the border to the
Frankenstein Complex and show all possible transitions from mylonites to ultramylonites. Sinistral and dextral transport directions were observed next to each
other (Stein, 2000). Sinistral strike slip zones are well-known from all over the
Bergstrasser Odenwald and are described in detail from the Melibocus Massive
(Altenberger et al., this volume). The dextral ones are restricted to the north.
Willner et al. (1991) and Krohe (1994) describe an important strike-slip zone
between the Flasergranitoid Zone and the adjacent Weschnitz Pluton to the south.
Therefore, they divided the central and the southern part of the Bergstrasser
Odenwald into two independent tectono-metamorphic units (``unit 2'' and ``unit
3''), although the structural and P-T data, obtained from both parts, do not
signicantly differ. Henes-Klaiber (1992) used this interpretation of the Bergstrasser
Odenwald in her study to propose a continuous increase of the intrusion depth from
the central toward the southern Bergstrasser Odenwald. She suggested that the
plutons, which were intruded into different crustal levels, were juxtaposed along a
major shear zone between the Flasergranitoid Zone and the southern Bergstrasser
Odenwald with a considerable vertical displacement. At rst glance the emplacement mechanisms seem to be very different to the north and south of this shear
zone. In the southern ``unit 3'' the calc-alkaline intrusions occur as large, distinct
diapiric plutons such as the granodiorite of the Weschnitz Pluton and the granites of
Tromm and Heidelberg, whereas in the Flasergranitoid Zone (``unit 2'') most
plutons are small, have an elliptical shape and are intimately associated with each
other (Stein, 2000).
A brief classication of the intrusives of the Flasergranitoid Zone
In the Flasergranitoid Zone of the central Bergstrasser Odenwald Stein (2000)
distinguished at least four different types of intrusions; all of them were sampled:
(1) Round to elliptical plutons with homogeneous and distinctive lithologies. They
are characterized by euhedral K-feldspar or plagioclase phenocrysts, which
together with the matrix minerals and microgranular enclaves are aligned
within a magmatic foliation or banding. Quartz is the only mineral with a
typical solid-state deformation imprint (Dietl and Stein, this volume). Typical
examples are the Melibocus Granodiorite, and the smaller Ludwigshohe Granite.
(2) Round to elliptical plutons with a concentric structure. They show a normal
zoning with diorites at the rims and granites in the cores. The intensity of
magmatic fabrics within these plutons increases from the centre towards the
margin. The most penetrative fabrics occur at lithological boundaries and
decrease continuously toward the centres of the intrusions. This is also true for

190

E. Stein and C. Dietl

solid-state deformations, so that the older diorites at the rim are strongest
overprinted by solid-state fabrics, whereas the younger granites in the core in
most cases do not show any solid-state deformation. Such congurations are
explained by forceful ballooning. One of the best examples in the Flasergranitoid Zone is the Billings Pluton.
(3) Ballooning plutons with a clear reverse zonation, with undeformed diorites in
their cores and granites with intense magmatic fabrics and solid-state overprint
at their margins. This intrusion mechanism characterizes nested diapirs, which
are typical of the southern part of the Flasergranitoid Zone. Good examples are
exposed at the Seemann quarry in Hochstadten or come from the boundary
region to the Weschnitz Pluton, where samples A9 and P53 were taken.
(4) Dykes, of dioritic, granodioritic and granitic composition. Most of them are
ne-grained, but porphyritic dykes with hornblende, plagioclase or K-feldspar
phenocrysts are also described (Nickel and Fettel, 1985). Sample T228IV comes
from a granodioritic porphyry dyke.
Sample description
Sample T228IV comes from a granodioritic porphyry dyke from the northern
Flasergranitoid Zone. It shows abundant micrographic intergrowths of quartz and
K-feldspar. Furthermore, this granodiorite is characterized by several cm-large
plagioclase phenocrysts and myrmekites. The complete assemblage contains quartz,
K-feldspar, plagioclase (An2833), green hornblende, Ti-rich biotite, titanite
ilmenite, apatite and zircon.
The Ludwigshohe Granite is a light, medium to coarse-grained, porphyritic
granite with ca. 2 cm large euhedral K-feldspar phenocrysts. Furthermore it contains schlieren and dark, ne-grained, microdioritic enclaves with dioritic composition. The granite consists of the assemblage plagioclase (An25 ) (38 vol.-%),
quartz (23 vol.-%), K-feldspar (21 vol.-%), biotite (12 vol.-%), hornblende (5 vol.-%)
and the accessories epidote, ilmenite, magnetite, hematite, titanite, zircon and
monazite. In the northern part of the Flasergranitoid Zone Stein (2000) observed
magmatic fabrics, which are almost obliterated by successive solid-state strike-slip
deformation.
The Billings Pluton consists of four lithologies: diorite, granodiorite, granite and
porphyry. The diorite is the only rock type that contains the required mineral
assemblage for geothermobarometry (see below). It is characterized by an intense,
steeply to the NW dipping magmatic foliation with an only weak solid-state
overprint. The diorite is medium- to coarse-grained with large euhedral plagioclase
and hornblende, which are aligned within the foliation plane. The diorite consists
of plagioclase (An2540) (6065 vol.-%), hornblende (15 vol.-%), biotite (10 vol.%), K-feldspar, quartz and chlorite (5 vol.-% each) and zircon, ilmenite, hematite,
calcite and titanite.
Sample A9 comes from a ``Flasergranodiorite'' north of the Hauptdioritzug. It
shows a strong magmatic foliation, which is steeply inclined (318/51). The
granodiorite consists of plagioclase (An2540) (35 vol.-%), quartz (30 vol.-%),
hornblende (15 vol.-%), biotite (10 vol.-%), K-feldspar, and chlorite (5 vol.-% each)
and zircon, ilmenite, hematite, calcite and titanite.

Hornblende thermobarometry of granitoids from the Central Odenwald

191

Sample P53 is a granodiorite from the southernmost edge of the Flasergranitoid


Zone, where it is associated with the metamorphic rocks of the Silbergrubenkopf
area. It consists of plagioclase (An2530) (35 vol.-%), hornblende (25 vol.-%), quartz
(20 vol.-%), biotite (10 vol.-%), K-feldspar, and chlorite (5 vol.-% each) and zircon,
ilmenite, hematite, apatite and titanite.
Amphiboles as thermobarometric ``index'' minerals
All described samples contain mineral assemblages with a certain amount of
amphiboles, which can be used as good pressure and temperature indicators in
igneous rocks. Therefore several plutons of the Flasergranitoid Zone were sampled,
to gain insight into the magmatic history of the Central Odenwald area.
Care was taken to examine only unzoned and unaltered amphiboles of a clear
magmatic origin to get intrusion-related P-T data.
Nomenclature of amphiboles
Amphiboles have been classied according to Leake et al. (1997). Mineral formula
calculations are based on 23 oxygens, standardized on 13 cations (without Ca, Na
and K).
All the investigated
amphiboles plot in the eld of calcic amphiboles, which is
P
dened by
(Ca Na) on M4  1.00 with Na < 0.50, and Ca  1.50 on M4
(Leake et al., 1997). Within the calcic amphiboles Leake et al. (1997) have
distinguished 4 groups:
a Na KA  0:50

and

Ti < 0:50;

b Na KA  0:50 and Ti  0:50;


c Na KA < 0:50 and CaA < 0:50;
d Na KA < 0:50 and CaA  0:50:
The amphiboles of the examined plutons belong either to group ``a)'' or ``c)''
with a distinct relation to individual plutons and lithologies (see also Fig. 2): The
porphyritic Ludwighshohe Granite as well as its microdioritic enclaves for the most
part contain hastingsites accompanied by a small number of ferroedenites, ferropargasites and ferrohornblendes in the granite, and some ferroedenites, magnesiohastingsites, magnesio- and ferrohornblendes in the enclaves. All hornblendes of the
diorites and granodiorites from the Billings quarry are magnesiohornblendes. Most
of the amphiboles from sample A9 are ferrohornblendes, but also some ferrotschermakites and one tschermakite occur. Hornblendes in sample P53 have
magnesiohornblenditic composition, but also edenites, pargasites and magnesiohastingsites occur. In the granodioritic porphyry (sample T228IV) mainly magnesiohornblendes are found, together with edenites and magnesiohastingsites.
Calcic amphiboles are typical for I-type intrusives (Chappel and White, 1974;
Wyborn et al., 1981; White and Chappel, 1983; Clemens and Wall, 1984), supporting
the results of Henes-Klaiber (1992).
In the following the terms ``amphibole'' and ``hornblende'' will be used
synonymously.

192

E. Stein and C. Dietl

Fig. 2. Classication of amphiboles according to the nomenclature of Leake et al. (1997),


LuHo Ludwigshohe

Thermobarometry
Amphiboles are the most useable minerals for geothermobarometry in calc-alkaline
igneous rocks, because they occur in nearly all calc-alkaline intrusives, regardless of
mac, intermediate or felsic compositions. They are stable over a wide P-T range
from 123 kbar and 4001150  C (Blundy and Holland, 1990). Many geothermobarometers are based on the Al-content of hornblende: The Al-in-hornblende
barometer (Hammarstrm and Zen, 1986; Hollister et al., 1987; Johnson and

Hornblende thermobarometry of granitoids from the Central Odenwald

193

Rutherford, 1989; Thomas and Ernst, 1990; Schmidt, 1992; Anderson and Smith,
1995) is controlled by the total Al-content of hornblende. The amphiboleplagioclase thermometer (Blundy and Holland, 1990; Holland and Blundy, 1994) is
based on the number of Si- and Al-cations on the tetraeder positions of amphiboles.
Factors inuencing the Al-content of amphiboles
The intensive parameters pressure, temperature, oxygen fugacity, as well as the
whole rock composition and the coexisting phases determine the Al-content of
hornblende. According to Hollister et al. (1987) the tschermak substitution
Si R2 AlIV AlVI is pressure-sensitive; with increasing pressure the Alcontent in the hornblende lattice increases, too. Other reactions, such as the edenite
substitution Si vacA AlIV K NaA , and reactions involving Ti (e.g.
Ti R2 2AlVI and Ti AlIV AlVI Si) are controlled more by temperature
than by pressure (Anderson and Smith, 1995): The higher the temperature, the more
effective the edenite substitution. This results in an increasing Al-content of
hornblende.
Besides these important substitutions the oxygen fugacity plays a decisive role,
as it controls the Fe # f Fe=Mg Feg and Fe3 =Fe2 Fe3 ratios: The
lower the oxygen fugacity, the more Fe2 is present. Spear (1981) and Anderson
and Smith (1995) classify a Fe # in the range from 0 to 0.6 as high, between 0.6 and
0.8 as medium and up to 1 as low oxygen fugacity. The relationship of substitution
reactions involving Al and of the oxygen fugacity is based on the fact, that a low
oxygen fugacity favors the insertion of Fe2 in the hornblende lattice. A high Fe2 /
Fe3 -ratio preferently favors the substitution of Mg by Al during the tschermak
substitution. A low oxygen fugacity therefore leads to high Al-contents of
hornblende. Therefore Anderson (1997) recommends just to use hornblendes with a
Fe #  0:65 for geobarometry. On the other hand, a high oxygen fugacity leads to a
preferred incorporation of Fe3 into the lattice, which preferably substitutes Al.
This can keep the Al-content of hornblende low. Anderson and Smith (1995)
therefore recommend just to use amphiboles with a Fe3 =Fe2 Fe3 -ratio  0:25
for barometric analyses. The general disadvantage of both these criteria is, that they
are just based on stoichiometric calculations and not on direct measurements of the
Fe3 and Fe2 contents. Therefore Fe # and Fe3 =Fe2 Fe3 ratios cannot
stand as the only criteria, determining the oxygen fugacity.
Possible further objectives may be derived from the presence of accessory
minerals. According to Ishihara (1977) magnetite-bearing igneous rocks (so called
``magnetite series'') point to crystallization conditions under a high oxygen fugacity,
whereas ilmenite-bearing ones (``ilmenite series'') indicate a low oxygen fugacity.
Moreover, the abundance of titanite indicates a high f O2 .
Generally it can be concluded, that hornblende crystallizing under high f O2
give better and more reliable geothermobarometry results than those growing under
low f O2.
Factors inuencing the Al-content of the investigated hornblendes
Both, the Tschermak and the edenite substitution are important for amphiboles of
the investigated plutons indicating that both temperature and pressure have

194

E. Stein and C. Dietl

inuenced the compositions of hornblende of plutons from the Flasergranitoid


Zone.
The role of the oxygen fugacity during the hornblende crystallization is much
more difcult to evaluate. Most of the investigated samples fulll the f O2 -criteria
of Anderson (1997) and Anderson and Smith (1995) Fe #  0:65 and
Fe3 =Fe2 Fe3  0:25 only hornblendes from the Ludwigshohe Pluton
and its enclaves do not. Results from this locality partly show too low Fe3 =
Fe2 Fe3 ratios and a too high Fe #.
On the other hand the porphyritic Ludwigshohe Granite is an example for the
very few magnetite-bearing granitoids of the Flasergranitoid Zone. This composition indicates a high oxygen fugacity and therefore crystallization conditions, which
are suitable for the geobarometric investigations can be assumed. Generally all other
sampled rocks, even some enclaves in the Ludwigshohe Granite, are characterized
by the occurence of ilmenite without magnetite as accessory oxide phase with small
amounts of titanite, i.e. they probably crystallised under low to medium f O2 conditions.
Because all the amphiboles of all plutons fulll at least one criteria for a high
oxygen fugacity, all samples were used for geothermobarometry.
The amphibole-plagioclase thermometer
General comments
Although the amphibole-plagioclase thermometer is still under debate, there is no
other geothermometer that can be applied to calc-alkaline igneous rocks. Furthermore, according to our experience, the resulting temperatures correlate very well
with independently determined temperatures of metamorphic rocks, e.g. using the
garnet-biotite thermometer.
Blundy and Holland (1990) and Holland and Blundy (1994) published three
different calibrations of the amphibole-plagioclase thermometer. Two are based on
the edenite-tremolite reaction:
4 quartz edenite albite tremolite:
One is based on the edenite-richterite reaction:
edenite albite richterite anorthite:
Blundy and Holland (1990) rst proposed a very simple, empirical thermometer
on the basis of the edenite-tremolite reaction, which could be applied only to quartzbearing, intermediate to felsic igneous rocks with plagioclase An  0:92 and Si in
hornblende  7:8 atoms p.f.u. This thermometer is calibrated for temperatures
between 500  C and 1100  C. It already takes into account, that the Al-content of
hornblende does not only depend on temperature, but also on pressure. The
thermometer is described by the following formula:
0:677Pkbar 48:98



T311 K
Si 4 Plag
X
0:0429 0:0083144 ln
8 Si Ab
with Si atoms p. f. u. in hornblende, and XAb in plagioclase in decimal units.

Hornblende thermobarometry of granitoids from the Central Odenwald

195

The formula presented in this paper has already been changed slightly from the
original version, as it has been adopted to the plagioclase compositions observed in
the investigated igneous rocks, which range from albite to andesine.
As this calibration resulted in too high temperatures for some lithologies (e.g.
Poli and Schmidt, 1992), Holland and Blundy (1994) recalibrated the amphiboleplagioclase thermometer. They extended the data base to all components, which
take part in the edenite-tremolite reaction. Moreover, they considered non-ideality
instead of ideality. These changes enable an application of this thermometer A
(Holland and Blundy, 1994) to quartz-bearing metabasites. It now is:
T313 K

A
M2
76:95 0:79Pkbar 39:4XNa
22:4XKA 41:5 2:89PkbarXAl
!
A
T1 Plag
27Xvac
XSi
XAb
0:0650 0:0083144 ln
A
T1
256XNa XAl

Additionally they calibrated a second version, thermometer B, which is based on


the edenite-richterite reaction (Holland and Blundy, 1994), and which is applicable
also to quartz-free igneous rocks:
T313 K

M4
M2
T1
A
81:44 33:6XNa
:66:88 2:92PkbarXAl
78:5XAl
9:4XNa
!
M4 T1 Plag
27XNa
XSi XAn
0:0721 0:0083144 ln
M4
T1 X Plag
64XCa XAl
Ab

Although these two thermometers can be used for a wide range of lithologies,
they have one main disadvantage; they take too many components into account,
which all inuence the calculated temperature, and which therefore can all act as
sources of error.
Again the presented formulae of thermometers A and B have been changed
slightly. They are also adopted to the An-contents of plagioclase in the investigated
samples.
Application of the amphibole-plagioclase thermometer to the Ludwigshohe
and Billings Plutons
One imperative prerequisite for the application of all three thermometers is the
availability of independently determined pressure data. Pressures determined by
Willner et al. (1991) from outcrops close to the sampled plutons were used for this
purpose (Table 1). P-T data of these authors from metamorphic wall rocks of the
Ludwigshohe and Billings Plutons as well as the Silbergrubenkopf area (sample
P53) correlate with the intrusion of the plutons. Pressures from the Felsberg and the
Helgengrund localities, close to the Ludwigshohe Pluton range from 4.4 kbar to
4.6 kbar; data from the Rimdidim outcrop, close to the Billings Pluton, give ca.
2.7 kbar. Unfortunately no hornblende-plagioclase pairs of sample P53 were
investigated, although at least one pressure value of the neighbouring Silbergrubenkopf (4.9 kbar) lies on the prograde branch of the P-T path and seems to
correlate with the intrusion of the granitoids (Willner et al., 1991). P-T data from
the Gadernheim locality, close to sample A9, represent only the retrograde

196

E. Stein and C. Dietl

Table 1. Comparison of P-T data determined by Willner et al. (1991) and P-T data from this
study
Willner et al. (1991)

this paper

outcrop

P [kbars] T [ C]

Helgengrund
Felsberg
Rimdidim
Muhlberg quarry
Kolmbach
Gadernheim
Silbergrubenkopf
Oberhambach

4.44.6
3.84.7
2.7
3.0

2.54.3
2.04.9
2.7

560660
550
625

700
600610
600625
625

close to

outcrop

P [kbars] T [ C]

T 228 IV
Ludwigshohe

2.94.1
4.16.2

626787

Billings

1.93.0

605697

A9

4.56.0

P 53

4.35.7

development of originally medium pressure rocks. As no independent pressure


data of metamorphic rocks from the northern Flasergranitoid Zone exist, the
amphibole-plagioclase thermometer was not applied to sample T228IV. Therefore,
the three different calibrations of the amphibole-plagioclase thermometer were
applied only to the porphyritic Ludwigshohe Granite, its enclaves and the Billings
Pluton.
Temperature data were calculated for individual hornblende-plagioclase pairs of
the different samples (4 from the porphyritic Ludwigshohe Granite, 16 from its
enclaves and 2 from the Billings Pluton), from which average temperatures for each
locality were derived. The resulting temperature range for both investigated plutons
is presented in Table 1. Average plagioclase and hornblende compositions are listed
in Table 2. Typical structural relationships between hornblende and plagioclase are
shown in Fig. 3.
As obvious in Fig. 4, temperatures calculated with the 1990 thermometer
(Blundy and Holland, 1990) are signicantly higher than those computed with both
1994 calibrations. Hornblende and cogenetic plagioclase of the porphyritic
Ludwigshohe Granite crystallized at temperatures of about 768  27  C, the
enclaves at 787  37  C and the Billings Pluton at 697  2  C. All temperatures lie
above the wet granitic solidus.
Using thermometer A (Holland and Blundy, 1994) signicantly lower temperatures were determined: for the porphyritic Ludwigshohe Granite T 704  47  C,
for the Ludwigshohe enclaves T 743  24  C and for the Billings Pluton
T 643  10  C. Even lower temperatures were calculated, using thermometer B
(Holland and Blundy, 1994): for the Ludwighshohe Granite T 626  65  C, for its
microdioritic enclave T 660  29  C and for the Billings Pluton T 605 
19  C. Data calculated with thermometer B (Holland and Blundy, 1994) are below
the granitic solidus. These appear not to be reliable, because structural relationships
between hornblende and plagioclase in the Ludwigshohe Granite and its enclaves
indicate a magmatic origin of both minerals (Dietl and Stein, this volume).

Hornblende thermobarometry of granitoids from the Central Odenwald

197

Table 2. Average hornblende and plagioclase compositions of the investigated igneous rocks (LH, g
Ludwigshohe, granite; LH, e Ludwigshohe, enclave)

The Al-in-hornblende barometer


General comments
Hammarstrm and Zen (1986) were the rst to suggest a relationship between the
Altot -content of amphiboles and the conning pressure, under which amphiboles
crystallized. Based on microprobe measurements of amphiboles from granitoids, for

198

E. Stein and C. Dietl

Fig. 3. BSE photographs of typical structural and petrographic relationships among


a hornblende, K-feldspar and quartz in the porphyritic Ludwigshohe Granite and
b hornblende, biotite, plagioclase, and quartz in its microdioritic enclaves

the intrusion depth of which have been calculated independently at 2 kbar and
8 kbar respectively, they formulated a rst empirical geobarometer:
P3 kbars 3:92 5:03Altot
Hollister et al. (1987) conrmed this correlation and empirically extended the
barometer to granitoids, which crystallized at pressures between 4 and 6 kbar. At

Hornblende thermobarometry of granitoids from the Central Odenwald

199

Fig. 4. Comparison of the crystallization temperatures of hornblende-plagioclase pairs


from the Ludwigshohe and Billings Plutons, respectively. They were determined with the
three calibrations of the amphibole-plagioclase thermomter: a Holland and Blundy (1994)
thermometer B on the x-axis versus Holland and Blundy (1994) thermometer A on the yaxis; b Holland and Blundy (1994) thermometer B on the x-axis versus Blundy and Holland
(1990) on the y-axis. It is clear from these two graphs, that the Blundy and Holland (1990)
calibration provides the highest temperatures and thermometer B of Holland and Blundy
(1994) the lowest, in some cases even unrealistic low values (below the wet granitic
solidus)

the same time they reduced the error bar of the barometer with their recalibrated
formula:
P1 kbar 4:76 5:64Altot
A rst experimental calibration of this barometer was carried out by Johnson and
Rutherford (1989) at temperatures between 720  C and 780  C, taking a CO2 H2 O
mixture with two different compositions (CO2 :H2 O 50:50 and 75:25) as uid
phase, to reach pressures between 2 and 8 kbar. Their formula reads as follows:
P0:5 kbar 3:46 4:23Altot

200

E. Stein and C. Dietl

Fig. 5. a The four most important


temperature-independent calibrations of the Al-in-hornblendebarometer: H & Z '86: Hammarstrm and Zen (1986); H et al.
'87: Hollister et al. (1987); J & R
'89: Johnson and Rutherford
(1989); S '92: Schmidt (1992).
b The calibration of the Al-inhornblende-barometer by Anderson and Smith (1995) applied to
different temperatures

Thomas and Ernst (1990) carried out further experiments, using a pure H2 O
uid at 750  C and a pressure range of 6 to 12 kbar. They achieved similar results as
Johnson and Rutherford, at least for the pressure range between 6 and 8 kbar.
Schmidt (1992) calibrated his experimental barometer at temperatures between
655  C and 700  C under water saturated conditions in the pressure range from 2.5
to 13 kbar. His Al-in-hornblende barometer reads:
P0:6 kbar 3:01 4:76Altot
All these four calibrations provided very similar pressures (Fig. 5).
According to the cited authors several prerequisites have to be fullled strictly
for a correct application of the barometers:
(1) the assemblage quartz, plagioclase, K-feldspar, hornblende, biotite, titanite and
magnetite/ilmenite has to be present contemperaneously with melt,

Hornblende thermobarometry of granitoids from the Central Odenwald

201

(2) the barometer can be applied only to rocks, which crystallized in a pressure
range between 2 and 13 kbar,
(3) plagioclase coexisting with hornblende should range between An25 and An35 ,
(4) hornblende should have crystallized near the granitic solidus,
(5) the Si-activity of the melt must have been  1, i.e. it must have been SiO2 saturated, because the Al-content of hornblende is directly related to its Sicontent and therefore also to the Si-activity of the entire system,
(6) amphibole should coexist with K-feldspar, because its activity also inuences
the Al-content of hornblende,
(7) due to the last three prerequisites only rims of hornblende in contact with quartz
and/or K-feldspar should be measured.
Taking all these preconditions into account, the Al-content of hornblende should
only be controlled by the pressure dependent Tschermak substitution and therefore
it can be used as a good barometer.
Already Blundy and Holland (1990) emphasized, that temperature plays a more
important role for the Al-content of amphiboles as the above cited authors
conceeded. Anderson and Smith (1995) therefore presented a new formulation of the
Al-in-hornblende barometer, which considers all the three intensive parameters
pressure, temperature and oxygen fugacity which control the Al content of
hornblendes.
The recalibration of Anderson and Smith (1995) is based on the Al-inhornblende barometers of Johnson and Rutherford (1989) and Schmidt (1992). They
introduce a temperature correction term on the basis of the amphibole-plagioclase
thermometer of Blundy and Holland (1990). The introduction of the temperature
sensitive edenite substitution to the barometer enables pressure calculations even for
igneous amphiboles, which did not crystallize at or close to the granitoids solidus.
Oxygen fugacity is a new limiting factor in the Al-in-hornblende barometer of
Anderson and Smith (1995), as they restricted its application to amphiboles, which
crystallized at high f O2 . The authors take the Fe # and the Fe3 =Fe3 Fe2
ratio as a measure for f O2. They recommend to use only hornblende with a Fe
#  0:65 and a Fe3 =Fe3 Fe2 ratio  0:25 for barometric purposes, because
all experimental calibrations of their Al-in-hornblende barometer were carried out
under medium to high oxygen fugacities. The new formula of Anderson and Smith
(1995) reads as follows:
 

T C 675
P0:6 kbar 3:01 4:76Altot
85
 f0:53Altot 0:005294  T C 675g
Application of the Al-in-hornblende barometer to the investigated plutons
All ve introduced calibrations of the Al-in-hornblende barometer (Hammarstrm
and Zen, 1986; Hollister et al., 1987; Johnson and Rutherford, 1989; Schmidt, 1992;
Anderson and Smith, 1995) were applied to the sampled plutons of the Flasergranitoid
Zone (Fig. 1). The calibration of Anderson and Smith (1995) can only be applied to
the Ludwigshohe and Billings Plutons, because only for these two granitoids

202

E. Stein and C. Dietl

Fig. 6. All ve calbrations of the Al-in-hornblende barometer applied to the six sample
fractions: a) Hammarstrm and Zen (1986); b) Hollister et al. (1987); c) Johnson and
Rutherford (1989); d) Schmidt (1992); e) Anderson and Smith (1995)

plagioclase analyses are available, which are necessary for the temperaturecorrection term based on the amphibole-plagiocase thermometer of Blundy and
Holland (1990). The temperature correction has been carried out, using average temperatures for each individual pluton (Billings; Ludwigshohe Granite and enclaves).
For the investigated plutons the following pressures were determined (for a
graphical compilation of all data see also Fig. 6, for an overview of the pressure
ranges see Table 1):

Hornblende thermobarometry of granitoids from the Central Odenwald

203

Sample T228IV: 9 amphibole rim measurements from the granodioritic porphyry of


sample T228IV (for an average analysis see Table 2) provided an average Altot of
1.501  0.106, equivalent to pressures of 3.6  0.5 kbar (applying the calibration of
Hammarstrm and Zen, 1986), 3.7  0.6 kbar (Hollister et al., 1987), 2.9  0.4 kbar
(Johnson and Rutherford, 1989) and 4.1  0.5 kbar (Schmidt, 1992). Because no
plagioclase compositions were measured for the granodioritic porphyry, pressures
were not temperature-corrected.
Ludwigshohe Pluton: For pressure determination of the Ludwigshohe Pluton in total
17 hornblende measurements could be used, 13 from the porphyritic granite itself
and 4 from enclaves and schlieren. The microprobe analyses were obtained from
amphibole rims, which are in contact either with quartz or with K-feldspar. The
structural relationship of hornblende and quartz/K-feldspar is displayed in Fig. 3a.
Average Altot -contents of hornblendes are 1.931  0.222 for the granite and
1.936  0.050 for the enclaves. Consequently, pressures without applying the
temperature correction term calculated for the Ludwigshohe Pluton range from 4.7
to 6.2 kbar. Both sample fractions, the porphyritic granite and the enclaves gave the
same average pressure for each calibration. The lowest pressure (4.7  0.9 kbar) is
calculated with the calibration of Johnson and Rutherford (1989), the highest
(6.2  1.1 kbar) with the calibration of Schmidt (1992). The other two give intermediate values of 5.8  1.1 kbar (Hammarstrm and Zen, 1986) and 6.1  1.3 kbar
(Hollister et al., 1987). Since the amphibole-plagioclase thermometer of Blundy and
Holland (1990) yielded temperatures, which are signicantly above the solidus
(granite: 768  C, enclaves: 787  C), a temperature correction, according to Anderson
and Smith (1995), seems to be reasonable. This correction generally leads to lower
pressure values. Values of 4.5  0.9 kbar for the granite and 4.1  0.2 kbar for the
enclaves were determined, which correlate with those computed with the calibration
of Johnson and Rutherford (1989). This result, obtained with the Anderson and
Smith calibration seems to t very well, because Johnson and Rutherford (1989)
calibrated their experiments at temperatures between 720 and 780  C, and this is
exactly the temperature range, in which hornblendes of the Ludwigshohe Pluton
crystallized.
Billings Pluton: In the sample from the Billings Pluton 7 measurements of
hornblende rims in contact with quartz were carried out. The average Altot -content
is 1.259  0.111. Derived pressures range from 1.9  0.5 kbar (Johnson and
Rutherford, 1989) to 3.0  0.5 kbar (Schmidt, 1992). Temperature data of ca.
697  C indicate that the amphiboles crystallized well above the solidus. Therefore,
the calibration of Anderson and Smith (1995) with its temperature correction term
was applied. According to this calibration the Billings Pluton was intruded under a
conning pressure of 2.8  0.5 kbar.
Samples A9 and P53: For samples A9 and P53 no plagioclase microprobe analyses
were carried out. Therefore, only the 4 calibrations without a temperature correction
term were applied to these samples. In total 7 hornblende measurements of sample
A9 and 8 of P53 fulll the requirements of the Al-in-hornblende barometry (for an
average analysis see Table 2). Both provided similar average Altot -values with

204

E. Stein and C. Dietl

1.889  0.151 for A9 and 1.839  0.042 for P53, resulting in similar pressure
data. The pressures range from 5.6  0.8 to 6.0  0.7 kbar for sample A9 and
from 5.3  0.2 to 5.7  0.2 kbar for sample P53, using those barometers calibrated
at lower temperatures. Only the Johnson and Rutherford (1989) barometer
gives lower pressures at 4.5  0.6 kbar for sample A9 and 4.3  0.2 kbar for sample
P53.
Implication of the thermobarometric results for the importance
of two major shear zones in the Bergstrasser Odenwald
Determinations of the intrusion depth of several plutons in the Flasergranitoid Zone
using the Al-in-hornblende geobarometer show that the analyzed plutons intruded at
pressures between ca. 2.5 and 6 kbar, what correlates with intrusion depths of 8 to
19.5 km. Neither a regional distribution pattern within the Flasergranitoid Zone (e.g.
an increasing intrusion depth from north to south), nor any correlations with the
different intrusion types and mechanisms can be derived from these data.
Samples from the southernmost edge of the Flasergranitoid Zone (sample P53)
and the Weschnitz Pluton (samples WP20 and 26; Henes-Klaiber, 1992) at the
northern edge of the southern Bergstrasser Odenwald provided very similar pressure
data, between 5.0 and 5.7 kbar. Such similar data do not allow to suggest that
vertical displacement between the Flasergranitoid Zone and the southern
Bergstrasser Odenwald took place as proposed by Henes-Klaiber (1992): This
interpretation is supported by several other arguments (Stein, 2000):
(1) Radiometric data of Kreuzer and Harre (1975), Rittmann (1984) and Todt et al.
(1995) do not indicate either a hiatus between the intrusions of the central and
southern Bergstrasser Odenwald, respectively, or between the metamorphic
imprint in both these ``units''.
(2) No major difference in the geochemical signatures of plutons from the central
and southern Bergstrasser Odenwald was reported (Altherr et al., 1999).
(3) Magmatic fabrics in the southernmost Flasergranitoid Zone do not show any
pervasive overprint by solid state deformation as it is expected in a major shear
zone.
(4) The different sizes of the plutons in the southern Bergstrasser Odenwald and
their homogeneity compared to the Flasergranitoid Zone cannot be used as
strong argument for a large vertical displacement at the boundary, because also
in the Flasergranitoid Zone large homogeneous plutons, e.g. the Melibocus
Granodiorite, occur.
On the other hand our P-T data, in agreement with those of Willner et al. (1991),
clearly establish different intrusion depths for the Frankenstein Gabbro and for the
plutons of the Flasergranitoid Zone. Furthermore radiometric, geochemical and
structural data again point to important differences between the Frankenstein
Complex and the Flasergranitoid Zone:
(1) The Frankenstein gabbro intruded ca. 360 Ma ago (Kirsch et al., 1986), that is
about 20 Ma before pluton emplacement in the Flasergranitoid Zone started
(Kreuzer and Harre, 1975).

Hornblende thermobarometry of granitoids from the Central Odenwald

205

(2) The Frankenstein Gabbro is derived of mantel melts (Kirsch et al., 1986), while
all plutons in the southern part of the Bergstrasser Odenwald have a crustal
signature (Altherr et al., 1999).
(3) Solid state fabrics transpose earlier magmatic fabrics within the felsic granites
of the northern Flasergranitoid Zone (Stein, 2000).
All these facts support the model of a major tectonic boundary, probably
developed as strike-slip shear zone with a strong vertical component between these
two units.
Moreover, the results of this study indicate that care must be taken with the
interpretation of the pervasive high temperaturelow pressure metamorphism in the
Bergstrasser Odenwald. This metamorphic event is generally regarded as regional
metamorphism and not as contact metamorphism (Taborszky et al., 1975 and
references therein). However, our P-T data for igneous rocks of the Flasergranitoid
Zone are very close to the P-T data of Willner et al. (1991) for metamorphic rocks of
this zone (Table 1). Considering that 90% of the entire Bergstrasser Odenwald
consist of plutonic rocks and only 10% are made up of metamorphic country rocks,
it appears possible to interpret the high temperaturelow pressure metamorphism as
dynamic contact-metamorphism due to the widespread intrusions in the Bergstrasser
Odenwald.
Regarding the entire Crystalline Odenwald, we favour the interpretation of Stein
(2000), who separated the Odenwald into three main geotectonic units: The
Bollstein Odenwald, the Frankenstein Complex and the Bergstrasser Odenwald. All
these different units may have had a common sedimentation and early tectonometamorphic development related to a medium pressure metamorphic event at
about 400-375 Ma. Afterwards these units had been separated and experienced
different tectono-magmatic and tectono-metamorphic histories. Finally they were
juxtaposed again along two major strike-slip zones: one between the Frankenstein
Complex and the Flasergranitoid Zone, and the second, the so-called Otzberg Zone,
between the Bollstein and the Bergstrasser Odenwald.
Acknowledgements
This paper forms part of the habilitation thesis of E. Stein. This research was nanced by
Deutsche Forschungsgemeinschaft (DFG) grants STE 678-1 and STE 678-2. Microprobe
analyses were carried out at the Material Sciences Department of the TU Darmstadt and
guided by Dr. S. Weinbruch and S. Riedel. We thank Prof. Dr. P. Blumel, Dr. J. Reinhardt
and Dr. D. Scheuvens for constructive discussions. We owe special thanks to Dr. D. Tanner
for improving the English of the manuscript, and to Prof. Dr. W. Schubert for his very
constructive review.
References
Altenberger U, Besch T (1993) The Boellstein Odenwald; evidence for pre- to early Variscan
plate convergence in the Central European Variscides. Geol Rdsch 82: 475488
Altenberger U, Oberhansli R, Stein E, Moghni M (2001) Geochemistry, tectonic setting and
geodynamic signicance of late orogenic dikes in the Melibocus Massif, Bergstrasser
Odenwald. Mineral Petrol (this volume)

206

E. Stein and C. Dietl

Altherr R, Henes-Klaiber U, Hegner E, Satir M, Langer C (1999) Plutonism in the


Variscan Odenwald (Germany): from subduction to collision. Int J Earth Sci 88: 422
443
Anderson JL (1997) Status of thermobarometry in granitic batholiths. Trans Roy Soc Edinb
Earth Sci 87: 125138
Anderson JL, Smith DR (1995) The effects of temperature and f O2 on the Al-in-hornblende
barometer. Am Mineral 80: 549559
Blundy JD, Holland TJB (1990) Calcic amphibole equilibria and a new amphiboleplagioclase geothermometer. Contrib Mineral Petrol 104: 208224
Chappel BW, White AJR (1974) Two contrasting granite types. Pacic Geol 8: 173174
Clemens JD, Wall VJ (1984) Origin and evolution of a peraluminous silicic ignimbrite
suite: the Violet Town Volcanics. Contrib Mineral Petrol 88: 354371
Dietl C, Stein E (2001) The diapiric emplacement and related magmatic fabrics of the
porphyritic Ludwigshohe granite, Central Odenwald (Germany). Mineral Petrol (this
volume)
Hammarstrom JM, Zen E-an (1986) Aluminium in hornblende: an empirical igneous
geobarometer. Am Mineral 71: 12971313
Henes-Klaiber U (1992) Zur Geochemie der variszischen Granitoide des Bergstrasser
Odenwaldes. Thesis, TU Karlsruhe, 264 pp (unpublished)
Hess JC, Schmidt G (1989) Zur Altersstellung der Kataklastite im Bereich der OtzbergZone, Odenwald. Geol Jb Hessen 117: 6977
Holland T, Blundy J (1994) Non-ideal interactions in calcic amphiboles and their bearing
on amphibole-plagioclase thermometry. Contrib Mineral Petrol 116: 433447
Hollister LS, Grissom GC, Peters EK, Stowell HH, Sisson VB (1987) Conrmation of the
empirical correlation of Al in hornblende with pressure of solidication of calc-alkaline
plutons. Am Mineral 72: 231239
Ishihara S (1977) The magnetite-series and ilmenite-series. Min Geol 27: 293305
Johnson MC, Rutherford MJ (1989) Experimental calibration of the aluminium-inhornblende geobarometer with applications to Long Valley caldera (California) volcanic
rocks. Geology 17: 837841
Kirsch H, Kober B, Lippolt HJ (1988) Age of intrusion and rapid cooling of the
Frankenstein gabbro (Odenwald, SW-Germany) evidenced by 40 Ar/39 Ar and singlezircon 207 Pb/206 Pb measurements. Geol Rdsch 77/3: 693711
Kreher B (1994) Petrologie und Geochemie der Gabbrointrusionen des Frankensteins
(Odenwald). Geol Jb Hessen 122: 81122
Kreuzer H, Harre W (1975) K/Ar-Altersbestimmungen an Hornblenden und Biotiten des
Kristallinen Odenwalds. In: Amstutz GC, Meisl S, Nickel E (eds) Mineralien und
Gesteine im Odenwald. Aufschlu Sbd 27: 7177
Krohe A (1994) Verformungsgeschichte in der mittleren Kruste eines magmatischen
Bogens. Geotekt Forsch 80: 1147
Leake BE, Woolley AR, Arps CES, Birch WD, Gilbert MC, Grice JD, Hawthorne FC, Kato
A, Kisch HJ, Krivovichev VG, Linthout K, Laird J, Mandarino J, Maresch WV, Nickel
EH, Rock NMS, Schumacher JC, Smith DC, Stephenson NCN, Ungaretti L, Whittaker
EJW, Youzhi G (1997) Nomenclature of amphiboles. Report of the Subcommittee on
Amphiboles of the International Mineralogical Association Commission on New
Minerals and Mineral Names. Eur J Mineral 9: 623651
Nickel E, Fettel M (1985) Odenwald. Vorderer Odenwald zwischen Darmstadt und
Heidelberg. Sammlung geol Fuhrer 65: 1231
Okrusch M (1995) Metamorphic evolution. In: Dallmeyer RD, Franke W, Weber K (eds)
Pre-Permian geology of Central and Eastern Europe. Springer, Berlin Heidelberg New
York Tokyo, pp 201213

Hornblende thermobarometry of granitoids from the Central Odenwald

207

Poli S, Schmidt MW (1992) A comment on ``Calcic amphibole equilibria and a new


amphibole-plagioclase geothermometer'' by Blundy JD and Holland TJB (Contrib
Mineral Petrol 1990, vol 104, pp 208224). Contrib Mineral Petrol 111: 273282
Rittmann KL (1984) Argon in Hornblende, Biotit und Muskovit bei der geologischen
Abkuhlung-40 Ar/39 Ar-Untersuchungen. Thesis, University of Heidelberg, 278 pp
(unpublished)
Schmidt MW (1992) Amphibole composition in tonalite as a function of pressure: an
experimental calibration of the Al-in-hornblende barometer. Contrib Mineral Petrol
110: 304310
Spear FS (1981) Amphibole-plagioclase equilibria: an empirical model for the reaction
albite tremolite edenite 4 quartz. Contrib Mineral Petrol 77: 355364
Stein E (2000) Zur Platznahme von Granitoiden Vergleichende Fallstudien zu Gefugen
und Platznahmemechanismen aus den White-Inyo Mountains, California, USA und
dem Bergstrasser Odenwald. Geotekt Forsch 93: 1330
Taborszky FK, Taupitz D, Gehlen K von (1975) Der Auerbacher Marmor. In: Amstutz GC,
Meisl S, Nickel E (eds) Mineralien und Gesteine im Odenwald. Aufschlu Sbd 27: 149
157
Thomas WM, Ernst WG (1990) The aluminium content of hornblende in calc-alkaline
granitic rocks: a mineralogic barometer calibrated experimentally to 12 kbar. In:
Spencer RJ, Chou IM (eds) Fluid-mineral interactions: a tribute to HP Eugster.
Geochem Soc Spec Publ 2: 5963
Todt WA, Altenberger U, Raumer JF von (1995) U-Pb data on zircons for the thermal peak
of metamorphism in the Variscan Odenwald, Germany. Geol Rdsch 84: 466472
White AJR, Chappel BW (1983) Granitoid types and their distribution in the Lachlan Fold
Belt, southeastern Australia. Mem Geol Soc Am 159: 2134
Willner AP, Massonne HJ, Krohe A (1991) Tectono-thermal evolution of a part of a
Variscan magmatic arc: the Odenwald in the Mid-German Crystalline Rise. Geol Rdsch
80: 369389
Wyborn D, Chappel BW, Johnston RM (1981) Three S-type volcanic suites from the
Lachlan Fold Belt, southeast Australia. J Geophys Res 86: 1033510348
Authors' addresses: Dr. E. Stein, Institut fur Mineralogie, TU Darmstadt, Schnittspahnstrasse 9, D-64287 Darmstadt, Federal Republic of Germany; e-mail: eckardt.stein@
t-online.de; C. Dietl, Geologisch-Palaontologisches Institut, Universitat Heidelberg, Im
Neuenheimer Feld 234, D-69120 Heidelberg, Federal Republic of Germany

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Anda mungkin juga menyukai