Eckelmann et al. 2014

Transcription

Eckelmann et al. 2014
Gondwana Research 25 (2014) 1484–1500
Contents lists available at ScienceDirect
Gondwana Research
journal homepage: www.elsevier.com/locate/gr
Plate interactions of Laurussia and Gondwana during the formation of
Pangaea — Constraints from U–Pb LA–SF–ICP–MS detrital zircon ages
of Devonian and Early Carboniferous siliciclastics of the
Rhenohercynian zone, Central European Variscides☆
Katja Eckelmann a,⁎, Heinz-Dieter Nesbor b, Peter Königshof c, Ulf Linnemann a, Mandy Hofmann a,
Jan-Michael Lange a, Anja Sagawe a
a
b
c
Senckenberg Naturhistorische Sammlungen Dresden, Museum für Mineralogie und Geologie, GeoPlasma Lab, Königsbrücker Landstr. 159, D-01109 Dresden, Germany
Hessisches Landesamt für Umwelt und Geologie, Abteilung Geologie und Boden, Geologischer Landesdienst, Rheingaustraße 186, 65203 Wiesbaden, Germany
Senckenberg Naturmuseen und Forschungsinstitute, Senckenberganlage 25, 60325 Frankfurt am Main, Germany
a r t i c l e
i n f o
Article history:
Received 24 January 2013
Received in revised form 1 May 2013
Accepted 21 May 2013
Available online 20 June 2013
Handling Editor: R.D. Nance
Keywords:
Gondwana provenance
Laurussia provenance
Variscan nappe tectonics
Rheic Ocean
Rhenohercynian Ocean
a b s t r a c t
The southern Rheinisches Schiefergebirge, which is part of the Rhenohercynian zone of the Central European
Variscides, exhibits several allochthonous units: the Gießen-, and the Hörre nappe, and parts of the
Frankenbach imbrication zone. These units were thrust over autochthonous and par-autochthonous
volcano-sedimentary complexes of the Lahn and Dill–Eder synclines. This paper reports a representative
data set of U–Pb LA–SF–ICP–MS ages of 1067 detrital zircon grains from Devonian and Lower Carboniferous
siliciclastic sediments of the autochthonous and the allochthonous areas, respectively. The cluster of U–Pb
ages from the allochthonous units points to a provenance in the Saxothuringian zone. Zircon populations
from the Saxothuringian zone are representative of a Gondwanan hinterland and are characterized by age
clusters of ~ 530–700 Ma, ~1.8–2.2 Ga, ~2.5–2.7 Ga, and ~ 3.0–3.4 Ga. Further samples were taken from the
autochthonous and par-autochthonous units of the Lahn–Dill and Kellerwald areas. A Lower Devonian sandstone sample from the Siegen anticline provides a reference for siliciclastic sediments derived from the Old
Red Continent. These samples show a provenance representative of Laurussia with debris primarily derived
from Baltica and Avalonia. U–Pb zircon age clusters occur at ~400–450 Ma, 540–650 Ma, 1.0–1.2 Ga, ~1.4–1.5 Ga,
~1.7–2.2 Ga, and 2.3–2.9 Ga. Provenance analysis and geochemical data of the Rhenohercynian zone provide new
information on the evolution of magmatic arcs in the Mid-Paleozoic. The data set constrains top-SE and top-NW
directed subduction of the oceanic crust of the Rheic Ocean. Subduction-related volcanism lasted from the Early
Devonian to the Early Carboniferous and thus confirms the existence of the Rheic Ocean until the Early
Carboniferous. The tectonic model outlined for the Rhenohercynian zone suggests a wide Rheic Ocean.
© 2013 International Association for Gondwana Research. Published by Elsevier B.V. All rights reserved.
1. Introduction
The Central-European part of the Variscan Orogen is one of the most
significant type areas for the continent–continent collision between
Laurussia and Gondwana that resulted in the formation of the supercontinent Pangaea. Traditionally, the Central-European Variscides
have been subdivided across strike from NW to SE into three zones,
the Rhenohercynian, the Saxothuringian, and the Moldanubian zones,
respectively (Kossmat, 1927). Later, Brinkmann (1948) included a separate Mid-German Crystalline zone to this model, which is interpreted
as a structural high between the Rhenohercynian and Saxothuringian
☆ This paper is dedicated to Dr. Peter Bender, Marburg, Germany.
⁎ Corresponding author. Tel.: +49 351 7958414424.
E-mail address: [email protected] (K. Eckelmann).
zones (Fig. 1). This concept has been adapted to modern plate tectonic
views in which the Variscan Orogen resulted from the accretion of
peri-Gondwanan terranes or microplates by the closure of several
oceans, among which the Rheic Ocean has been considered the most
important.
The Rhenohercynian zone, which crops out preferentially in the
Rheinisches Schiefergebirge, constitutes a classical fold-and-thrust belt
and is one of the structural units of the European Variscides interpreted
as a collage of microplates (e.g. Matte, 1986; Franke and Oncken, 1990;
Franke et al., 1990; Franke and Oncken, 1995; Franke, 2000).
It is now widely accepted that the Rheinisches Schiefergebirge is
part of the Avalonia terrane, which separated from Gondwana in the
Early Ordovician and drifted northwards (e.g. Oncken et al., 2000; Tait
et al., 2000; Torsvik and Cocks, 2004; Linnemann et al., 2008, 2010;
Romer and Hahne, 2010). This resulted in the opening of the Rheic
Ocean. In Late Ordovician, around 450 Ma, Avalonia collided with
1342-937X/$ – see front matter © 2013 International Association for Gondwana Research. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.gr.2013.05.018
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
8°
10°
12°
s
Gommern
Harz
mountains
autochthonous and par-autochthonous units
52°
Kammquartzite imbrication zone
allochthonous units
La
1485
u
ur
a
si
ex
Rheinisches
Schiefergebirge
51°
an
ni
y
rc
he e
o
en zon Lahn-Dill
area
Rh
pl
Kassel
pp
rz
Ha
i
G
or
N
Lahn and DillEder syncline
Frankfurt
th
er
n
y
Ph
llit
n
zo
S
e
e
an n
m zo
r
e
e n
-G lli
id ta
M rys
C
Spessart
Odenwald
100 km
na
n-
e
eß
see
Fig. 2
50°
e
Kellerwald
c
ei
m
co
Erfurt
re
u
ut
s
Rh
a
an
w
d
G
on
ian
g
in
ur e
h
ot on
ax z
Fig. 1. Schematic map showing the position of the field working area (Lahn–Dill and Kellerwald areas) within the Rhenohercynian zone of the Central European Variscides (see
rectangle for location of Fig. 2A, B).
Map is modified after Huckriede et al. (2004).
Baltica, which led to the closure of the Tornquist Sea. The ensuing collision of both Baltica and Avalonia with Laurentia closed the intervening Iapetus Ocean at about 420 Ma. As a result the continent Laurussia
(= Old Red Continent) was formed (Kroner et al., 2007; Linnemann et al.,
2008; Nance et al., 2010). According to Sánchez Martínez et al. (2007)
and Nance et al. (2010), the closure of the Rheic Ocean began in the
Late Silurian–Early Devonian and continued until the Early Carboniferous
by successive advancing from west to east.
Several plate tectonic models exist for the Variscan orogeny. The
two-plate model involves the collision of Laurussia (Baltica +
Avalonia + Laurentia) with West Gondwana (Amazonia + West
Africa + Cadomia) as a result of the closure of the Rheic Ocean
(e.g. Linnemann et al., 2004; Kroner et al., 2007; Linnemann et al.,
2010). The multi-plate model accounts for the existence of a number
of terranes and narrow oceans between the principal plates of Laurussia
and Gondwana (e.g. Franke, 2000; Tait et al., 2000; Torsvik and Cocks,
2004; von Raumer and Stampfli, 2008). Different microplates are
assigned to the “Armorican Terrane Assemblage (ATA)” by Franke
(1995) and are composed of Bohemia, Saxothuringia, Iberia and the
Armorican Massif. ATA is believed to have separated from Gondwana
during the Ordovician, following Avalonia in northward direction.
Between Avalonia and ATA, an island arc has been proposed, which
formed during the closure of the Rheic Ocean and was accreted to
Avalonia during the Early Devonian (Franke and Oncken, 1995;
Franke, 2000). This arc is documented by magmatic rocks of Silurian
age in the “Northern Phyllite zone” and in the “Mid-GermanCrystalline zone (MGCZ)” (Altenberger et al., 1990; Sommermann,
1993; Dombrowski et al., 1995; Reischmann et al., 2001).
One model for the Rhenohercynian proposes successive rifting of a
passive southern margin of Laurussia during Devonian times caused by
the opening of a Rhenohercynian Ocean (e.g. Franke, 2000; Tait et al.,
2000; Doublier et al., 2012). Conversely, Floyd (1982), Flick and Nesbor
(1988), Oczlon (1994), Smith (1996), von Raumer and Stampfli (2008),
and Zeh and Gerdes (2010) interpreted the Rhenohercynian as an
active continental margin whereby a northward dipping subduction
zone triggered the opening of a Rhenohercynian back arc ocean.
The MGCZ, a NE–SW trending basement wedge, represents the
Variscan collision zone and separates the Rhenohercynian zone
(Avalonia) from the Saxothuringian zone (Armorica; Zeh and
Gerdes, 2010; Zeh and Will, 2010). However, Zeh and Gerdes
(2010) presented provenance data that the MGCZ was not completely
Armorican, but was partly Avalonian as indicated by certain rock
units of the Ruhla Crystalline complex. Accordingly, the Spessart
and the Böllstein Odenwald can be interpreted as belonging to
Avalonia as well.
According to Plesch and Oncken (1999), the final Variscan continental collision occurred along two suture zones, one separating the
external Rhenohercynian zone from the Saxothuringian zone, the
other separating the latter from the Tepla-Barandium part of the
internal Moldanubian zone. However, collisions occurred earlier
between the different microcontinents of the ATA. For example,
Bohemia and Saxothuringia collided with the Armorican Massif and
consequently became part of Armorica during the Late Devonian
(Schäfer et al., 1997; Dörr and Zulauf, 2010). The final stage of the
Variscan orogeny is primarily characterized by diorites and granites,
usually of calc-alkaline composition (Altherr et al., 1999; Stein, 2001)
and with a distinct subduction signature (Henes-Klaiber, 1992). But
the rock pile was also overprinted by regional HT/LP metamorphism
and segmentation in folds and thrusts (Oncken, 1997; Kroner et al.,
2007). Within the Rheinisches Schiefergebirge, deformation combined
with very low to low-grade metamorphism progressed from the SE
(330 Ma) to the NW (300 Ma) in the Late Carboniferous (Ahrendt et
al., 1978, 1983). The Variscan orogeny led to a crustal shortening there
of, on average, about 50% (Behrmann et al., 1991; Oncken et al., 1999).
This is also suggested for the allochthonous units (Doublier et al., 2012).
Geochronological and kinematic studies (e.g. Ahrendt et al., 1983;
Engel and Franke, 1983; Klügel, 1997; Plesch and Oncken, 1999;
Huckriede et al., 2004) have improved our knowledge of the structural evolution of the Rhenohercynian fold and thrust belt. 40K/40Ar-ages
of detrital muscovites from non- to low-grade metamorphic sediments in the Hörre nappe provide information only about final
cooling below ~ 350 °C (Huckriede et al., 2004). U–Pb-ages of zircon
1486
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
grains, on the other hand, exhibit a much higher temperature regime
since zircons equilibrate at 800 °C or higher. Therefore, the original
crystallization age of the zircon or rather the age of the last magmatic
event survives later sedimentary or metamorphic processes. Accordingly, the provenance of these sediments is best deduced from the
distribution of zircon age populations.
The aim of this study is to present U–Pb LA–SF–ICP–MS ages of detrital zircons derived from Devonian and Lower Carboniferous siliciclastic
sediments of the southeastern Rheinisches Schiefergebirge. The samples were collected from autochthonous to par-autochthonous and
allochthonous units in order to clarify their stratigraphic, sedimentary,
magmatic and plate tectonic relationships and thereby obtain a better
understanding of Variscan collisional tectonics and the provenance of
particularly allochthonous units within the Rhenohercynian zone.
2. Geological setting and geotectonic units
2.1. Introduction
The Rheinisches Schiefergebirge is part of the Rhenohercynian zone
of the Variscan Orogen (Fig. 1). In its eastern part, several autochthonous, par-autochthonous (e.g. Wachendorf, 1986; Meischner, 1991;
Schwan, 1991; Bender and Königshof, 1994) and allochthonous units
(e.g. Engel et al., 1983; Oczlon, 1992, 1994; Franke, 2000; Salamon,
2003; Huckriede et al., 2004; Salamon and Königshof, 2010) have
been defined, the latter having been recognized by Kossmat (1927)
who was the first to propose a nappe tectonic concept for the
Rhenohercynian zone. In the eastern Rheinisches Schiefergebirge, samples were taken from different structural units: the Lahn–Dill area in the
A
40
Autochthonous to
par-autochthonous units
50
Lahn and Dill-Eder synclines
Lower Carboniferous
Devonian
60
Allochthonous units
34
70
80
10 km
Gießen nappe
Lower Carboniferous
Devonian
50
Lohra nappe
LahnLower Carboniferous
Devonian
Bicken-Ense and Wildestein
imbrication structure
Kammquartzite
imbrication structure
Steinhorn nappe
Lower Carboniferous
Devonian
Hörre nappe
Lower Carboniferous
Devonian
40
Dill
44440
Marburg
30
45439
44467
44442
45347
Dillenburg
44441
20
44468
45438
10
44443
Wetzlar
Gießen
56 00
Fig. 2. Geological maps of (A) the Lahn–Dill area and (B) the Kellerwald areas (modified after Meischner, 1966; Bender, 2006). Locations of samples are indicated by red dots.
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
B
34
98
02
1487
35
06
Autochthonous to par−autoch−
thonous units
10
Bad Wildungen
Dill−Eder syncline
Lower Carboniferous
64
Devonian
Bicken−Ense and Wildestein
imbrication structure
Kammquartzite
imbrication structure
Allochthonous units
Gießen nappe
Lohra nappe
44358
Lower Carboniferous
.
60
Devonian
Steinhorn nappe
Hörre nappe
Lower Carboniferous
Devonian
Armsfeld
56
44356
Haina
52
Jesberg
44352
45360
44353
56
0
0.4
0.8
1.2
1.6
2 km
48
44350
Gilserberg
Fig. 2 (continued).
southeast, and in the Kellerwald area, a partly separated region in the
northeastern Rheinisches Schiefergebirge. The Lahn–Dill area can be
subdivided into the autochthon, known as Lahn syncline and Dill–
Eder syncline, which are separated by the allochthon of the Hörre
nappe. Southeast of the latter, more allochthonous units occur, such as
the Steinhorn nappe and the Lohra nappe as a part of the Frankenbach
imbrication zone, and the Gießen nappe (Fig. 2A). Additionally, there
are several par-autochthonous units, which can be found imbricated
within the above mentioned autochthon and allochthon as a result of
strong deformation and thrusting. Except for the Lahn syncline all of
these structures extend into the Kellerwald area (Fig. 2B). The deformation over the entire area shows distinctive differences: the allochthonous and par-autochthonous units have experienced more intense
folding and thrusting, as compared to the autochthonous areas. For detailed sedimentological and facies analysis we refer to the references
cited in the text. The region's detailed biostratigraphic record is beyond
1488
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
the scope of this paper, although new and literature-derived biostratigraphic data were used in this study. Here we refer to the short description in the text (see below), which is essential for a better understanding
of the complex structures. The general stratigraphy and geotectonic
assignment of the investigated samples are shown in Table 1.
2.2. Lahn and Dill–Eder synclines
The geological record of the Lahn syncline and the Dill–Eder syncline generally extends from about the Middle Devonian to the end
of the Early Carboniferous and the onset of the Variscan orogeny.
Sedimentation in this area was generally controlled by basin development and volcanism, both reflecting the thinning of the crust due to
extensional tectonics. In terms of depositional development, the
Lahn and Dill–Eder synclines represent a single unit with the Dill–
Eder syncline in the north being more influenced by arenaceous
siliciclastic sediments derived from the Old Red Continent.
Intensive intraplate volcanism started in the early Givetian, lasting
with interruptions for about 50 Ma into the Carboniferous (Nesbor et
al., 1993; Nesbor, 2004; see chapter 3). The Devonian volcanic successions allowed for the overgrowth by reef limestones (so-called
“Massenkalk”), which flourished during the Givetian and Frasnian.
Depending on their depositional setting, reef structures are variable
in their morphology and thicknesses (e.g., Königshof et al., 1991;
Buggisch and Flügel, 1992; Braun et al., 1994; Königshof et al., 2010).
Volcanic activity and reef development led to a considerable submarine
relief in the Lahn–Dill and Kellerwald areas, resulting in widespread facies variations during the Upper Devonian. Further Upper Devonian
sedimentation in the Lahn and Dill–Eder synclines produced slates
and nodular calcareous slates in numerous basins, and condensed limestones on the swells. The slates were supplemented by thick sandstone
sequences, especially in the Dill–Eder-syncline (e.g., in the Thalenberg
Formation: Pirwitz, 1986; Wierich, 1999), whereas these are rare in
the Lahn syncline. Locally, minor amounts of primitive basaltic melts
ended the Devonian volcanic cycle.
Crustal thinning on the southern shelf of Laurussia continued into
the Early Carboniferous resulting in a moderate increase of water
depth in the eastern Rheinisches Schiefergebirge. This resulted in a
special facies realm called the “Culm facies” that caused a reduced
sedimentation initially characterized by black shales and cherts. This
sedimentation was later associated with intensive activity of the
Carboniferous volcanic cycle. The volcanic activity was followed by
argillaceous sedimentation and, subsequently, by graywackes at the
end of the Lower Carboniferous. The latter ones can reach thicknesses
of hundreds of meters and represent the final stage of deposition prior
to the Variscan orogeny in the area of interest (for overview and references see Bender and Stoppel, 2006; Bender, 2008; Königshof et al.,
2008; Nesbor, 2008; Königshof et al., 2010).
Table 1
Stratigraphic position and source areas of the sandstones and graywackes used in this study.
Series
Stage
Sample
Geographic region
Lithology
number
Jesberg
Kulm syncline
Armorican terrane A.
Kammquartzite
imbrication structure
Baltica
graywacke
Gießen nappe
Armorican terrane A.
Lahn−Dill area
graywacke
Hörre nappe
Armorican terrane A.
44353
KE−6
Kellerwald
graywacke
Lohra nappe
Armorican terrane A.
44442
Will−1
Lahn−Dill area
graywacke
44358
KE−11
Kellerwald
sandstone
Dill−Eder
syncline
Baltica
45347
Bot−1
Lahn−Dill area
sandstone
Famennian
44441
Bisch−2
Lahn−Dill area
graywacke
Hörre nappe
Armorican terrane A.
Frasnian/
Famennian
44350
KE−3
Kellerwald
graywacke
45438
Kirch−1
Gießen nappe
Armorican terrane A.
Frasnian
Lahn−Dill area
graywacke
Late Emsian
44467
Haig−1
Lahn−Dill area
sandstone
Siegen anticline
Baltica
Early Emsian
45360
KE−17
Kellerwald
graywacke
Steinhorn nappe
Armorican terrane A.
Early Emsian
45439
Herm−1
Lahn−Dill area
graywacke
44352
KE−5a
Kellerwald
graywacke
Tournaisian/
Visean
44356
KE−9c
Kellerwald
sandstone
Visean
44468
Kamm−1
Lahn−Dill area
sandstone
Visean
44443
Heu−2
Lahn−Dill area
Tournaisian/
Visean
44440
Eln−2
Tournaisian/
Visean
Late
Early
Carboniferous
Tournaisian
Famennian
Early
Late
Devonian
Famennian
Early
Initial geotectonic position
geotectonic unit
Visean
Tournaisian/
Devonian
Geological/
Early
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
2.3. Hörre nappe
The sedimentary development of the Hörre nappe began in the early
Famennian. It differs profoundly from that of coeval strata in the Lahn
and Dill–Eder synclines (e.g., Bender, 1978; Bender and Homrighausen,
1979; Bender, 1989, 2008). For example, volcanic rocks are notably
absent in contrast to the entire Rhenohercynian zone where repeated
intense and sometimes explosive volcanism occurred. It is important
to note, however, that ash layers are intercalated in the sedimentary
pile of some formations of the Hörre nappe, such as in the Gladenbach
Formation (Early Carboniferous). The succession of the Hörre nappe
starts with shales overlain by sandstones and graywackes (Ulmbach
Formation). Occasionally detrital limestones are intercalated. The overlaying Weitershausen Formation is mainly composed of shales and
detrital, thin-bedded carbonates. These rocks are overlain by shales,
quartzitic sandstones and thin-bedded graywackes of the Endbach Formation of Early Carboniferous age. The overlying Gladenbach Formation
is mainly composed of siliceous shales, alum black shales, cherts and
turbiditic limestones. These sediments are succeeded by the Bischoffen
Formation consisting of shales with intercalations of graywackes, and
finally the Elnhausen Formation composed mainly of shales and thickbedded graywackes (Bender, 2008).
2.4. Bicken–Ense and Wildestein imbrication structures
The Bicken–Ense and Wildestein imbrication structures have long
been regarded as parts of the Dill–Eder syncline bordering the Hörre
unit (e.g., Bender, 1997), whereby the Wildestein imbrication structure
is imbricated with rocks of the Hörre nappe. Because of differences in
their sedimentological records compared to adjoining rock units of the
same age and their tectonic contact (Bender et al., 1997), they are to
be considered to be par-autochthonous units. The Bicken–Ense imbrication structure consists almost entirely of the Bicken Formation, which
is mainly composed of shales and fossiliferous, pelagic carbonates
(e.g. cephalopod limestones) of Early Devonian (Emsian) to Early
Carboniferous age (Bender, 1997). The Wildestein imbrication structure
is composed of reddish and gray shales and siliceous shales of the
Wildestein Formation, which is assumed to be mainly of late Famennian
age.
2.5. Kammquartzite imbrication structure
The Kammquartzite imbrication structure can be traced for more
than 300 km throughout the Rhenohercynian zone (Figs. 1, 2A, B).
This and the dominance within the unit of quartzitic sandstones of
Early Carboniferous age have been the focus of geological investigations
for a long time (Koch, 1858; Kockel, 1958; Schriel and Stoppel, 1958;
Homrighausen, 1979; Wierich and Vogt, 1997; Jäger and Gursky,
2000; Huckriede et al., 2004). Originally regarded as part of the Hörre
unit, the Kammquartzite imbrication structure is now considered a separate tectonic unit since it can be found within the Dill–Eder syncline,
the Wildestein imbrication structure, the Hörre nappe, the Frankenbach
imbrication zone, and beneath the Gießen nappe. Reflecting its intensive deformation, the quartzitic sandstone slices range in a size from
more than 100 m to less than 1 cm, embedded in a melange composed
of very strongly deformed dark gray and siliceous shales.
2.6. Frankenbach imbrication zone
The Frankenbach imbrication zone, as defined by Bender (2006), has
a stratigraphical range from Early Devonian to Early Carboniferous. It
is generally strongly deformed and exhibits numerous imbricate structures. The zone consists, on the one hand, of slices of autochthonous
rocks typical for the Lahn and Dill–Eder synclines, although massive
reef limestones (Massenkalke), which are common in the Lahn syncline
and in the northern part of the Rheinisches Schiefergebirge (e.g., Pas et
1489
al., 2013 and references therein), are absent. On the other hand, the
allochthonous units of the Lohra and Steinhorn nappes, as well as slices
of the Hörre nappe, are intercalated in the Frankenbach imbrication
zone. The par-autochthonous Kammquartzite is also imbricated locally
into this zone.
2.7. Lohra nappe
The Lohra nappe represents an isolated allochthonous unit in the
Kellerwald area, whereas to the southwest in the Lahn–Dill area, it
is tectonically integrated into the Frankenbach imbrication zone
(Fig. 2A, B). The Lohra nappe consists of the Famennian Lohra unit
composed of shales and cherts, with intercalations of sandstones
and graywackes, and the Kehna unit of Lower Carboniferous age composed of graywackes and shales.
2.8. Steinhorn nappe
The Steinhorn nappe is here defined as a new tectonic unit with an
isolated allochthonous setting in the Kellerwald area (see Fig. 2A, B).
Comparable to the Lohra nappe, it can be traced into the Frankenbach
imbrication zone as part of the Lahn–Dill area. The Steinhorn nappe consists primarily of Early Devonian sandstones, graywackes (ErbslochGrauwacke), shales and carbonates (e.g., the Greifenstein and Schönau
limestones), which exhibit Bohemian faunal affinities (e.g. Flick, 1999).
Middle/Upper Devonian sequences are represented by shales, alum
shales, cherts and graywackes. Those of Early Carboniferous age are
composed of shales, cherts and a melange unit of Devonian and Lower
Carboniferous rocks. The entire sedimentological record differs from
other structural units described in this paper, and is here used to define
a new tectonic unit.
2.9. Gießen nappe
The long-disputed Gießen nappe (Fig. 2A, B) was the first allochthonous unit established in the Rheinisches Schiefergebirge (e.g. Eder et al.,
1977; Engel et al., 1983), and covers an area of about 250 km2. The rocks
of the Gießen nappe consist of shales, clayish sandstones, cherts and
graywackes, with the latter predominating. At its base, condensed
phyllitic shales and radiolarian cherts are imbricated with tectonic slices
of MORB-type metabasalts (Wedepohl et al., 1983; Grösser and Dörr,
1986). The Gießen nappe sequence spans the Early Devonian to Early
Carboniferous and can be subdivided into two structural units containing graywackes of different ages (Birkelbach et al., 1988; Dörr,
1990). Graywackes of the smaller northern unit are of Frasnian age,
whereas those of the larger southern unit are Early Carboniferous (for
a detailed stratigraphic and lithologic description see Dörr, 1986).
2.10. Lower Devonian of the Siegen anticline
The Siegen anticline consists of Lower Devonian rocks which are
characterized by generally monotonous siliciclastic sequences delivered
from the Old Red Continent. Sediments of the Lochkovian are up to
1000 m thick and are mainly composed of reddish mudstones and siltstones with intercalated coarse-grained sandstones and quartzites.
Siegenian (mainly Pragian) sediments (the Siegenian stage is used
herein in a traditional German sense) of the Siegen anticline comprise
more than 4000 m of mudstones, siltstones and fine-grained sandstones with distinct fossiliferous horizons containing abundant plant remains. The lithological character of the lower Emsian sediments is very
similar but changes in the upper Emsian due to a general transgressive
trend (Poschmann and Jansen, 2003; Mittmeyer, 2008). Shallow water,
coarse-grained quartzitic sandstones are overlain by argillaceous sediments. Importantly, an intense period of volcanism is documented by
numerous intercalations of silica-rich, mainly volcaniclastic rocks in
the Lower Devonian sediments (see chapter 3).
1490
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
1
ash tuff; Early
Carboniferous
(Kubanek & Zimmerle
1986; Dehmer et al.
1989)
Comendite
Pantellerite
Phonolite
Rhyolite
ash tuff; Early
Carboniferous
ash tuff; Frasnian
ash tuff; Eifelian
ash and lapilli tuff;
Early Devonian
(Kirnbauer 1991)
0.1
Trachyte
Zr/ TiO2
Rhyodacite
Dacite
Trachyandesite
Andesite
0.01
Basanite
Nephelinite
Andesite/ Basalt
Alkali-Basalte
Sub-Alkaline Basalt
Kellerwald area;
Early Carboniferous
phase
Lahn-Dill area;
Early Carboniferous
phase
Lahn-Dill area;
Famennian
phase
Kellerwald area;
Givetian-Frasnian
phase
Lahn-Dill area;
Givetian-Frasnian
phase
0.001
0.1
1
10
Nb/Y
Fig. 3. Classification of the volcanic rocks of the Rhenohercynian zone in the Zr/TiO2 versus Nb/Y-diagram of Winchester and Floyd (1977). For locations of samples see Tables 1–3,
Electronic supplement.
3. Volcanism
Sedimentation on the southern shelf of Laurussia was accompanied by a wide diversity of volcanic processes. The oldest volcanic
rocks are preserved in the Phyllite zone at the southern margin of
the Rheinisches Schiefergebirge. Radiometric dating of these metarhyolites, -rhyodacites and -andesites shows them to be Silurian age
(Sommermann et al., 1992, 1994; Meisl, 1995). Rhyodacitic to rhyolitic magmatism continued through the Devonian until the Visean.
Widespread ignimbrites and volcanic fall out deposits of the Lower
Devonian (often redeposited) can be found between the Ebbe anticline (Heyckendorf, 1985) in the northwest and the southern part
of the Taunus (Kirnbauer, 1991) at the southern margin of the
Rheinisches Schiefergebirge (Figs. 3, 4). In the southeastern part of
the Rheinisches Schiefergebirge the source of these pyroclastics,
which are frequently intercalated within the siliciclastic rock pile, is
uncertain. Kirnbauer (1991) proposed a source area to the NE, E or
1000
Lower Carboniferous
rock / primitive mantle
Frasnian
Eifelian
100
Early Devonian
(Kirnbauer 1991)
10
Cs
Rb
Ba
Th
U
Nb
Ta
K2O
La
Ce
Pb
Pr
Nd
Sr
Sm
Hf
Zr
TiO2
Eu
Gd
Tb
Dy
Ho
Y
Er
Tm
Yb
Lu
1
Fig. 4. Element concentrations in dacitic/rhyodacitic lapilli and ash tuffs intercalated in the
sedimentary succession of the Devonian and Lower Carboniferous of the Rheinisches
Schiefergebirge in the order of increasing compatibility normalized to primitive mantel
(Hofmann, 1988). For locations of samples see Table 3, Electronic supplement.
SE of the Rheinisches Schiefergebirge with deposition of the pyroclastic material on the outer shelf area or on the slope of the southern
margin of Laurussia.
Most of the volcanic products are well preserved because of very
low-grade metamorphic overprinting. As a result, the typical phenocrysts in the rhyodacites (predominantly quartz and plagioclase,
with some alkali feldspar and biotite) can be identified. The feldspar
is generally albitized and biotite is often altered to chlorite by diagenetic and metamorphic processes. The original glassy matrix is also
completely altered to secondary minerals. Nevertheless, the rock
type can be classified by immobile elements (Fig. 3; Tables 1–3, see
Electronic supplement). Primary particles in the frequently coarse
grained pyroclastic deposits are often highly vesicular, and Y-shaped
former glass shards are a typical feature (Scherp, 1983).
At the beginning of the Eifelian a distinct facies change is evident
as very distal ash fall deposits occur. Nevertheless, the dacitic/
rhyodacitic composition can still be identified by immobile elements.
The volcanic ashes are intercalated as numerous extremely finegrained layers within the thick pile of autochthonous sedimentary
rocks, and are therefore often overlooked (Dehmer et al., 1989). In
the Visean, the grain size of the ash layers increases again (van
Amerom et al., 2001). Over the entire time span from the Early Devonian to the Early Carboniferous the mineralogical and geochemical
variation of these pyroclastic deposits is remarkably low (Fig. 4,
Table 3 Electronic supplement). But all of these rocks have been
subjected to diagenetic alteration and metamorphic overprint. The
main difficulty, however, lies in the evaluation of the intensity of contamination with epiclastic (sedimentary) debris and the degree to
which sorting processes during their redeposition modified their
original composition. Consequently, the immobile HFSE values do
not correspond definitively to those of recent fresh samples of subduction related dacites or rhyodacites.
While this dacitic/rhyodacitic volcanism originated far from the
Rheinisches Schiefergebirge, intensive intraplate volcanism started
on the subsiding southern shelf of Laurussia in the early Givetian.
Hence, a Devonian and an Early Carboniferous cycle can be distinguished (Nesbor, 1997, 2004, 2007). The Devonian cycle comprised
two phases, a bimodal (Givetian–Frasnian) main phase and a primitive basaltic (Famennian) late phase.
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
During the Early Carboniferous continued extension on the southern shelf of Laurussia combined with an increasing subsidence of the
continental crust led to a moderate increase of water depth. These
processes are reflected in a change in the sedimentary environment
and renewed volcanic activity of the Early Carboniferous cycle. The
new melts were almost entirely tholeiitic, in contrast with the Devonian alkali basaltic to basanitic compositions, but still exhibit intraplate affinities. Noticeable is the increase of aluminum in picotites
enclosed in the olivine phenocryts of the basalts. This is an indication
of ascending juvenile magma from deep mantle sources.
Sills and dikes of the Devonian and Carboniferous cycles are widely distributed in the Rheinisches Schiefergebirge, demonstrating the
widespread extent of the volcanic activity, well beyond the Lahn–
Dill and Kellerwald areas.
4. Methods
4.1. Sample material and stratigraphy
Thirty-three samples were investigated, but only 15 representative
examples are presented because of page limitations. The additional
1491
samples will be published separately in the context of a more regional
framework. The samples collected for U–Pb analysis from the different
structural units (Fig. 2A, B and Table 1) are described in stratigraphic
order.
The stratigraphically oldest sample (sample # 45439 [Herm-1],
Damm-Mühle Formation) of early Emsian age was taken from the
northeastern part of Steinhorn nappe in the Lahn–Dill area. This graywacke was formerly interpreted as being equivalent to the so-called
“Erbsloch-Grauwacke” in the Kellerwald (e.g. Kupfahl, 1953; Ziegler,
1957), but according to Jahnke (1971) profound facies and faunal differences separate the two. Sample # 45360 (KE-17) was taken at the
type location of the “Erbsloch-Grauwacke” in the Steinhorn nappe in
the Kellerwald area and is also of early Emsian age (Plusquellec and
Jahnke, 1999; Fig. 3). Sample # 44467 (Haig-1) was taken from a
quartzitic sandstone (the so-called “Ems-Quarzit”) of the Siegen
anticline and is of late Emsian age. It is used as a reference sample
for siliciclastic sediments derived from the Old Red Continent.
Sample # 45438 (Kirch-1) comes from the (Devonian) graywackes of the Gießen nappe in the Lahn–Dill area, which are covered
by siliceous shales and, according to Dörr (1986), is of Frasnian age.
Equivalent rocks from the Kellerwald area (sample # 44350 [KE-3])
Table 2
Analytical results from all investigated samples.
Sample name
44352
(KE-5a)
44356
(KE-9c)
44468
Lower
Carboniferous
(Kamm-1)
44443
(Heu-2)
Concordant grains
(90-110%)
Youngest
grain (in Ma)
(2σ in Ma)
Oldest
grain (in Ma)
(2σ in Ma)
118
58
368 ± 9
3006 ± 21
Neoproterozoic
142
108
432 ± 11
2873 ± 23
Mesoproterozoic
133
102
874 ± 26
2716 ± 23
Mesoproterozoic
90
71
345 ± 8
2602 ± 16
Neoproterozoic
141
64
322 ± 10
3036 ± 23
Lower Carboniferous
158
83
330 ± 10
2350 ± 41
Neoproterozoic
176
69
393 ± 8
2187 ± 17
Neoproterozoic
180
108
396 ± 15
1968 ± 45
Paleoproterozoic
82
36
423 ± 11
2740 ± 18
Mesoproterozoic
189
62
413 ±13
3358 ± 65
Neoproterozoic
119
51
359 ± 9
2149 ± 32
Neoproterozoic
115
57
384 ± 9
2459 ± 27
Neoproterozoic
128
74
419 ± 14
2008 ± 45
Mesoproterozoic
58
26
544 ± 14
2431 ± 49
Neoproterozoic
173
98
471 ± 10
3189 ± 29
1925
1069
Measured grains
44440
(Eln-2)
44353
(KE-6)
44442
(Will-1)
44358
(KE-11)
45347
(Bot-1)
Upper
Devonian
44441
(Bisch-2)
44350
(KE-3)
45438
(Kirch-1)
44467
(Haig-1)
Lower
Devonian
45360
(KE-17)
45439
(Herm-1)
Total:
Main part of all grains
Neoproterozoic
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
0.012
0.004
14
44352 (KE-5a)
n=58/118
Jesberg-Kulm
10
6
Syncline
2
12
0.000
0.006
0.002
0.008
0.004
0.000
0.012
Probability
0.008
0.004
44356 (KE-9c)
n=108/142
Kammquartzite
Imbric. Struct.
8
4
0
10
6
44468 (Kamm-1)
n=102/133
Kammquartzite
Imbric. Struct.
2
20
16
44443 (Heu-2)
n=71/90
Gießen
Nappe
12
8
4
0
0.000
20
0.014
0.010
0.006
Frequency
0.008
Ar
ch
ea
n
C
ar
b
D oni
ev fe
Si on rou
l u ia s
O ria n
rd n
o
C vic
am ia
br n
ia
n
N
eo
pr
ot
.
A
Pa
le
op
ro
t.
Dill area. Lycopsidan plant remains have been found in the Kehna
graywacke suggesting a late Tournaisian age (Kerp et al., 2006; Fig. 3).
The equivalent in the Kellerwald is the Hundshausen graywacke
(sample # 44353 [KE-6]), which is also of Early Carboniferous age
(Bender and Stoppel, 2006).
Sample # 44440 (Eln-2) is from the Hörre nappe in the Lahn–Dill
area (Fig. 2A) and is a graywacke of Early Carboniferous age (Elnhausen
Formation). This formation is the youngest sequence of the Hörre nappe
beginning in the Pericyclus stage (cu II), its stratigraphical extension is
unknown (Bender, 1989; Entenmann, 1991; Bender et al., 1997).
Sample # 44443 (Heu-2, Fig. 2A, Table 1) is a coarse-grained graywacke from the Gießen nappe in the Lahn–Dill area, and is of Early
Carboniferous age based on plant remains of Lepidodendron lossenii
Weiss (Henningsen, 1962). According to Mosseichik (2007) these
are also composed of graywackes and siliceous shales and are of
Frasnian/Famennian age. The oldest sample from the Hörre nappe
in the Lahn–Dill area (sample # 44441 [Bisch-2]), is a graywacke
belonging to the Ulmbach Formation and is of early Famennian age.
Samples from Famennian sandstones of the Dill–Eder syncline
have been taken from the Lahn–Dill area (sample # 45347 [Bot-1];
early Famennian) and from the Aschkoppen sandstone (Fig. 2A, B,
Table 1) in the northern Kellerwald (sample # 44358 [KE-11];
Famennian). Both samples are reference samples for the autochthonous Dill–Eder syncline.
The remaining samples are from the Lower Carboniferous: sample
# 44442 (Will-1) belongs to Lohra nappe (Kehna Graywacke Formation), which is a part of the Frankenbach imbrication zone. This unit is
located between the Hörre nappe and the Gießen nappe in the Lahn–
M
es
op
ro
t.
1492
16
44440 (Eln-2)
n=64/141
Hörre
Nappe
12
8
4
0.002
0
0.016
0.012
0.008
20
44353 (KE-6)
n=83/158
Lohra
Nappe
16
12
8
4
0.004
0.000
0.008
0.004
0
44442 (Will-1)
n=69/176
Lohra
Nappe
12
8
4
0
3600
3200
2800
2400
2000
1600
1200
800
600
500
400
300
0.000
age (Ma)
Fig. 5. A, B: Combined binned frequency and probability density distribution plots in the age range of 300 and 3600 Ma from 15 representative samples of the Lahn–Dill and
Kellerwald areas. All grains are concordant in the range between 90 and 100%.
Pa
l
eo
Ar
ch
e
pr
an
ot
.
1493
t.
ro
es
op
M
ar
C
D
B
bo
n
ev ifer
o
Si ni ous
lu a n
O r ia n
rd
o
C vici
am an
br
ia
n
N
eo
pr
ot
.
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
0.012
0.008
0.004
14
44358 (KE-11)
n=108/180
Dill-Eder
Syncline
10
6
2
12
0.000
0.006
0.002
0.008
0.004
45347 (Bot-1)
n=36/82
Dill-Eder
Syncline
8
4
0
44441 (Bisch-2)
n=62/189
Hörre
Nappe
10
6
2
0.000
16
Probability
0.006
0.002
0.008
0.004
44350 (KE-3)
n=51/119
Gießen
Nappe
12
8
4
0
10
45438 (Kirch-1)
n=57/115
Gießen
Nappe
6
2
12
0.000
0.006
0.002
0.010
0.006
0.002
Frequency
0.010
44467 (Haig 1)
n=74/128
Siegen
Anticline
8
4
0
45360 (KE-17)
n=26/58
Steinhorn
Nappe
12
8
4
0
0.016
0.012
0.008
0.004
20
45439 (Herm-1)
n=98/173
Steinhorn
Nappe
16
12
8
4
0
3600
3200
2800
2400
2000
1600
1200
800
600
500
400
300
0.000
age (Ma)
Fig. 5 (continued).
plants have a stratigraphic range from the late Tournaisian to the
Visean. The plant is also known from a Lower Carboniferous section
in Plauen (Vogtland) and has been described by Knaus (1994) and,
earlier, by Daber (1959) as being of probable Visean age.
Two samples were taken from the complex Kammquartzite imbrication structure. Sample # 44468 [Kamm-1] is a quartzitic sandstone
(Kammquartzite) of Early Carboniferous age from the Lahn–Dill area
(Jäger and Gursky, 2000). Wierich and Vogt (1997) proposed a new
formation name (“Bruchberg-Sandstein”-Formation) and placed the
sandstones and quartzites in the Pericyclus stage (early Visean)
based on spore assemblages. According to Jäger and Gursky (2000)
deposition of the Kammquartzite Formation started at the base of
the Visean and lasted into the late Visean. Sample # 44356 (KE-9c)
of the Kammquartzite imbrication structure comes from the central
part of the Kellerwald.
Another graywacke of Early Carboniferous age (sample # 44352
[KE-5a]; Fig. 2B, Table 1) was collected from the “Jesberg–Kulm syncline”
in the southern part of the Kellerwald. These graywackes conformably
overlie alum shales, which are Early Carboniferous (Tournaisian to
Visean) according to Jahnke and Paul (1968).
Samples used for geochemical analyses in the different units are
shown in Figs. 3 and 4 and in Tables 1–3 (Electronic supplement).
4.2. U–Pb detrital zircon ages and geochemistry
Zircon concentrates were separated from 2 to 4 kg sample material at the Senckenberg Naturhistorische Sammlungen Dresden (SNSD)
using magnetic and heavy liquid (LST) methods. Final selection of the
zircon grains for U–Pb dating was achieved by hand-picking under a
binocular microscope. Zircons of all grain sizes and morphological
1494
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
types were selected, mounted in resin blocks and polished to half
their thickness. Cathodoluminescence (CL)-images of zircon were
produced with a Zeiss scanning electron microscope EVO 50. All
grains were analyzed for U, Th, and Pb isotopes by LA–SF–ICP–MS
techniques at the Museum für Mineralogie und Geologie (GeoPlasma
Lab, SNSD), using a Thermo-Scientific Element 2 XR sector field
ICP–MS coupled to a New Wave UP-193 Excimer Laser System. A
teardrop-shaped, low volume laser cell constructed by Ben Jähne
(Dresden) and Axel Gerdes (Frankfurt/M.) was used to enable sequential sampling of heterogeneous grains (e.g. growth zones) during time
resolved data acquisition. Each analysis consisted of approximately
15 s background acquisition followed by 30 s data acquisition, using a
laser spot-size of 25 and 35 μm, respectively. A common-Pb correction
based on the interference- and background-corrected 204Pb signal and
a model Pb composition (Stacey and Kramers, 1975) was carried out if
necessary. The necessity of the correction is judged on whether the
corrected 207Pb/206Pb lies outside of the internal errors of the measured
ratios. Discordant analyses were generally interpreted with care. Raw
data were corrected for background signal, common Pb, laser induced
elemental fractionation, instrumental mass discrimination, and timedependant elemental fractionation of Pb/Th and Pb/U using an Excel®
spreadsheet program developed by Axel Gerdes (Institute of Geosciences,
Johann Wolfgang Goethe-University Frankfurt, Frankfurt am Main,
Germany). Reported uncertainties were propagated by quadratic addition of the external reproducibility obtained from the standard zircon
GJ-1 (~0.6% and 0.5–1% for the 207Pb/206Pb and 206Pb/238U, respectively)
during individual analytical sessions and the within-run precision of
each analysis. Concordia ages (95% confidence level) were produced
using Isoplot/Ex 2.49 (Ludwig, 2001) and frequency and relative probability plots made use AgeDisplay (Sircombe, 2004). The207Pb/206Pb age
was taken for interpretation for all zircons >1.0 Ga, and the 206Pb/238U
ages for younger grains. Further details of the instrument settings are
available from Table 4 (see Electronic supplement). For further details
on analytical protocol and data processing see Gerdes and Zeh (2006)
and also Frei and Gerdes (2009). The uncertainty in the degree of concordance of Precambrian–Paleozoic grains dated by the LA–ICP–MS
method is relatively large and results obtained from just a single analysis have to be interpreted with care. A typical uncertainty of 2–3% (2σ)
in 207Pb/206Pb for a Late Neoproterozoic grain (e.g., 560 Ma) corresponds to an absolute error on the 207Pb/206Pb age of 45–65 Myr.
Such a result leaves room for interpretation of concordance or slight
discordance — the latter one could be caused by episodic lead loss, fractionation, or infiltration Pb isotopes by a fluid or on micro-cracks. Thus,
zircons showing a degree of concordance in the range of 90–100% in this
paper are classified as concordant because of the overlap of the error
ellipse with the concordia (e.g. Jeffries et al., 2003; Linnemann et al.,
2007; Frei and Gerdes, 2009). Th/U ratios (Tables DR-3 to DR-9, data repository) were obtained from the LA–ICP–MS measurements of investigated zircon grains. U and Pb content and Th/U ratio were calculated
relative to the GJ-1 zircon standard and are accurate to approximately
10%.
Geochemical samples were analyzed at the Hessisches Landesamt
für Umwelt und Geologie (HLUG) and the Geowissenschaftliches
Zentrumof the Universität Göttingen (GZG). RFA analyses were determined with sequential X-ray Spectrometer SRS 3000 (Siemens) at the
HLUG and the Axios X-ray Spectrometer (PANalytical) at the GZG.
REE were analyzed with ICP–MS Elan 500 (Perkin Elmer) at the
HLUG and with DRC II (Perkin Elmer) at the GZG.
4.3. Results
In total, we analyzed U–Th–Pb isotopes of 1925 detrital zircon
grains, of which 1067 showed U–Pb ages with a degree of concordance
in the range of 90–100%. Zircons were obtained from 15 arenaceous
samples of the Lahn–Dill and Kellerwald areas. Data are available in
Tables 5 to 19 in the data redepository (Electronic supplement).
Table 2 lists the results regarding measured and concordant grains,
youngest and oldest age, and largest zircon population. Combined
frequency and probability density distribution plots of each sample
are given in Fig. 5A, B.
All samples of the nappe units (Hörre-, Steinhorn-, Lohra- and
Gießen nappe) and of the Jesberg–Kulm syncline show a gap in
Mesoproterozoic ages. This is typical for a provenance in the West
African craton or a related terrane (e.g. Linnemann et al., 2007). The
rock units in the nappe complexes of the Lahn–Dill and Kellerwald
areas showing such zircon populations are related to the Gondwanan
rock units of the Saxothuringian zone. Only few single grains (up to 2
grains per sample) show ages between 1.6 and 1.0 Ga. Such low input
can be explained by long-distance along-shore transport and is not significant. All other samples have their main peaks in the Mesoproterozoic
and, hence, have a Laurussian affinity (Linnemann et al., 2012). In 3 samples of the Lower Carboniferous nappe units, distinct peaks around
360 Ma occur. This peak is unique and reflects the input from a magmatic arc source that was active during the closure of the Rheic Ocean.
The results based on the geochemical analyses have been already
presented in Section 3 within the framework of the volcanic development of the study area.
5. Discussion
The principal plates and microplates of Gondwanan Europe and
their adjoining cratons show distinct differences in the timing of
tectonomagmatic and metamorphic events. These orogenic and plate
tectonic activities are mirrored in the U–Pb zircon ages of siliciclastic
sediments. Because of different geological histories, each continent exhibits its own particular zircon population in the archive of detrital
sediments such as sandstones and graywackes. The typical zircon
populations from cratons and microplates relevant to this paper are
presented in Fig. 6. The results are compiled from our own and published
data derived from Precambrian and Paleozoic sandstones (Linnemann et
al., 2012 and references therein). The northern margin of the Rheic
Ocean is characterized by an Avalonian and Baltic hinterland and
therefore delivered Lower Paleozoic sediments that are characterized
by a significant population of zircon grains of Mesoproterozoic age
(1600–1000 Ma) as well as those with ages of ~450 Ma and ~420 Ma.
Both latter ages originated from events caused by the closure of the
Tornquist Sea (Avalonia docking on Baltica at ~450 Ma) and the closure
of the Iapetus Ocean (collision of Baltica + Avalonia with Laurentia at
~420 Ma). In contrast, Lower Paleozoic sandstones formed at the southern margin of the Rheic Ocean are derived from a West African hinterland. In this case, ages of detrital zircon populations show a distinct gap
in the record of Mesoproterozoic grains. These special features of zircon
population patterns from sandstones from different paleocontinents
such as Laurussia and Gondwana are a tool for the differentiation and
characterization of siliciclastic sediments in orogenic collisional zones.
Based on these requirements, we assign the provenance of the Devonian
and Lower Carboniferous siliciclastics from the Lahn–Dill and Kellerwald
areas to the Rhenohercynian and Saxothuringian margins. As a result,
we present a plate tectonic model based on the U–Pb ages of the detrital
zircon grains obtained from the 15 investigated sandstones and graywackes as well as on geochemical proxies of the volcaniclastics.
The spectra of zircon ages from siliciclastic Devonian to Lower
Carboniferous sediments derived from different tectonic units of the
Lahn–Dill and the Kellerwald areas allow a distinction of two source
areas. The samples of sedimentary rocks of the autochthonous to parautochthonous units show a distinct provenance from Baltica whereas
those from the allochthonous units were derived from a Gondwanan
source (Fig. 5). Typical Baltican signatures have been found in the
Lower Devonian sandstones of the Siegen anticline, the Upper Devonian
sandstones of the Dill–Eder syncline and in the “Kammquartzite” of the
Lower Carboniferous. The mainly Mesoproterozoic age spectrum points
to Svecofennian and Karelian source areas. Remarkably, the signal
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
3200
3300
3400
3500
3400
3500
3100
3100
3300
3000
3000
Palaeoarch.
3200
2800
2900
2900
2700
2400
2400
Mesoarch.
2800
2300
2300
2700
2200
2200
2500
2100
2100
2600
2000
2000
2600
1900
1900
2500
1700
1800
1600
1800
1500
1600
Neoarch.
1700
1400
1500
1100
1300
1000
1100
1400
900
1000
1200
800
900
1300
700
800
Palaeoproterozoic
1200
600
Mesoprot.
700
age (Ma)
Neoprot.
600
400
500
Pz
1495
Baltica
Amazonia
East Avalonia
(Brabant
Massif)
Dev
Sil
Ord
Camb
Cadomia
Saxo-Thuringian
Moldanubian
age (Ma)
400
500
West African Craton
Pz
Neoprot.
Mesoprot.
Palaeoproterozoic
igneous and metamorphic ages (all methods)
igneous (incl. inheritance) and metamorphic zircon
detrital zircon
igneous (incl. inheritance) from pre-Silurian rocks/protoliths
Neoarch.
Mesoarch.
Palaeoarch.
Dev - Devonian
Sil - Silurian
Ord - Ordovician
Cam - Cambrian
Pz - Palaeozoic
Fig. 6. Detrital zircon age distributions for Baltica, Amazonia, East Avalonia (Brabant massif), Armorica (Saxothuringian and Moldanubian zones), and the West African craton (data
compilation from Linnemann et al., 2004; Drost et al., 2010; Linnemann et al., 2011, 2012). Note the general occurrence of Mesoproterozoic zircon ages in Baltica, Amazonia
(Avalonian hinterland), and Avalonia, and the contrasting scarcity of the same zircon populations in Armorica and the West African craton.
for Avalonia is very weak. This implies that the southern shelf of
Laurussia (Avalonia) was covered by a thick pile of debris derived
from Precambrian rocks of Baltica. Only few outcrops of the Avalonia
terrane must have projected through of the plain of Baltic sediments,
and so were subject to erosion. Another option could be tectonic
stacking, caused by plate tectonic movements (i.e., Baltica was thrust
over Avalonia). But such a scenario requires the occurrence of highgrade metamorphic rocks, which are absent in the Rhenohercynian
zone.
In contrast, all graywacke samples of the Hörre, Steinhorn, Lohra
and Gießen nappes exhibit a gap between 1.7 and 1.0 Ga within the
Mesoproterozoic zircon ages. Only very few isolated zircons of
Mesoproterozoic age have been detected. Altogether, the analyzed detrital zircon populations are characteristic for Gondwana, especially
for the Saxothuringian terrane. These results confirm former interpretations of the geotectonic history of the Hörre and Gießen nappes. Based
on the profound sedimentological differences, such as the occurrence
of graywackes as early Famennian and the absence of volcanic rocks
(Bender, 1978; Bender and Homrighausen, 1979; Homrighausen,
1979; Entenmann, 1991; Herbig and Bender, 1992), as well as facies
differences (e.g. Herbig and Bender, 1992; Groos-Uffenorde et al., 2000;
Bender and Blumenstengel, 2003), the Hörre unit has been interpreted
to be part of the allochthonous unit of the Hörre/Gießen nappe
which was derived from the south of the Rheinisches Schiefergebirge
(e.g. Franke, 1995; Oncken and Weber, 1995; Bender, 2006). Bender
and Königshof (1994) considered the Hörre zone to be an autochthonous unit due to the metamorphic patterns based on conodont color
alteration of conodonts (CAI), but this interpretation has been
disproved.
Further to the east, similar rocks occur in the Kellerwald and the Harz
Mountains, which have been assigned to the “Hörre–Gommern zone”
(e.g., Stoppel, 1961; Eder et al., 1969; Bender and Homrighausen, 1979;
Engel et al., 1983; Meischner, 1991; Bender, 2006; Fig. 1). This structural
unit strikes SW/NE for about 300 km and is about 5 to 15 km wide.
Based on our results the “Hörre–Gommern zone” as defined by many authors (see above), is not a single structural unit because of zircon data
point to different provenances. The Hörre nappe clearly shows zircons
delivered from a West African hinterland, whereas the Kammquartzite
imbrication structure exhibits a Baltica signature. In contrast to these
results, Huckriede et al. (2004) assumed a depositional setting for the
Hörre sediments close to the margin of Laurussia. Haverkamp et al.
(1992) assigned the “Hörre–Gommern Quartzite” (=Kammquartzite)
to the distal passive margin of Laurussia. Doublier et al. (2012) proposed
a pelagic facies setting for the successions of the Hörre and Gießen
nappes and a lower slope setting for the “Kammquartzite”. Despite
some distal turbidites present for the latter, which were deposited in a
narrow channel from NE to SW (e.g. Homrighausen, 1979; Jäger and
Gursky, 2000), there are no convincing indications of a deep water facies
in the “Kammquartzite”. However, there are clear indications for a shallow water environment. Some sections within the “Kammquartzite”
show prominent hummocky cross stratification. Such ripple structures
are known to occur in the lower part of a shoreline sequence, especially
in fine grained sandstones (Reineck and Singh, 1980). Tongue-like or
lobe-like small current ripple marks (lingoid ripples) can also be recognized, indicating a shallow water environment (see also Jäger and
Gursky, 2000). Additionally, fossil assemblages in the Steinhorn nappe,
situated south of the Hörre nappe (Fig. 2A, B), show advice that the
sediments of the allochthonous units were deposited at the northern
margin of Gondwana. For example, comparisons in the trilobite fauna
between the Suchomsty and Acanthopyge limestones in Bohemia
(Chlupáč, 1983) and the Greifenstein limestone in the Steinhorn nappe
show a close affinity of the latter to the Barrandian (Flick, 1999).
The diversity of provenance and fossil assemblages in the autochthonous to par-autochthonous units and those of the allochthonous units
require a distinct spatial separation of the source areas during the
Devonian and the lower part of the Lower Carboniferous. Consequently,
this period was dominated by oceans separating Laurussia from
Armorica as well as the latter one from Gondwana (Fig. 7). Based on
the results presented in this paper we conclude that subduction of a
wide Rheic Ocean lasted from Late Silurian until Visean times.
Prerequisite for the subduction of such an oceanic realm are different active margins in the north and in the south, as well as intra-
1496
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
30° N
Rhenohercynian
Ocean
nee
Gr land
ern
rth pe
o
N ro
Eu
Rheic Ocean
n
cea
cO
hei
rt
No
R
Eq
ua
r
me
hA
Saxothuringia
Bohemia
to
ic
a
Palaeotethys
r
30°
S
ri
Af
er i ca
t h Am
Sou
ca
South Pole
Fig. 7. Plate-tectonic reconstruction for the Late Devonian to Early Carboniferous. Red star: Position of the autochthonous Rheinisches Schiefergebirge. Blue star: Source area of the nappe
units in the south east of the Rheinisches Schiefergebirge.
oceanic arcs such as the Frankenstein complex of the Bergsträsser
Odenwald (Fig. 8). However, remnants of such structures are rare or
must be detected indirectly. Thus, subduction at intra-oceanic arcs and
subduction erosion must have played a decisive role (Nance et al., 2012).
Evidence for a wide Rheic Ocean and an active margin to the north
that was later associated with a backarc setting has been postulated
in earlier publications (Floyd, 1982; Flick and Nesbor, 1988; Oczlon,
1994; Smith, 1996; von Raumer and Stampfli, 2008; Zeh and Gerdes,
2010). This wide ocean was subducted through the Devonian and the
early part of the Early Carboniferous along a northward dipping subduction zone below Laurussia and in a southern direction below the
Saxothuringian terrane (Fig. 8). Zircon analysis confirms the existence
of magmatic arcs (especially in the south) until the late Visean and
less prominently in the north until the Tournaisian (Fig. 5). Our model
contradicts the interpretation of sedimentary deposition occurring on
a passive margin along the northern border of a relatively small ocean
proposed by Franke and Oncken (1990, 1995), Oncken et al. (1999),
Franke (2000) and Doublier et al. (2012).
The most important argument suggesting a broad Rheic Ocean is
provided by the provenance data of the allochthonous units. These correspond to the spectra of zircon ages derived from the Saxothuringian
zone, which proves a source area located on Armorica in contradiction
to the interpretation of Huckriede et al. (2004) and Doublier et al.
(2012). Differences in provenance data reflecting Baltica and Armorica
sources require wide separation between the two source areas during
the Devonian and the lower part of the Early Carboniferous. Moreover,
all samples of Visean age exhibit an Armorican pattern of zircon ages —
i.e., not only the samples from the allochthonous units, but also those
from the autochthonous and par-autochthonous units as well. The
reason for this reflects is the geotectonic situation during the late
Early Carboniferous. At that time the Rheic Ocean was closed, and
sedimentation into the Rhenohercynian zone was generally supplied
from the south. Hence, the distinction of different source areas for the
graywackes is no longer possible. Our data support the existence of
the Rheic Ocean until the Early Carboniferous, which contradicts the
results published by, for example, Kroner et al. (2007) and Linnemann
et al. (2010).
Further support for the plate tectonic model presented is provided by
the volcanic activity that accompanied the subduction of the Rheic
Ocean. This activity is documented in numerous ash layers intercalated
within the sedimentary rock pile of the Rhenohercynian (Heyckendorf,
1985; Kubanek and Zimmerle, 1986; Kirnbauer, 1991; van Amerom
et al., 2001). The magma composition is characterized by predominant
dacitic/rhyodacitic and minor rhyolitic melts (see Fig. 3). Such silica
rich volcanism is typical of thick continental crusts above the subducted
slab. Furthermore, the uniform composition of the magmas generated
over the whole time span (about 80 Ma) requires a constant process of
melt generation that is only provided by the subduction of a large ocean.
The geodynamic development of the southern margin of Laurussia
is documented by the evolution of the grain sizes of the pyroclastic
deposits. During the Lower Devonian typical fall out and ignimbrite
deposits occur, often composed of coarse-grained particles. Y-shaped
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
A
1497
Late Devonian
Baltica
North
Avalonia
Ulmbach Formation
„Frasnian Gießen graywacke“
South
Avalonia
Rhenohercynian
Ocean
Rheic Ocean
Laurussia
Armorica
roll back
B
Latest Devonian to Early Carboniferous
Baltica
North
Avalonia
South
Avalonia
Rhenohercynian
Ocean
Ka
m
m
qu
ar
tz
ite
Rheic
Frankenstein Elnhausen Formation
Complex
„Kehna graywacke“
Early Carboniferous
„Gießen graywacke“
Ocean
Laurussia
Armorica
slab break up
Fig. 8. Simplified geotectonic model for the (A) Late Devonian and (B) Latest Devonian to Early Carboniferous.
shards and highly vesicular pyroclasts are common. During that time
the dacitic/rhyodacitic and rhyolitic volcanism was active in the
Rheinisches Schiefergebirge or nearby. Thus, the subduction zone was
situated directly at the southern margin of Laurussia. In the Middle
Devonian, extremely fine grained ash fall was deposited, which marks
the opening of the Rhenohercynian Ocean, triggered by a roll back of
the subducted slab (von Raumer and Stampfli, 2008). As a result
Avalonia split up into a northern and a southern segment — North and
South Avalonia (Fig. 8). Because of the increasing distance, the grain
size of the pyroclastic fall out deposits decreases. These are witnesses
of highly explosive Plinian and Ultraplinian eruptions, which can be
traced over wide distances. The grain size of the Visean volcanic ashes
increased again as a result of the closing of the Rhenohercynian Ocean
and the approach of the subduction zone dipping beneath South
Avalonia.
The existence of South Avalonia is demonstrated by the detrital zircons in the Ruhla Crystalline complex of the Mid-German Crystalline
zone (MGCZ), which suggest two different source areas (Zeh and
Gerdes, 2010). The northern part of the Ruhla Crystalline complex exhibits a provenance in Avalonia, whereas the southern part has a provenance in Armorica. Consequently, the main suture between Avalonia
and Armorica must lie within the MGCZ and not at the southern Taunus
margin as previously believed (e.g., Franke and Oncken, 1995; Oncken
et al., 2000). Besides the northern part of the Ruhla Crystalline, the
Spessart and the Böllstein Odenwald can also be interpreted as belonging to South Avalonia. This assumption is supported by subductionrelated magmatism at the Silurian/Devonian boundary (Dombrowski
et al., 1995; Reischmann et al., 2001). This magmatism can be correlated
with Silurian volcanic activity at the southern margin of the Rheinisches
Schiefergebirge (Sommermann et al., 1992, 1994; Meisl, 1995).
Doublier et al. (2012) suggest that the volcanic belt of the Lahn–
Dill area is separated from the volcanic evolution of the Gießen-Lizard
Ocean (= Rhenohercynian Ocean) by a large non-volcanic belt
comprising the Bicken–Ense unit, Hörre nappe, Lindener Mark,
Hessische Schieferserie and the Taunus. However, new mappings
shows that the volcanic rocks of the Lahn- and Dill–Eder synclines
were not only limited to their present distribution area. Rather, their
subvolcanic dikes and sills were distributed throughout the entire
autochthon of the Rheinisches Schiefergebirge. Likewise, fragments of
volcaniclastic rocks of the Givetian–Frasnian phase are found in volcanic
breccias of a phreatomagmatic diatreme of Tertiary age in the Taunus
anticline (Requadt and Weidenfeller, 2007).
At the southern rim of the Rheic Ocean subduction processes beneath
Armorica are demonstrated by typical subduction-related plutonic rocks
in the Mid-German Crystalline Rise (MGCZ), the Saxothuringian- and
the Moldanubian zone. These plutons often show an I-type signature
and a calc alkaline composition (e.g., Altherr et al., 1999; Stein, 2001).
Their geochronological ages range from the Late Devonian/Early
Carboniferous boundary to the earliest part of the Late Carboniferous.
In addition to these active continental margins, intraoceanic island
arcs also occur, such as the Frankenstein complex of the Bergsträsser
Odenwald with a Late Devonian/Early Carboniferous age (Kirsch et al.,
1988). The final stage of the Variscan orogeny with thrust faulting,
nappe transport, and strike–slip faults leads to the present complex
geotectonic picture.
6. Conclusions
• 33 samples obtained from Devonian and Carboniferous autochthonous to par-autochthonous and allochthonous units of the eastern
Rheinisches Schiefergebirge have been investigated for U–Pb detrital
zircon ages. A representative selection of 15 samples is presented in
this paper in the context of their stratigraphic, sedimentary, magmatic
and plate tectonic implications.
• Based on the distribution of zircon age populations, the autochthonous
to par-autochthonous sediments exhibits a typical Baltica provenance
with weak Avalonian signals, whereas those of the allochthonous
units were derived from a Gondwanan source.
• Sediments of the same age of autochthonous to parautochthonous and
allochthonous units show profound differences in their composition
1498
•
•
•
•
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
and facies development. The allochthonous units such as the Lohra,
Steinhorn, and Hörre nappes are characterized by the absence of
volcanic rocks. At the base of the Gießen nappe metabasalts with
MORB-type affinities occur as tectonic slices.
Subduction-related volcanism lasted from the Early Devonian to
the Early Carboniferous with a source area outside the Rheinisches
Schiefergebirge. Intraplate volcanism was confined to the autochthonous to par-autochthonous units of the Rheinisches Schiefergebirge.
The existence of different magmatic arcs is mirrored in the frequency
distribution of zircon grains. Based on both the provenance analysis
and geochemical results, the evolution of these suggests subduction
to the north as well as to the south.
The plate tectonic model presented suggests the existence of a wide
Rheic Ocean and an active continental margin to the north, which
resulted in the opening of the Rhenohercynian Ocean. This was associated with extension and subsidence of the continental crust of
North Avalonia (=southern shelf area of Laurussia), which was
accompanied by an intensive intraplate volcanism.
In contrast to earlier publications (e.g., Kroner et al., 2007; Linnemann
et al., 2010) the data presented here suggest the existence of the Rheic
Ocean until the Early Carboniferous.
Supplementary data to this article can be found online at http://
dx.doi.org/10.1016/j.gr.2013.05.018.
Acknowledgments
We are much indebted to Dr. Peter Bender (Marburg) whose biostratigraphic work over the past decades and whose knowledge of the
regional geology provided the background for this study. This paper is
therefore dedicated to him. We thank Gerhard Wörner (University of
Göttingen) who performed the geochemical analysis of the Kellerwald
samples. We also acknowledge the support by Ms. Schaffner and
Mr. Hoeßle (both HLUG) for preparing the geological maps. We gratefully acknowledge constructive reviews by Sonia Sánchez Martínez
(Madrid) and Daniel Pastor Galan (Utrecht). Furthermore, we are
thankful for the critical comments by Heiner Flick (Marktoberdorf).
All comments helped to improve the paper. We also thank Johnny
Waters (Boone, USA) for his careful language check. Finally, we thank
the Editor-in-Chief and the Associate Editor who provided the possibility to publish this paper.
References
Ahrendt, H., Hunziker, J.C., Weber, K., 1978. K/Ar-Altersbestimmungen an schwachmetamorphen Gesteinen des Rheinischen Schiefergebirges. Zeitschrift der
Deutschen Geologischen Gesellschaft 129, 229–247.
Ahrendt, H., Clauer, N., Hunziker, J.C., Weber, K., 1983. Migration of folding and metamorphism in the Rheinisches Schiefergebirge deduced from K–Ar and Rb–Sr age
determinations. In: Martin, H., Eder, F.W. (Eds.), Intracontinental Fold Belts.
Springer, Berlin, Heidelberg, New York, pp. 323–338.
Altenberger, U., Besch, T., Mocek, B., Zaipeng, Y., Yong, S., 1990. Geochemie und
Geodynamik des Böllsteiner Odenwaldes. Mainzer Geowissenschaftliche Mitteilungen
19, 183–200.
Altherr, R., Henes-Klaiber, U., Hegner, E., Satir, M., Langer, G., 1999. Plutonism in the
Variscan Odenwald (Germany): from subduction to collision. International Journal
of Earth Sciences 88, 422–443.
Behrmann, J., Drozdzewski, K., Heinrichs, T., Huch, M., Meyer, W., Oncken, O., 1991.
Crustal-scale balanced cross sections through the Variscan fold belt, Germany:
the central EGT-segment. Tectonophysics 196, 1–21.
Bender, P., 1978. Die Entwicklung der Hörre-Zone im Devon und Unterkarbon.
Zeitschrift der Deutschen Geologischen Gesellschaft 129, 131–140.
Bender, P., 1989. Die Hörre und ihre Stellung im östlichen Rheinischen Schiefergebirge.
Jahresberichte und Mitteilungen des Oberrheinischen Geologischen Vereins, Neue
Folge 71, 347–356.
Bender, P., 1997. Paläozoikum der Bickener- und Wildestein-Schuppe. In: Bender, P.,
Lippert, H.-J., Nesbor, H.D. Geologischen Karte von Hessen 1:25.000, Blatt 5216
Oberscheld, 2. Auflage, Erläuterungen. Hessisches Landesamt für Bodenforschung,
Wiesbaden, 91–111.
Bender, P., 2006. Geologische Karte der Hörre (zwischen Dill und Lahn) und der
Frankenbacher Schuppenzone, 1:40.000. Hessisches Landesamt für Umwelt und
Geologie, Wiesbaden.
Bender, P., 2008. Lahn- und Dill-Mulde. In: Deutsche Stratigraphische Kommission
(Ed.), Stratigraphie von Deutschland VIII, Devon. Schriftenreihe der Deutschen
Gesellschaft für Geowissenschaften, 52, pp. 221–246.
Bender, P., Blumenstengel, H., 2003. Ostracoden aus der Weitershausen Formation
(Oberdevon, Hörre, Rheinisches Schiefergebirge). Geologisches Jahrbuch Hessen
131, 191–201.
Bender, P., Homrighausen, R., 1979. Die Hörre-Zone, eine Neudefinition auf
lithostratigraphischer Grundlage. Geologica et Palaeontologica 13, 257–260.
Bender, P., Königshof, P., 1994. Regional maturation patterns of the Devonian strata in
the eastern Rheinisches Schiefergebirge (Lahn–Dill area) based on conodont colour alteration (CAI). Courier Forschungsinstitut Senckenberg 168, 335–345.
Bender, P., Stoppel, D., 2006. Der Ost- und Südostrand des Rheinischen Schiefergebirges:
Lahn-Dill-Gebiet, Kellerwald, Taunus. In: Deutsche Stratigraphische Kommission
(Ed.), Stratigraphie von Deutschland VI, Unterkarbon (Mississippium). Schriftenreihe
der Deutschen Gesellschaft für Geowissenschaften, 41, pp. 358–378.
Bender, P., Lippert, H.-J., Nesbor, H.D., 1997. Geologischen Karte von Hessen 1:25.000,
Blatt 5216 Oberscheld, 2. Auflage, Erläuterungen. Hessisches Landesamt für
Bodenforschung, Wiesbaden, 1–421.
Birkelbach, M., Dörr, W., Franke, W., Michel, H., Stibane, F., Weck, R., 1988. Die
geologische Entwicklung der östlichen Lahnmulde (Exkursion C am 7. April 1988).
Jahresberichte und Mitteilungen des Oberrheinischen Geologischen Vereins, Neue
Folge 70, 43–74.
Braun, R., Oetken, S., Königshof, P., Kornder, L., Wehrmann, A., 1994. Development and
biofacies of reef-influenced carbonates (Central Lahn Syncline, Rheinisches
Schiefergebirge). Courier Forschungsinstitut Senckenberg 169, 351–386.
Brinkmann, R., 1948. Die Mitteldeutsche Schwelle. Geologische Rundschau 36, 56–66.
Buggisch, W., Flügel, E., 1992. Mittel- bis oberdevonische Karbonate auf Blatt Weilburg
(Rheinisches Schiefergebirge) und in Randgebieten: Initialstadien der Riffentwicklung
auf Vulkanschwellen. Geologisches Jahrbuch Hessen 120, 77–97.
Chlupáč, I., 1983. Trilobite assemblages in the Devonian of the Barrandian area and
their relations to palaeoenvironments. Geologica et Palaeontologica 17, 45–73.
Daber, R., 1959. Die Mittel-Visé-Flora der Tiefbohrung von Doberlug-Kirchhain.
Geologie 8 (Beihefte 26), 1–83.
Dehmer, J., Hentschel, G., Horn, M., Kubanek, F., Nöltner, T., Rieken, R., Wolf, M.,
Zimmerle, W., 1989. Die vulkanisch-kieselige Gesteinsassoziation am Beispiel der
unterkarbonischen Kieselschiefer am Ostrand des Rheinischen Schiefergebirges.
Geologie–Petrologie–Geochemie. Geologisches Jahrbuch Hessen 117, 79–138.
Dombrowski, A., Henjes-Kunst, F., Höhndorf, A., Kröner, A., Okrusch, M., Richter, P., 1995.
Orthogneisses in the Spessart Crystalline Complex, Northwest Bavaria: Silurian
granitoid magmatism at an active continental margin. Geologische Rundschau 84,
399–411.
Dörr, W., 1986. Stratigraphie, Stoffbestand und Fazies der Gießener Grauwacke
(östliches Rheinisches Schiefergebirge). (Dissertation) Universität Gießen.
Dörr, W., 1990. Stratigraphie, Stoffbestand und Fazies der Gießener Grauwacke
(östliches Rheinisches Schiefergebirge). Geologische Abhandlungen Hessen 91, 1–94.
Dörr, W., Zulauf, G., 2010. Elevator tectonics and orogenetic collapse of a Tibetan-style
plateau in the European Variscides: the role of the Bohemian shear zone. International Journal of Earth Sciences 99, 299–325.
Doublier, M.P., Potel, S., Franke, W., Roache, T., 2012. Very low-grade metamorphism of
Rheno-Hercynian allochthons (Variscides, Germany): facts and tectonic consequences. International Journal of Earth Sciences 101, 1229–1252.
Drost, K., Gerdes, A., Jeffries, T., Linnemann, U., Storey, C., 2010. Provenance of
Neoproterozoic and early Paleozoic siliciclastic rocks of the Teplá–Barrandian
unit (Bohemian Massif): evidence from U–Pb detrital zircon ages. Gondwana Research 19, 213–231.
Eder, W., Engel, W., Uffenorde, H., 1969. Stratigraphische und fazielle Gliederung
des Quarzit-Zuges im Kellerwald (Mitteldevon bis Unterkarbon; Rheinisches
Schiefergebirge). Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen
133, 223–244.
Eder, W., Engel, W., Franke, W., 1977. Gießener Grauwacke (Aufschlüsse 15 bis 17). In:
Bender, P., Eder, W., Engel, W., Franke, W., Langenstrassen, F., Walliser, O.H.,
Witten, W. (Eds.), Exkursion A. Paläogeographische Entwicklung des östlichen
Rheinischen Schiefergebirges, demonstriert an einem Querschnitt. Exkursionführer
Geotagung '77 Göttingen, I, pp. 49–53.
Engel, W., Franke, W., 1983. Flysch sedimentation: its relationship to tectonism in the
European Variscides. In: Martin, H., Eder, F.W. (Eds.), Intracontinental Fold Belts.
Springer, Berlin, Heidelberg, New York, pp. 289–322.
Engel, W., Franke, W., Grote, C., Weber, K., Ahrendt, H., Eder, F.W., 1983. Nappe tectonics in
the southeastern part of the Rheinisches Schiefergebirge. In: Martin, H., Eder, F.W.
(Eds.), Intracontinental Fold Belts. Springer, Berlin, Heidelberg, New York, pp. 267–287.
Entenmann, W., 1991. Feinstratigraphie und Sedimentologie der Elnhausen-Schichten
(Unterkarbon) der Hörre-Zone. Geologica et Palaeontologica 25, 79–85.
Flick, U., 1999. Die Trilobiten von Greifenstein/Hörre - Ableger einer böhmischen Fauna im
Rheinischen Schiefergebirge. Jahrbücher des Nassauischen Vereins für Naturkunde
120, 23–50.
Flick, H., Nesbor, H.-D., 1988. Der Vulkanismus in der Lahnmulde. Jahresberichte und
Mitteilungen des Oberrheinischen Geologischen Vereins, Neue Folge 70, 411–475.
Floyd, P.A., 1982. Chemical variation in Hercynian basalts relative to plate tectonics.
Journal of the Geological Society of London 139, 505–520.
Franke, W., 1995. Rhenohercynian foldbelt: autochthon and nonmetamorphic nappe
units — stratigraphy. In: Dallmeyer, D., Franke, W., Weber, K. (Eds.), Pre-Permian
Geology of Central and Eastern Europe. Springer, Berlin, pp. 33–39.
Franke, W., 2000. The mid-European segment of the Variscides: tectonostratigraphic
units, terrane boundaries and plate tectonic evolution. In: Franke, W., Haak, V.,
Oncken, O., Tanner, D. (Eds.), Orogenic Processes: Quantification and Modelling
in the Variscan Belt. Geological Society London, Special Publication, 179, pp. 35–61.
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
Franke, W., Oncken, O., 1990. Geodynamic evolution of the northcentral Variscides — a
comic strip. In: Freeman, R., Giese, P., Mueller, S. (Eds.), The European Geotraverse.
European Science Foundation, pp. 187–194.
Franke, W., Oncken, O., 1995. Zur prädevonischen Geschichte des Rhenohercynischen
Beckens. Nova Acta Leopoldina, Neue Folge 71, 53–72.
Franke, W., Bortfeld, R.K., Brix, M., Drozdzewski, G., Dürbaum, H.J., Giese, P., Janoth, W.,
Jödicke, H., Reichert, Ch., Scherp, A., Schmoll, J., Thomas, R., Thünker, M., Weber, K.,
Wiesner, M.G., Wong, H.K., 1990. Crustral structure of the Rhenish Massif: results
of deep seismic reflection lines DEKORP 2-North and 2-North-Q. Geologische
Rundschau 79, 523–566.
Frei, D., Gerdes, A., 2009. Precise and accurate in situ U–Pb dating of zircon with high
sample throughput by automated LA–SF–ICP–MS. Chemical Geology 261, 261–270.
Gerdes, A., Zeh, A., 2006. Combined U–Pb and Hf isotope LA–(MC–) ICP–MS analysis of
detrital zircons: comparison with SHRIMP and new constraints for the provenance
and age of an Armorican metasediment in Central Germany. Earth and Planetary
Science Letters 249, 47–61.
Groos-Uffenorde, H., Lethiers, F., Blumenstengel, H., 2000. Ostracodes and Devonian
stratigraphy. Courier Forschungsinstitut Senckenberg 220, 99–111.
Grösser, J., Dörr, W., 1986. MOR-Type-Basalte im östlichen Rheinischen Schiefergebirge.
Neues Jahrbuch für Geologie und Paläontologie, Monatshefte 12, 705–712.
Haverkamp, J., von Hoegen, J., Kramm, U., Walter, R., 1992. Application of U–Pbsystems from detrital zircons for palaeogeographic reconstructions — a case
study from the Rhenohercynian. Geodinamica Acta 5, 69–82.
Henes-Klaiber, U., 1992. Zur Geochemie der variskischen des Bergsträßer Odenwaldes.
(Ph.D. thesis) Universität Karlsruhe.
Henningsen, D., 1962. Untersuchungen über Stoffbestand und Paläogeographie der
Gießener Grauwacke. Geologische Rundschau 51, 600–626.
Herbig, H.-G., Bender, P., 1992. A eustatically driven calciturbidite sequence from the
Dinantian II of the eastern Rheinisches Schiefergebirge. Facies 27, 245–262.
Heyckendorf, K., 1985. Die unterdevonischen Lenne-Vulkanite im nordöstlichen
Rheinischen Schiefergebirge: Beiträge zur Stratigraphie, Paläogeographie, Petrographie
und Geochemie. Mitteilungen aus dem Geologisch-Paläontologischen Institut der
Universität Hamburg 68, 1–52.
Hofmann, A.W., 1988. Chemical differentiation of the earth: the relationship between
mantle, continental crust, and oceanic crust. Earth and Planetary Science Letters 90,
297–314.
Homrighausen, R., 1979. Petrographische Untersuchungen an sandigen Gesteinen der
Hörre-Zone (Rheinisches Schiefergebirge, Oberdevon-Unterkarbon). Geologische
Abhandlungen Hessen 79, 1–84.
Huckriede, H., Wemmer, W., Ahrendt, H., 2004. Palaeogeography and tectonic structure
of allochthonous units in the German part of the Rheno-Hercynian Belt (Central
European Variscides). International Journal of Earth Sciences 93, 414–431.
Jäger, H., Gursky, H.-J., 2000. Alter, genese und Paläogeographie der Kammquarzit Formation (Vise) im Rhenoherzynikum – neue Daten und neue Deutungen. Zeitschrift
der Deutschen Geologischen Gesellschaft 151, 415–439.
Jahnke, H., 1971. Fauna und Alter der Erbslochgrauwacke (Brachiopoden und
Trilobiten, Unterdevon. Rheinisches Schiefergebirge und Harz). Göttinger Arbeiten
zur Geologie und Paläontologie 9, 1–105.
Jahnke, H., Paul, J., 1968. Das Alter der Grauwacken im südlichen Kellerwald (Oberdevon
und Unterkarbon, Rheinisches Schiefergebirge). Notizblatt des Hessischen
Landesamtes für Bodenforschung zu Wiesbaden 96, 68–84.
Jeffries, T.E., Fernandez-Suarez, J., Corfu, F., Alonso, G.G., 2003. Advances in U–Pb geochronology using a frequency quintupled Nd:YAG based laser ablation system
(lambda = 213 nm) and quadrupole based ICP–MS. Journal of Analytical Atomic
Spectrometry 18, 847–855.
Kerp, H., Kampe, A., Schultka, S., van Amerom, H.W.J., 2006. Makrofloren. In: Deutsche
Stratigraphische Kommission (Ed.), Stratigraphie von Deutschland VI, Unterkarbon
(Mississippium). Schriftenreihe der Deutschen Gesellschaft für Geowissenschaften,
41, pp. 271–293.
Kirnbauer, T., 1991. Geologie, Petrographie und Geochemie der Pyroklastika des Unteren
Ems/Unterdevon (Porphyroide) im südlichen Rheinischen Schiefergebirge. Geologische
Abhandlungen Hessen 92, 1–228.
Kirsch, H., Kober, B., Lippolt, H.J., 1988. Age of intrusion and rapid cooling of the
Frankensteingabbro (Odenwald, SW-Germany) evidenced by Ar 40/Ar 39 and
single-zircon Pb 207/Pb measurements. Geologische Rundschau 77, 693–711.
Klügel, T., 1997. Geometrie und Kinematik einer variszischen Plattengrenze – der Südrand
des Rhenoherzynikums im Taunus. Geologische Abhandlungen Hessen 101, 1–215.
Knaus, J.M., 1994. Triphyllopteris collombiana: a clarification of the generic concept
based on rediscovered specimens from Kossberg bei Plauen, Germany, and a
reassignment of the North American species of Triphyllopteris to Genselia gen.
nov. International Journal of Plant Sciences 155, 97–116.
Koch, C., 1858. Paläozoische Schichten und Grünsteine in den HerzoglichNassauischen Ämtern Dillenburg und Herborn, unter Berücksichtigung allgemeiner
Lagerungsverhältnisse in angrenzenden Ländertheilen. Jahrbücher des Nassauischen
Vereins für Naturkunde 13, 85–329.
Kockel, C., 1958. Schiefergebirge und Hessische Senke um Marburg/Lahn. Sammlung
geologischer Führer, 37. Bornträger, Berlin.
Königshof, P., Gewehr, B., Kornder, L., Wehrmann, A., Braun, R., Zankl, H., 1991.
Stromatoporen-Morphotypen aus einem zentralen Riffbereich (Mitteldevon) in
der südwestlichen Lahnmulde. Geologica et Palaeontologica 25, 19–35.
Königshof, P., Flick, H., Nesbor, H.-D., Schäfer, A., Stets, J., 2008. Rheno-Hercynian Zone. In:
Königshof, P., Linnemann, U. (Eds.), The Rheno-Hercynian, Mid-German CrystallineZone and Saxo-Thuringian Zones (Central European Variscides). 20th International
Senckenberg-Conference and 2nd Geinitz-Conference, From Gondwana to Laurussia
to Pangaea, Dynamics of Oceans and Supercontinents, Final Meeting of IGCP 497 and
499. Excursion Guide. Senckenberg, Frankfurt am Main, Dresden, pp. 11–78.
1499
Königshof, P., Nesbor, H.-D., Flick, H., 2010. Volcanism and reef development in
the Devonian: a case study from the Lahn syncline, Rheinisches Schiefergebirge
(Germany). Gondwana Research 17, 264–280.
Kossmat, F., 1927. Gliederung des varistischen Gebirgsbaus. Abhandlungen des
Sächsischen Geologischen Landesamts 1, 1–39.
Kroner, U., Hahn, T., Romer, R.L., Linnemann, U., 2007. The Variscan orogeny in the
Saxo-Thuringian zone — heterogenous overprint of Cadomian/Palaeozoic periGondwana crust. In: Linnemann, U., Nance, R.D., Kraft, P., Zulauf, G. (Eds.), The Evolution of the Rheic Ocean: From Avalonian–Cadomian Active Margin to Alleghenian–
Variscan collision. Geological Society of America, Special Paper, 423, pp. 153–172.
Kubanek, F., Zimmerle, W., 1986. Tuffe und kieselige Tonschiefer aus dem tieferen
Unterkarbon der Bohrung Adlersberg (West-Harz). Geologisches Jahrbuch D 78,
207–268.
Kupfahl, H.G., 1953. Untersuchungen im Gotlandium und Unterdevon des Kellerwaldes
und bei Marburg. Notizblatt des Hessischen Landesamtes für Bodenforschung zu
Wiesbaden 81, 96–128.
Linnemann, U., McNaughton, N.J., Romer, R.L., Gehmlich, M., Drost, K., Tonk, C., 2004.
West African provenance for Saxo-Thuringia (Bohemian Massif): did Armorica
ever leave pre-Pangean Gondwana? — U–Pb-SHRIMP zircon evidence and the
Nd-isotopic record. International Journal of Earth Sciences 93, 683–705.
Linnemann, U., Gerdes, A., Drost, K., Buschmann, B., 2007. The continuum between
Cadomian Orogenesis and opening of the Rheic Ocean: constraints from LA–ICP–
MS U–Pb zircon dating and analysis of plate-tectonic setting (Saxo-Thuringian
zone, NE Bohemian massif, Germany). In: Linnemann, U., Nance, R.D., Kraft, P.,
Zulauf, G. (Eds.), The Evolution of the Rheic Ocean: From Avalonian–Cadomian
Active Margin to Alleghenian–Variscan Collision. Geological Society of America,
Special Paper, 423, pp. 61–96.
Linnemann, U., Pereira, F., Jeffries, T.E., Drost, K., Gerdes, A., 2008. The Cadomian orogeny and the opening of the Rheic Ocean: the diachrony of geotectonic processes
constraints by LA–ICP–MS U–Pb zircon dating (Ossa-Morna and Saxo-Thuringian
Zones, Iberian and Bohemian Massifs). Tectonophysics 461, 21–43.
Linnemann, U., Hofmann, M., Romer, R.L., Gerdes, A., 2010. Transitional stages between
the Cadomian and Variscan Orogenies: basin development and tectonomagmatic
evolution of the southern margin of the Rheic Ocean in the Saxo-Thuringian
Zone (North Gondwana shelf). In: Linnemann, U., Romer, R.L. (Eds.), Pre-Mesozoic
Geology of Saxo-Thuringia — From the Cadomian Active Margin to the Variscan
Orogeny. Schweizerbart Science Publishers, Stuttgart, pp. 59–98.
Linnemann, U., Ouzegane, K., Drareni, A., Hofmann, M., Becker, S., Gärtner, A., Sagawe,
A., 2011. Sands of West Gondwana: an archive of secular magmatism and plate interactions — a case study from the Cambro-Ordovician section of the Tassili Ouan
Ahaggar (Algerian Sahara) using U–Pb-LA–ICP–MS detrital zircon ages. Lithos
123, 188–203.
Linnemann, U., Herbosch, A., Liégeois, J.-P., Pin, C., Gärtner, A., Hofmann, M., 2012. The
Cambrian to Devonian Odyssey of the Brabant Massif within Avalonia: a review with
new zircon ages, geochemistry, Sm–Nd isotopes, stratigraphy and palaeogeography.
Earth-Science Reviews 112, 126–154.
Ludwig, K.R., 2001. Users Manual for Isoplot/Ex rev. 2.49. Berkeley Geochronology
Center Special Publication, 1a 1–56.
Matte, Ph., 1986. Tectonics and plate tectonics model for the Variscan Belt of Europe.
Tectonophysics 126, 329–374.
Meischner, D., 1966. Der geologische Bau des Kellerwaldes. Habilitationsschrift,
Universität Göttingen.
Meischner, D., 1991. Kleine Geologie des Kellerwaldes (Exkursion F). Jahresberichte und
Mitteilungen des Oberrheinischen Geologischen Vereins, Neue Folge 73, 115–142.
Meisl, S., 1995. Igneous activity. In: Dallmeyer, R.D., Franke, W., Weber, K. (Eds.), PrePermian Geology of Central and Eastern Europe. Springer, Berlin, pp. 118–131.
Mittmeyer, H.G., 2008. Unterdevon der Mittelrheinischen und Eifeler Typ-Gebiete (Teile von
Eifel, Westerwald,Hunsrück und Taunus). In: Deutsche Stratigraphische Kommission
(Ed.), Stratigraphie von Deutschland VIII, Devon. Schriftenreihe der Deutschen Gesellschaft für Geowissenschaften, 52, pp. 139–203.
Mosseichik, Yu.V., 2007. Correlation of Visean floral successions from the Equatorial
Belt. Doklady Earth Sciences 415, 701–704.
Nance, R.D., Gutiérrez-Alonso, G., Keppie, J.D., Linnemann, U., Brendan Murphy, J.,
Quesada, C., Strachan, R.A., Woodcock, N., 2010. Evolution of the Rheic Ocean.
Gondwana Research 17, 194–222.
Nance, R.D., Gutierrez-Alonso, G., Keppie, D., Linnemann, U., Murphy, J.B., Quesada, C.,
Strachan, R.A., Woodcock, N.H., 2012. A brief history of the Rheic Ocean. Geoscience Frontiers 3 (2), 125–135.
Nesbor, H.-D., 1997. Petrographie der vulkanischen Gesteine. in: Bender, P., Lippert, H.-J.,
Nesbor, H.D. Geologischen Karte von Hessen 1:25,000, Blatt 5216 Oberscheld, 2.
Auflage, Erläuterungen. Hessisches Landesamt für Bodenforschung, Wiesbaden,
159–207.
Nesbor, H.-D., 2004. Paläozoischer Intraplattenvulkanismus im östlichen Rheinischen
Schiefergebirge – Magmenentwicklung und zeitlicher Ablauf. Geologisches Jahrbuch
Hessen 131, 145–182.
Nesbor, H.-D., 2007. Paläozoischer Vulkanismus im Lahn-Dill-Gebiet – südliches
Rheinisches Schiefergebirge (Exkursion E am 12 April 2007). Jahresberichte und
Mitteilungen des Oberrheinischen Geologischen Vereins, Neue Folge 89, 193–216.
Nesbor, H.-D., 2008. Bimodaler Vulkanismus im Devon des Rheinischen Schiefergebirges.
In: Deutsche Stratigraphische Kommission (Ed.), Stratigraphie von Deutschland
VIII, Devon. Schriftenreihe der Deutschen Gesellschaft für Geowissenschaften, 52,
pp. 247–251.
Nesbor, H.-D., Buggisch, W., Flick, H., Horn, M., Lippert, H.-J., 1993. Vulkanismus im
Devon des Rhenoherzynikums. Fazielle und paläogeographische Entwicklung
vulkanisch geprägter mariner Becken am Beispiel des Lahn-Dill-Gebietes.
Geologisches Jahrbuch Hessen 98, 3–87.
1500
K. Eckelmann et al. / Gondwana Research 25 (2014) 1484–1500
Oczlon, M.S., 1992. Gondwana and Laurussia before and during the Variscan Orogeny in
Europe and related areas. Heidelberger Geowissenschaftliche Abhandlungen 53, 1–56.
Oczlon, M.S., 1994. North Gondwana origin for exotic Variscan rocks in the Rhenohercynian
zone of Germany. Geologische Rundschau 83, 20–31.
Oncken, O., 1997. Transformation of a magmatic arc and an orogenic root during
oblique collision and its consequences for the evolution of the European Variscides
(Mid-German crystalline rise). Geologische Rundschau 86, 2–10.
Oncken, O., Weber, K., 1995. Structure. In: Dallmeyer, R.D., Franke, W., Weber, K. (Eds.),
Pre-Permian Geology of Central and Eastern Europe. Springer, Berlin, pp. 50–58.
Oncken, O., Winterfeld, C., Dittmar, U., 1999. Accretion and inversion of a rifted passive
margin — the Late Palaeozoic Rhenohercynian fold and thrust belt. Tectonics 18, 75–91.
Oncken, O., Plesch, A., Weber, K., Ricken, W., Schrader, S., 2000. Passive margin detachment during arc–continent collision (Central European Variscides). In: Franke, W.,
Haak, V., Oncken, O., Tanner, D. (Eds.), Orogenic Processes: Quantification and
Modelling in the Variscan Belt. The Geological Society of London, Special Publication, 179, pp. 199–216.
Pas, D., DaSilva, A.C., Cornet, P., Bultynck, P., Königshof, P., Boulvain, F., 2013. Sedimentary development of a continuous Middle Devonian to Mississippian section from
the fore-reef fringe of the Brilon Reef Complex (Rheinisches Schiefergebirge,
Germany). Facies 57. http://dx.doi.org/10.1007/s10347-012-0351-z.
Pirwitz, K., 1986. Die Sandsteinfächer des höchsten Oberdevons der Dillmulde (ThalenbergFormation) am SE-Rand des Siegerländer Blocks – Stratigraphische, petrographische
und sedimentologische Untersuchungen und deren paläogeographische Deutung.
(Dissertation) Universität Marburg.
Plesch, A., Oncken, O., 1999. Orogenic wedge growth during collision — constraints on
mechanics of a fossil wedge from its kinematic record (Rhenohercynian FTB, Central
Europe). Tectonophysics 309, 117–139.
Plusquellec, Y., Jahnke, H., 1999. Les tabules de L'Erbsloch-grauwacke (Emsien Inferieur
du Kellerwald) et le problem des affinites paleogeographiques de l'allochthone
“Giessen-Harz”. Abhandlungen der Geologischen Bundesanstalt 54, 435–451.
Poschmann, T., Jansen, U., 2003. Lithologie und Fossilführung einiger Profile in den
Siegen-Schichten des Westerwaldes (Unter-Devon, Rheinisches Schiefergebirge).
Senckenbergiana Lethaea 83, 157–183.
Reineck, H.E., Singh, I.B., 1980. Depositional Sedimentary Environments. Springer, Berlin
Heidelberg, New York 1–549.
Reischmann, T., Anthes, G., Jaeckel, P., Altenberger, U., 2001. Age and origin of the
Böllsteiner Odenwald. Mineralogy and Petrology 72, 29–44.
Requadt, H., Weidenfeller, M., 2007. Geologischen Karte von Rheinland-Pfalz 1:25.000,
Blatt 5713 Katzenelnbogen, 2. Auflage, Erläuterungen. Landesamt für Geologie und
Bergbau Rheinland-Pfalz, Mainz, 1–240.
Romer, R.L., Hahne, K., 2010. Life of the Rheic Ocean: scrolling through the shale record.
Gondwana Research 17, 408–421.
Salamon, M., 2003. Grobklastische Beckensedimente (Olisthostrome) des oberen
Mitteldevons im Lahn-Dill-Gebiet – Zeugen einer aktiven Rift-Tektonik. Geologische
Abhandlungen Hessen 111, 1–209.
Salamon, M., Königshof, P., 2010. Middle Devonian olistostromes in the RhenoHercynian (Rheinisches Schiefergebirge) — an indication of back arc rifting on a
passive margin of Laurussia? Gondwana Research 17, 281–291.
Sánchez Martínez, S., Arenas, R., Díaz García, F., Martínez Catalán, J.R., Gómez-Barreiro,
J., Pearce, J., 2007. Careón ophiolite, NW Spain: suprasubduction zone setting for
the youngest Rheic Ocean floor. Geology 5, 53–56.
Schäfer, J., Neuroth, H., Ahrendt, H., Dörr, W., Franke, W., 1997. Accretion and exhumation
at a Variscan active margin, recorded in the Saxothuringian flysch. Geologische
Rundschau 86, 599–611.
Scherp, A., 1983. Unterdevonische Schmelztuffe im rechtsrheinischen Schiefergebirge.
Neues Jahrbuch für Geologie und Paläontologie, Monatshefte 1983, 47–58.
Schriel, W., Stoppel, D., 1958. Acker-Bruchberg und Kellerwald – Stratigraphie und
Tektonik. Zeitschrift der Deutschen Geologischen Gesellschaft 110, 260–292.
Schwan, W., 1991. Geologie des Acker-Bruchberg-Ilsenburg-Zuges (Oberharz) – Derzeitiger
Forschungsstand und Diskussion der Probleme. Zentralblatt für Geologie und
Paläontologie Teil 1 (7), 787–850.
Sircombe, K.N., 2004. AGEDISPLAY: an EXCEL workbook to evaluate and display univariate geochronological data using binned frequency histograms and probability
density distributions. Computers and Geosciences 30, 21–31.
Smith, A.G., 1996. Some aspects of the Phanerozoic paleogeographic evolution of
Europe. Zeitschrift der Deutschen Geologischen Gesellschaft 147, 147–168.
Sommermann, A.E., 1993. Zirkonalter aus dem Granit der Bohrung Saar 1.Berichte der
Deutschen Mineralogischen Gesellschaft, Beihefte zum. European Journal of Mineralogy 5, 145.
Sommermann, A.-E., Meisl, S., Todt, W., 1992. Zirkonalter von 3 verschiedenen
Metavulkaniten aus dem Südtaunus. Geologisches Jahrbuch Hessen 120, 67–76.
Sommermann, A.-E., Anderle, H.-J., Todt, W., 1994. Das Alter des Quarzkeratophyrs der
Krausaue bei Rüdesheim am Rhein (Bl. 6013 Bingen, Rheinisches Schiefergebirge).
Geologisches Jahrbuch Hessen 122, 143–157.
Stacey, J.S., Kramers, J.D., 1975. Approximation of terrestrial lead isotope evolution by a
two-stage model. Earth and Planetary Science Letters 26, 207–221.
Stein, E., 2001. The geology of the Odenwald crystalline complex. Mineralogy and
Petrology 72, 7–28.
Stoppel, D., 1961. Geologie des südlichen Kellerwaldgebirges. Abhandlungen des
Hessischen Landesamtes für Bodenforschung 34, 1–114.
Tait, J.A., Schätz, M., Bachtadse, V., Soffel, H., 2000. Palaeomagnetism and Palaeozoic
paleogeography of Gondwana and European terranes. In: Franke, W., Haak, V.,
Oncken, O., Tanner, D. (Eds.), Orogenic Processes: Quantification and Modelling
in the Variscan Belt. Geological Society London, Special Publication, 179, pp. 21–34.
Torsvik, T.H., Cocks, L.R.M., 2004. Earth geography from 400 to 250 Ma: a palaeomagnetic,
faunal and facies review. Journal of the Geological Society of London 161, 555–572.
Van Amerom, H.W.J., Heggemann, H., Herbig, H.-G., Horn, M., Korn, D., Nesbor, H.-D.,
Schrader, S., 2001. Das Grauwacken-Profil (Ober-Viseum) des Steinbruchs
Dainrode im Kellerwald (NW Hessen). Geologisches Jahrbuch Hessen 129, 5–25.
Von Raumer, J.F., Stampfli, G.M., 2008. The birth of the Rheic Ocean — Early Palaeozoic
subsidence patterns and subsequent tectonic plate scenarios. Tectonophysics 461,
9–20.
Wachendorf, H., 1986. Der Harz – variszischer Bau und geodynamische Entwicklung.
Geologisches Jahrbuch A 91, 3–67.
Wedepohl, K.H., Meyer, K., Muecke, G.K., 1983. Chemical composition and genetic relations of meta-volcanic rocks from the Rhenohercynian Belt of Central Europe. In:
Martin, H., Eder, F.W. (Eds.), Intracontinental Fold Belts. Springer, Berlin, Heidelberg, New York, pp. 231–256.
Wierich, F., 1999. Orogene Prozesse im Spiegel synorogener Sedimente –
Korngefügekundliche Liefergebietsanalyse siliziklastischer Sedimente im Devon des
Rheinischen Schiefergebirges. Marburger Geowissenschaften 1, 1–244.
Wierich, F., Vogt, A.W., 1997. Zur Verbreitung, Biostratigraphie und Petrographie
unterkarbonischer Sandsteine des Hörre-Gommern-Zuges im östlichen Rheinischen
Schiefergebirge. Geologica et Palaeontologica 31, 97–142.
Winchester, J.A., Floyd, P.A., 1977. Geochemical discrimination of different magma series and their differentiation products using immobile elements. Chemical Geology
20, 325–343.
Zeh, A., Gerdes, A., 2010. Baltica- and Gondwana derived sediments in the Mid-German
Crystalline Rise (Central Europe): implications for the closure of the Rheic Ocean.
Gondwana Research 17, 254–263.
Zeh, A., Will, T.M., 2010. The Mid-German crystalline zone. In: Linnemann, U., Romer,
R.L. (Eds.), Pre-Mesozoic Geology of Saxo-Thuringia — From the Cadomian Active
Margin to the Variscan Orogeny. Schweizerbart Science Publishers, Stuttgart,
pp. 195–220.
Ziegler, W., 1957. Das Marburger Gotlandium. Notizblatt des Hessischen Landesamtes
für Bodenforschung zu Wiesbaden 85, 67–74.