background image

GEOLOGICA CARPATHICA

, JUNE 2017, 68, 3, 248 – 268

doi: 10.1515/geoca-2017-0018

www.geologicacarpathica.com

Geochemistry, environmental and provenance study of  

the Middle Miocene Leitha limestones (Central Paratethys)

AHMED ALI

1, 2, 

 and MICHAEL WAGREICH

2

1

 Geology Department, Faculty of Science, Minia University, 61519 Menia, Egypt; 

ahmad.ali@mu.edu.eg

2

 Department of Geodynamics and Sedimentology, Centre for Earth Sciences, University of Vienna, 1090 Vienna, Austria

(Manuscript received July 20, 2016; accepted in revised form March 15, 2017)

Abstract: Mineralogical, major, minor, REE and trace element analyses of rock samples were performed on Middle 

Miocene limestones (Leitha limestones, Badenian) collected from four localities from Austria (Mannersdorf,  Wöllersdorf, 

Kummer and Rosenberg quarries) and the Fertőrákos quarry in Hungary. Impure to pure limestones (i.e. limited by Al

2

O

3

 

contents above or below 0.43 wt. %) were tested to evaluate the applicability of various geochemical proxies and indices 

in regard to provenance and palaeoenvironmental interpretations. Pure and impure limestones from Mannersdorf and 

Wöllersdorf (southern Vienna Basin) show signs of detrital input (REEs = 27.6 ± 9.8 ppm, Ce anomaly = 0.95 ± 0.1 and the 

presence of quartz, muscovite and clay minerals in impure limestones) and diagenetic influence (low contents of, e.g., 

Sr = 221 ± 49 ppm, Na is not detected, Ba = 15.6 ± 8.8 ppm in pure limestones). Thus, in both limestones the reconstruction 

of original sedimentary palaeoenvironments by geochemistry is hampered. The Kummer and Fertőrákos (Eisenstadt– 

Sopron  Basin)  comprise  pure  limestones  (e.g.,  averages  Sr = 571 ± 139  ppm,  Na = 213 ± 56  ppm,  Ba = 21 ± 4  ppm, 

REEs = 16 ± 3 ppm and Ce anomaly = 0.62 ± 0.05 and composed predominantly of calcite) exhibiting negligible diagenesis. 

Deposition under a shallow-water, well oxygenated to intermittent dysoxic marine environment can be reconstructed. 

Pure to impure limestones at Rosenberg–Retznei (Styrian Basin) are affected to some extent by detrital input and 

 volcano-siliciclastic admixture. The Leitha limestones at Rosenberg have the least diagenetic influence among the  studied 

localities (i.e. averages Sr = 1271 ± 261 ppm, Na = 315 ± 195 ppm, Ba = 32 ± 15 ppm, REEs = 9.8 ± 4.2 ppm and  Ce anomaly = 

0.77 ± 0.1 and consist of calcite, minor dolomite and quartz). The siliciclastic sources are characterized by immobile 

 elemental ratios (i.e. La/Sc and Th/Co) which apply not only for the siliciclastics, but also for marls and impure  limestones. 

At Mannersdorf the detrital input source varies between intermediate to silicic igneous rocks, while in Kummer and 

Rosenberg the source is solely silicic igneous rocks. The Chemical Index of Alteration (CIA) is only applicable in the 

shale-contaminated impure limestones. CIA values of the Leitha limestones from Mannersdorf indicate a gradual 

 transition from warm to temperate palaeoclimate within the limestone succession of the Badenian.

Keywords: Central Paratethys, Neogene, Leitha limestones, geochemistry, provenance.

Introduction

Sedimentary geochemistry, especially the distribution of rare 

earth elements (REEs) provides valuable information about 

marine depositional environments and sediment provenance. 

It is used in the assessment of the pathways of terrigenous and 

biogenic fluxes from the source to the sediments (Piper 1974; 

Elderfield & Greaves 1982; Haley et al. 2004; Anderson et al. 

2007; Johannesson et al. 2014; Garbelli et al. 2016). Absolute 

element concentrations, relative abundance, patterns of the 

REEs and Ce and Eu anomalies provide guiding tools and can 

be used as proxies for various environmental parameters, 

widely applied in sedimentological studies (e.g., Elderfield 

1988;  Piepgras  &  Jacobsen  1992;  Nozaki  2001;  Bau  et  al. 

2010). The absolute and relative concentrations of these ele-

ments in sediments help us to recognize, classify and unravel 

(1) the detrital input but also input from hydrothermal systems 

and related alterations; (2) the interaction with the biogeo-

chemical cycle involving removal of elements from water 

bodies by adsorption and oxidation on particle surfaces and 

deeper regeneration within the sedimentary column; and  

(3) the effects of advective transport through rocks (Elderfield 

1988; Kocsis et al. 2010; Azmy et al. 2011). REEs are charac-

terized by largely similar behaviour in sediments, based on the 

chemical similarity as a result of the progressive filling of  

the inner 4f electrons orbital with increasing atomic number 

(de Baar et al. 1985; Nozaki 2001; Kim et al. 2012). 

REEs patterns can be used to distinguish between different 

depositional environments. Based on the model proposed by 

Haley et al. (2004) three pattern types can be distinguished: 

(1) the linear pattern type which shows a constant, but mode-

rate increase in the Post-Archean Australian Shale (PAAS)- 

normalized REEs across the series, from lower to higher 

 atomic numbers. In this pattern type REEs are released mainly 

from the degradation of particulate organic matter (POC). This 

pattern is distinctive for oxygen levels, namely oxic and 

 suboxic environments; (2) HREE-enrichment pattern (higher 

(H) atomic number REEs enriched) in which the POC source 

plays an important role in carrying REEs; (3) MREE bulge 

pattern (enrichment in middle (M) REEs) which is strongly 

related to high Fe-concentrations. This correlation is inter-

preted as a result of the dissolution of a surficial, REE-  

enriched solid Fe-phase which becomes the source for these 

REEs.

background image

249

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

Further sediment geochemical interpretations can be per-

formed using cerium and europium. All lanthanide elements 

are generally positively trivalent charged (i.e. 

+3

), but Ce and 

Eu also exist in 

+4

 and 

+2

 states, respectively. The oxidation of 

Ce

+3

 (loss of electron) and production of Ce

+4

 takes place 

trough reaction with seawater and results in the formation of 

cerium dioxide (CeO

2

) which is highly insoluble in seawater. 

This criterion makes Ce a practical index for interpreting 

 

primary redox conditions in seawater using carbonates 

 (Elderfield  1988;  Bau  1991;  Bau  &  Möller  1992;  Bau  & 

 Dulski 1999; Hongo & Nozaki 2001; Dubinin 2004; Craddock 

et al. 2010; Schmidt et al. 2010). The reduction of Eu

+3 

to Eu

+2 

which has the ability to replace Ca

+2

 in feldspars, especially in 

plagioclase, is favourable in sites with submarine hydrother-

mal activity (Elderfield 1988) representing a secondary signi-

ficant  REEs  source  (e.g.,  Michard  et  al.  1983;  Hinkley  & 

 Tatsumoto 1987; Campbell et al. 1988). There, the REEs are 

scavenged quickly from the hydrothermal fluids, especially 

Eu which is removed faster than its neighbour elements, and 

thus produces a positive Eu anomaly (Olivarez & Owen 1991).

In this study we analysed the mineralogy and geochemistry 

of Neogene limestones (Middle Miocene, Badenian Leitha 

limestones) at several localities in Austria and Hungary. This 

work deals mainly with sediment geochemistry regarding 

 major, minor and REEs of these limestones transitional from 

impure to pure carbonate lithologies (distinguished by Al

2

O

3

 

contents below or above 0.43), to get information about sedi-

ment provenance, hinterland and depositional environments. 

Provenance and environmental sediment geochemistry signals 

are distinguished and evaluated for significance in a relatively 

straightforward carbonate setting of the Central Paratethys Sea 

to test the applicability of various methods in sediment geo-

chemistry on shallow-water carbonates in general.   

Geological settings

The Vienna Basin was formed mainly as a pull-apart basin 

along strike-slip and normal faults due to transtension along 

the junction of the Eastern Alps and Western Carpathians 

(Royden 1985; Wessely 2006). Subsidence took place during 

the late Early to Middle Miocene, and led to faulting along the 

basin margins. Subsidence rates strongly varied during basin 

development with increasing rates during the early Badenian 

in the southern part of the basin (Wessely 1983; Lankreijer et 

al. 1995; Wagreich & Schmid 2002; Lee & Wagreich 2017). 

Islands, shoals, and small carbonate platforms were wide-

spread in the Vienna Basin during the Middle Miocene (Dullo 

1983; Riegl & Piller 2000; Schmid et al. 2001; Kováč et al. 

2007;  Harzhauser  &  Piller  2010).  The  Leitha  Mountains 

(southern part of Vienna Basin), type area of the Leitha lime-

stones, represented a shallow carbonate platform with exten-

sive carbonate production (Schmid et al. 2001; Strauss et al. 

2006). According to Riegl & Piller (2000) these Leitha lime-

stones were deposited in a gently sloping, shallow subtidal 

environment.

The Eisenstadt-Sopron Basin formed a small satellite basin 

of the Vienna Basin to the southeast, similar in many respects 

also to the larger Styrian Basin (e.g., Ebner & Sachsenhofer 

1995) in the south of Austria, both at the junction to the 

 Pannonian Basin system, and having similar Leitha limestones 

deposits in shoaling parts of the basins. The Styrian Basin 

opened in the Early Miocene during the final stage of the 

 Alpine  orogeny  (Ebner  &  Sachsenhofer  1995),  and  was 

 divi ded by the N–S Middle Styrian Hills into the Eastern and 

Western Styrian subbasins (Kröll 1988). Basin subsidence 

during the Karpatian was a result of extensional tectonics and 

led to marine flooding (Ebner & Sachsenhofer 1995); eruptive 

volcanism accompanying the tectonic activity continued in the 

Badenian (Balogh et al. 1994; Slapansky et al. 1999; Seghedi 

& Downes 2011). 

Middle Miocene shallow-water limestones in Central Europe 

are termed the Leitha limestones, a classical name that was 

already used during the 19

th

 century (e.g., Suess 1860). Lateron, 

the  stratigraphic  unit  was  defined  by  Papp  &  Steininger  in 

more detail (in Papp et al. 1978), although this definition does 

not  conform  to  modern  stratigraphic  codes  (Hedberg  1976; 

Salvador 1994; Steininger & Piller 1999) — as a consequence 

the term is not defined formally, and we use the term Leitha 

limestone as an informal name throughout the following  paper. 

Leitha limestone thus refers to Middle Miocene shallow-water 

carbonate units composed mainly of corallinacean algae and 

subordinate coral-bearing strata. The latter were defined origi-

nally as being reefs (Papp et al. 1978; Dullo 1983; Tollmann 

1985), but later investigations (Piller & Kleemann 1991; Piller 

et al. 1996, 1997; Riegl & Piller 2000) confirmed that these 

carbonates are not reefal deposits sensu stricto but rather coral 

carpets such as at the type locality in the Leitha Mountains of 

eastern Austria (Riegl & Piller 2000). 

The unit is dated mainly to the Badenian (regional Central 

Paratethys Middle Miocene stage) which is equivalent to  

the Langhian–early Serravallian stages (Piller et al. 2007; 

 Hohenegger et al. 2014). The existence of the lost Paratethys 

Sea was already proposed by Laskarev (1924) due to its diffe-

rent biogeographic entity compared to the Mediterranean 

 Neogene. The Paratethys was subdivided into three palaeo-

geographic and geotectonic units, the Western, Central and 

Eastern Paratethys, respectively. The Eastern Alpine- Carpathian 

Foreland basins, from Lower Austria to Moldavia, the study 

area of the Vienna, Eisenstadt and Styrian basins in Austria, 

and the adjacent large Pannonian Basin System were included 

into the Central Paratethys (Piller et al. 2007).

Although a huge amount of palaeontological data is avai lable 

for  these  Miocene  limestones  (e.g.  Riegl  &  Piller  2000; 

Schmid et al. 2001), the exact chronostratigraphic position of 

individual units of Leitha limestones remains hard to unravel. 

In general, mostly a middle to late Badenian (Langhian to 

 early Serravalian) age was deciphered more recently (e.g., 

 Reuter et al. 2012; Wiedl et al. 2012, 2013, 2014; Hohenegger 

et al. 2014). Numerical ages for the Badenian are rare; tuffs 

intercalated with Leitha limestones in the Styrian basin at 

Retznei quarry were dated using 

40

Ar/

39

Ar method on biotite, 

background image

250

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

and  yielded  an  age  of  14.21± 0.07  Ma,  together  with  three 

 sanidine  crystals  of  14.34 ± 0.12  Ma  (Handler  et  al.  2006). 

 Hohenegger et al. (2009) and Hohenegger & Wagreich (2012), 

using cyclostratigraphy and astrochronology on combined 

outcrops and core data of the Badenian type locality in the 

southern Vienna Basin, dated the middle Badenian to 14.22 to 

13.98 Ma (see also Hohenegger et al. 2014). Strauss et al. 

(2006) used 3D seismic reflection data, well-drill data, surface 

outcrops and refined biostratigraphy in the interpretation of 

sequence stratigraphic framework for the southern and central 

Vienna Basin. Three Badenian 3

rd

 order depositional cycles, 

also recognized in the Styrian Basin (Schreilechner & Sachsen-

hofer 2007), are correlated with the TB2.3., TB2.4. and TB2.5. 

cycles from the global sea-level charts of Haq et al. (1987) 

with most Leitha limestones corresponding to the TB2.4 and 

TB2.5 sequences, starting at 14.8 Ma and 13.6 Ma, respecti-

vely (Hohenegger et al. 2014).

Sedimentary facies of Leitha limestones were investigated 

from many localities in Austria such as the type-locality at 

Fenk quarry (Grosshoeflein, Leitha Mts., Burgenland province, 

Austria, Fig. 1), where a sequence of ten coral intervals was 

observed (Riegl & Piller 2000). According to the authors these 

coral intervals represent a sequence of coral carpets and non-

frame building coral communities. 

Studied areas

Mannersdorf quarry (samples are denoted by M14/n, where 

n is sample number) is an active quarry of the Lafarge Zement-

werke GmbH close to the village Mannersdorf in  Lower 

Austria (Fig. 1), in the SW marginal part of the Vienna Basin. 

The quarry is located in the NE marginal part of the Leitha 

Mountains (Fig. 1) extends in NE–SW direction, and the 

 mining activity continues in a SW direction. The Leitha lime-

stones overlie mainly Middle Triassic dolostone and in some 

parts there is a deepening upward (transgressive) succession 

preserved, represented by basal breccias — gravel/conglo-

merate–sandstone–limestone (Fig. 2). In the NE part of the 

quarry a gravel layer underlies the limestones, described by 

Wiedl et al. (2012) as well rounded, poorly sorted, and com-

posed of granite, quartz and dolostone pebbles and fine quartz 

sand. Detailed stratigraphy and facies description were given 

by Wiedl et al. (2012); the facies types determined by these 

authors will be also used in our study. Besides cross-bedded 

gravel facies, basal breccia facies and rhodolith facies, the 

 bioclastic corallinacean facies include seven subfacies (i.e. 

Acervulina-rhodolite, Mollusc, Amphistegina, Bryozoan, Coral 

debris, Pholadomya and Pinna subfacies) (Wiedl et al. 2012). 

The abandoned Wöllersdorf quarry area (samples are deno-

ted by WÖ/n) is situated on the western margin of the Vienna 

Basin (Fig. 1). Here, Miocene sediments transgress mainly 

Triassic carbonates of the Northern Calcareous Alps. Several 

metres of Leitha limestones yield abundant Corallinacea, and 

molluscs (Wessely 2006). Mainly algal rudstones, grainstones 

and packstones were described from this area (Rohatsch 

2005).

The Kummer quarry (samples are denoted by SK/n) is located 

2 km to the East of St. Margarethen village in Burgenland, 

Austria (Fig. 2), within the Eisenstadt–Sopron Basin. The basin 

is surrounded by the Leitha Mountains to the North, 

Fertőrákos–Ruster Hügelland to the East, the Sopron Hills to 

Fig. 1. a — General map of Austria with Leitha mountains location; b — geological map of the Leitha mountains and the southern Vienna 

Basin, including Leitha limestones distribution (after Strauss et al. 2006).

background image

251

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

the South and Rosalia mountains to the West. The Leitha lime-

stones overly Neogene gravels and Austroalpine basement 

rocks. In the current paper we are following the sedimentology 

and stratigraphy of the quarry provided by Schmid et al. 

(2001) who defined three sedimentary facies (i.e. coralli-

nacean debris facies, laminated marl facies and rudstones 

 facies) from the quarry (Fig. 2). 

Fertőrákos quarry (samples are denoted by F/n) is located to 

the south of Kummer quarry in Hungary and has a largely 

 similar geological setting and stratigraphy as Kummer quarry, 

being also part of the Eisenstadt–Sopron Basin (Fig. 2).

Rosenberg quarry (samples are denoted by RR/n) (40 km 

south of Graz, Fig. 3) is part of the Retznei cement quarries  

of the Lafarge Zementwerke GmbH in the Styrian Basin.  

The Leitha limestones in the area are correlated with the 

 second Badenian transgression into the Central Paratethys 

which corresponds to TB 2.4 of the global sea level of Haq et 

al. 1987; see also Strauss et al. (2006). The stratigraphy of the 

studied locality shows a mixture of carbonates with silici- 

volcaniclastic sediments which were investigated and facies 

types were determined by Reuter et al. (2012). Four deposi-

tional units and six facies types (reef facies, inter-reef facies, 

coralline algal debris facies, rhodolith-Porites facies, quartz 

sand-Planostegina facies and coral carpet facies) are reported 

from the mixed silici-volcaniclastic succession at the 

 Rosenberg quarry (Reuter et al. 2012).

Methods

In total, 91 sediment samples were collected. The 41 sam-

ples from Mannersdorf include eight clastic samples for deci-

phering siliciclastic background values (i.e. one sand, one 

 conglomerate, one sandstone and five intercalated clay sam-

ples) and 33 samples represent various Leitha limestones 

 facies types. Six Leitha limestone samples were collected at 

the Wöllersdorf quarry. A total of thirty samples were taken 

from Kummer quarry near St. Margarethen village including 

twenty-three limestone samples, four marly facies and three 

clastic samples. Eleven samples were collected at Rosenberg 

quarry, two from the lower siliciclastics, four samples represent 

limestones, one tuff sample and four upper siliciclastics 

 samples. Only three limestone samples were taken from the 

Fertőrákos locality. Limestone samples include both pure and 

impure limestones featuring visible siliciclastic contamination.

Rock samples were cut into 5×5 cm slabs for thin sections, 

while all samples were powdered for X-Ray Diffraction 

(XRD) and whole rock geochemical analyses. The thin sec-

tions were investigated microscopically and photomicro-

graphs were taken by Leica microscope (Leica DM2700P) 

with Leica camera (Leica MC170 HD) connected to computer 

software (LAS v4.4.0). For XRD analysis the powdered sam-

ples  were  analysed  with  a  Panalytical  PW  3040/60  X’Pert 

PRO diffractometer (CuKα radiation, 40 kV, 40 mA, step size 

0.0167, 5 s per step). The X-ray diffraction patterns were inter-

preted using the Panalytical software “X´Pert High score plus” 

at the University of Vienna, Austria. SEM pictures were taken 

using a FEI Scanning Electron Microscope at the University 

of Vienna, Austria. The apparatus is equipped with secondary 

electron detectors for operation at high vacuum and low 

 vacuum, as well as a back scattered electron detector and an 

energy dispersive X-ray detection unit (EDAX Apollo XV). 

The chemical bulk rock analyses including major oxides, 

some  minor  elements  (see  Tables  1,  2  &  3)  and  loss  on  

ignition (LOI) were detected by ICP-emission spectroscopy  

by  

Acme 

Labs (Acme Analytical Laboratories Ltd., 

 www. acmelab.com; now Bureau Veritas Mineral Labs) in 

Vancouver (Canada). For the analysis of trace elements (REE 

Fig. 2. Simplified geological map of the Eisenstadt–Sopron Basin,  

SE Vienna Basin and westernmost Pannonian Basin with location of 

Kummer and Fertőrákos quarries (after Schmid et al. 2001).

Fig. 3. Geological map of the Styrian Basin with location of the  studied 

Rosenberg quarry near Retznei village (after Reuter et al. 2012).

background image

252

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

and base metals) ICP-mass spectroscopy was used 

 

(www.acmelab.com). Standard deviations are mainly in the 

range  of  0.1  to  1 %  (for  further  details  see  Neuhuber  et  al. 

2016). In addition, three samples of cements and individual 

shells (cement from M14/43, oyster shell from SK/28 and 

 pectinid shell from SK/19) were selected for chemical analy-

ses to be compared to the whole rock chemical analyses of the 

same samples. 

The PAAS-normalized REEs patterns were obtained by 

 dividing the measured elements by the same elements of 

 Taylor & McLennan (1985). The PAAS value of REEs are as 

follows  La = 38,  Ce = 80,  Pr = 8.9,  Nd = 32,  Sm = 5.6,  Eu = 1.1, 

Gd = 4.7, Tb = 0.77, Dy = 4.4, Ho = 1, Er = 2.9, Tm = 0.4, Yb = 2.8 

and Lu = 0.43 ppm (Taylor & McLennan 1985). Ce anomaly 

was quantified after the normalization of the neighbouring 

 elements La and Pr throughout the equation (De Baar et al. 

1983):  Ce/Ce* = 2CeN/(La+Pr)N,  and  the  Eu  anomaly  was 

evaluated  using  the  equation:  Eu/Eu* = EuN/√SmN*GdN, 

where the subscript N refers to normalized value to PAAS.   

Sr/Ca ratio was calculated after the conversion of CaO wt. % 

into Ca ppm by the equation (CaO wt. % * 0.7147) * 10000 =  

Ca ppm) to each sample then Ca was divided by Sr and expres-

sed in a reverse manner where the numerator represents Sr and 

equals 1 and denominator represents the result of the division. 

These calculations were summed and averaged to all lime-

stone samples.

Many indices were proposed to evaluate the chemical 

weathering such as chemical index of alteration (CIA) = [Al

2

O

3

(Al

2

O

3

 + CaO* + Na

2

O + K

2

O)] × 100,  chemical  index  of 

weathe ring  (CIW) = [Al

2

O

3 

/ (Al

2

O

3

 + CaO + Na

2

O)] × 100  or 

CIW` = [Al

2

O

3 

/ (Al

2

O

3

 + Na

2

O)] × 100, where CaO* represents 

the calcium content derived from silicates (Nesbitt & Young 

1982; Harnois 1988; Chittleborough 1991; Cullers 2000).  

The CIA is applied here to determine the degree of chemical 

weathering in which CaO* is set equal to Na

2

O because of the 

high content of calcite in all samples (McLennan et al. 1993). 

Mn*  values  were  calculated  by  the  equation  [(Mn*=  

Log (Mn

sample 

/ Mn

shale

) / (Fe

sample 

/ Fe

shale

) according to Wedepohl 

(1978) average Mn

shale

  value  is  600  ppm  and  average  Fe

shale

 

value is 46150 ppm].

Results

Mineralogy

The bulk mineralogy of limestone samples from Manners-

dorf examined by SEM, optical microscopy and XRD allowed 

the discrimination of four categories. The mineral composi-

tion of these categories is: (1) only calcite (pure limestone,  

Fig. 4a, b); (2) calcite and a small amount of quartz (also pure 

limestones); (3) calcite and a small amount of quartz and dolo-

mite and (4) calcite, quartz and clay minerals (impure lime-

stone,  Fig.  4c, d).  Besides  that,  the  optical  microscopy  and 

Scanning Electron Microscopy (SEM) examination of the 

samples proved the presence of muscovite (Fig. 4a), clay 

 minerals and feldspars with variable amounts in all studied 

limestones facies and sometimes pyrite and barite.  

In this regard the Leitha limestones lithofacies are treated 

here separated in two main groups: (1) visibly “pure” lime-

stones with minor amounts of detrital materials only con-

firmed by SEM and in some cases by optical microscopy,  

(2) “impure” limestones with variable to abundant amounts of 

the detritus revealed by XRD.

Samples from Wöllersdorf are subdivided into two groups 

according to the mineralogical composition, namely: (1) litho-

facies mainly consisting of calcite, quartz and muscovite and 

(2) lithofacies composed of calcite, dolomite and quartz, both 

lithofacies are considered here as impure limestones. Lime-

stones from the Kummer quarry display two main lithofacies: 

(1) limestones composed mainly of calcite and minor quartz 

abundance  (pure  limestones,  Fig.  4e, f),  and  (2)  marly  

facies composed of calcite with quartz and clay minerals.  

The Fertőrákos limestone samples are similar to the Kummer 

pure limestones with mainly calcite and quartz as mineral 

components. 

The Rosenberg Leitha limestones comprise calcite, dolo-

mite and quartz, dolomite is significantly more abundant 

(higher intensity peak in XRD chart, sample RR/7) in the 

 sample directly in contact with the tuff layer, the mineralogy 

of the latter layer is quartz, calcite, muscovite, albite,  orthoclase 

and gypsum. 

Geochemistry

Major elements

The average chemical compositions of the Mannersdorf 

samples are shown in Table 1. In pure Leitha limestones sam-

ples the major oxide is CaO average of 54.98 wt. % (n=25).  

In contrast the lithofacies with higher clay mineral contents 

(impure limestones) show a gentle variation in the major 

 oxides in which the average CaO content is lower with value 

of 51.56 wt. % (n=6) while average Al

2

O

3

 content is 1.19 wt. %.

At Wöllersdorf, the two Leitha limestones lithofacies which 

are impure limestones exhibit different chemical characte ris-

tics (Table 2) in which the major oxides compositions in the 

calcite lithofacies are dominated by CaO with an average 

 value  of  53.2  wt. %  (n=3)  and  MgO  average  value  is  

0.64  wt. %.  The  limestones  lithofacies  with  dolomite  have 

 average values of CaO = 47.3 wt. % and MgO = 5.7 wt. % (n=3).

At Kummer, the pure limestone lithofacies dominated by 

calcite have a CaO average value of 54.4 wt. % (n=23) and an 

average Al

2

O

3

 = 0.14  wt. %.  Marly  lithofacies  have  a  lower 

CaO average of 46.1 wt. % (n=4), average SiO

2

 of 8.7 wt. %, 

average Al

2

O

3

 of 3.2 wt. %, average Fe

2

O

of 0.9 wt. %, while 

MgO; K

2

O have average values of 0.8 and 0.6 wt. %, respec-

tively. Fertőrákos Leitha limestones samples do not show dif-

ferences from Kummer samples, showing that they are pure 

(Table 2) with a CaO average of 53.97 wt. % (n=3). 

The Rosenberg Leitha limestones samples (n=4) show dif-

ferences in chemical compositions with respect to their strati-

background image

253

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

Fig. 4. Mineralogy of Leitha limestones: a — SEM image of pure limestone with the existence of muscovite (Mu), M14/3; b — XRD 

 diagram of pure limestone from Mannersdorf; c — photomicrograph of impure limestone with a lithoclast composed mainly of microcline 

(Mc) and quartz (Qz), M14/21; d — XRD diagram of impure limestone from Mannersdorf; e — photomicrograph of pure limestone from 

Kummer, SK/1; f — XRD diagram of representative samples of pure limestone from Kummer.

background image

254

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

graphic position within the Leitha limestone interval. Two 

samples were collected from the lower and upper parts of  

the main limestone bank (Reutel et al. 2012), in contact to 

 underlaying and overlaying siliciclastics, and two samples 

from the middle portion of the bank (Table 3). In the former 

ones  the  average  CaO  is  51.46  wt. %  and Al

2

O

3

 average is  

0.75 wt. %. The latter samples have higher CaO with average 

of  55.12  wt. %  and  lower  Al

2

O

3

  average  of  0.06  wt. %.  

The  tuff  layer  (sample  RR/6)  has  the  lowest  SiO

2

 content 

among  the  sampled  siliciclastics  with  16.93  wt. %  and  the 

highest  content  of  MgO  with  8.31  wt. %.  The  CaO, Al

2

O

3

Fe

2

O

3

 and K

2

O contents are 29.91, 5.73, 3.57 and 1.26 wt. %, 

respectively. 

Minor elements

Minor element contents in the Mannersdorf samples are low 

in the limestone facies compared to the siliciclastics (Table 1). 

The only exception is strontium (Sr) which has a significantly 

higher content in the limestone facies. The average Sr content 

in the different limestone facies varies from pure (221 ppm,  

n= 25) to impure limestones (224 ppm, n=6) at Mannersdorf, 

recording Sr/Ca ratios with an average of 1/1845. Interes-

tingly, the arsenic (As) content in siliciclastic samples is 

 variable and ranges from low (9.6 ppm) to high (136.6 ppm) 

contents. In contrast As content in the limestones is conside-

rably lower.

Rock type

Underlying clastics

n=3

Intercalated clastics

n= 5

Pure limestone 

dominated by calcite

n= 8

Pure limestone with 

minor quartz

n= 17

Impure limestone with 

clay minerals

n= 6

Limestone with minor 

dolomite

n=2

SiO

2

54.46 ± 11.6

47.2 ± 21.8

0.34 ± 0.21

1 ± 0.43

3.68 ± 1.56

0.86 ± 0.26

Al

2

O

3

5.05 ± 2.99

13.92 ± 5.11

0.1 ± 0.07

0.31 ± 0.14

1.19 ± 0.58

0.24 ± 0.07

Fe

2

O

3

1.65 ± 0.1

4.71 ± 2.24

0.06 ± 0.04

0.13 ± 0.06

0.57 ± 0.15

0.1 ± 0.01

MgO

0.4 ± 0.2

1.18 ± 0.52

0.34 ± 0.02

0.31 ± 0.1

0.47 ± 0.12

0.94 ± 0.52

CaO

18.69 ± 7.8

12.76 ± 16.3

55.47 ± 0.8

54.76 ± 0.53

51.56 ± 1.23

54.2 ± 0.28

Na

2

O

0.19 ± 0.22

0.28 ± 0.34

0.02

K

2

O

0.81 ± 0.52

2.25 ± 1.05

0.02 ± 0.01

0.05 ± 0.03

0.21 ± 0.11

0.03 ± 0.01

MnO

0.02 ± 0.01

0.05 ± 0.04

0.05 ± 0.03

0.01

LOI

18.27 ± 6.89

16.3 ± 11.7

43.59 ± 0.66

43.34 ± 0.27

42.1 ± 1.1

43.5 ± 0.14

As

24.37 ± 6.57

80.6 ± 39.97

1.88 ± 0.95

2.65 ± 1.24

10.55 ± 6.54

3.55 ± 1.91

Co

2.63 ± 0.35

10.12 ± 6.63

0.47 ± 0.26

0.61 ± 0.38

1.6 ± 0.85

Ni

69.2 ± 25.9

Cr

111.8 ± 31.6

101± 35.4

5.7 ± 9.1

V

46.3 ± 13.2

114.6 ± 45.3

9 ± 3.2

15.7 ± 2.7

23.17 ± 6.24

18.5 ± 3.53

Sr

367.2 ± 424.6

138.2 ± 55.8

279. 6 ± 40.9

208.4 ± 22.45

195.4 ± 59.3

170. 5 ± 37.3

Rb

35.37 ± 30.17

103 ± 56.65

0.79 ± 0.53

2.2 ± 1.06

7.9 ± 4.49

1.8 ± 0.57

Cs

12.2 ± 15.77

12.9 ± 7.99

0.18 ± 0.1

0.29 ± 0.18

1.27 ± 0.53

0.35 ± 0.07

Ba

222 ± 141

610 ± 371

10.5 ± 2.67

18 ± 9.69

220.5 ± 323.4

119.5 ± 135.1

Pb

3.67 ± 0.7

13.66 ± 5.7

0.51 ± 0.2

0.76 ± 0.37

2.23 ± 0.5

0.6 ± 0.14

Zr

99.7 ± 37.7

175.3 ± 82.5

3.35 ± 0.65

4.68 ± 1.69

17.75 ± 7.75

10.15 ± 0.35

Y

4.9 ± 1.5

25.68 ± 10.3

0.9 ± 0.15

1.61 ± 0.38

14.83 ± 19. 55

2.1 ± 0.28

Nb

4.17 ± 2.28

12.36 ± 5.21

0.2 ± 0.07

0.62 ± 0.53

Hf

2.53 ± 0.75

4.74 ± 2.08

0.1 ± 0.5

0.15 ± 0.08

0.47 ± 0.23

0.1

Th

1.8 ± 1.28

11.74 ± 5.44

0.2 ± 0.07

0.35 ± 0.2

0.98 ± 0.46

0.2

U

1.3 ± 0.36

9.78 ± 7.33

1.18 ± 0.43

2.86 ± 1.5

4.6 ± 2.6

3.5 ± 0.28

La

6.9 ± 4.78

27.7 ± 15.31

0.98 ± 0.29

1.62 ± 0.52

4.57 ± 1.22

0.95 ± 0.21

Ce

13.53 ± 10.77

54.4 ± 27.77

1.33 ±0.36

2.45 ± 0.81

9.02 ± 2.46

1.45 ± 0.07

Pr

1.53 ± 1.26

6.71 ± 3.29

0.13 ± 0.04

0.29 ± 0.1

1.05 ± 0.31

0.19 ± 0.02

Nd

6.13 ± 5.24

27.1 ± 12.53

0.65 ± 0.23

1.19 ± 0.4

4.58 ± 1.48

0.75 0.35

Sm

1.1 ± 1.01

5.53 ± 2.47

0.11 ± 0.05

0.21 ± 0.07

1 ± 0.48

0.16 ± 0.04

Eu

0.36 ± 0.26

1.12 ± 0.51

0.03 ± 0.02

0.05 ± 0.02

0.33 ± 0.2

0.06 ± 0.01

Gd

1.24 ± 0.8

5.17 ± 2.16

0.13 ± 0.02

0.24 ± 0.07

1.7 ± 1.47

0.28 ± 0.01

Tb

0.2 ± 0.12

0.8 ± 0.37

0.02 ± 0.01

0.04 ± 0.01

0.3 ± 0.3

0.05 ± 0.01

Dy

1.11 ± 0.59

4.48 ± 1.89

0.12 ± 0.04

0.22 ± 0.05

1.88 ± 2.03

0.32

Ho

0.21 ± 0.07

0.91 ± 0.39

0.03 ± 0.01

0.05 ± 0.01

0.43 ± 0.5

0.07 ± 0.01

Er

0.57 ± 0.18

2.4 ± 1.13

0.1 ± 0.04

0.14 ± 0.04

1.24 ± 1.5

0.2 ± 0.06

Tm

0.09 ± 0.03

0.33 ± 0.18

0.02 ± 0.01

0.02 ± 0.01

0.18 ± 0.23

0.03

Yb

0.6 6 ± 0.17

2.13 ± 1.13

0.08 ± 0.04

0.12 ± 0.03

1.16 ± 1.52

0.17 ± 0.01

Lu

0.11 ± 0.03

0.33 ± 0.17

0.02 ± 0.01

0.02 ± 0.01

0.2 ± 0.25

0.03

Ce/Ce*

0.94 ± 0.04

0.93 ± 0.02

0. 84 ± 0.18

0.91 ± 0.07

0. 95 ± 0.08

0.92

Eu/Eu*

1.38 ± 0.16

0.98 ± 0.1

1.56 ± 0.76

1.05 ± 0.22

1.16 ± 0.17

1.13 ± 0.09

ΣREEs

33.74 ± 24.76

139.1 ± 68.9

3.59 ± 0.93

6.65 ± 1.98

27.61 ± 9.79

4.7 ± 0.18

(La/Yb)

N

0.74 ± 0.42

0.94 ± 0.07

0.92 ± 0.39

0.99 ± 0.2

0.65 ± 0.41

0.42 ± 0.08

Y/Ho

23.24 ± 1.2

28.5 ± 1.8

30.6 ± 9.4

36.7 ± 11.6

33.2 ± 7.4

30.2 ± 2.1

* oxides in wt. %, elements in ppm and (–) indicates not detected 

Table 1: Average chemical compositions* and standard deviations of Mannersdorf Leitha limestones facies types and siliciclastics.

background image

255

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

In the Wöllersdorf samples the minor element concentra-

tions are still low without major variations between the two 

lithofacies and similar to those from Mannersdorf quarry, 

 except for Sr which is lower in the lithofacies containing 

 dolomite with an average value of 141.5 ppm, with Sr/Ca ratio 

of 1/2412, and the other lithofacies has an average value of 

253 ppm and Sr/Ca ratio of 1/1506.   

In general minor elements in the Kummer samples are again 

low except for Sr. Unlike the Mannersdorf and Wöllersdorf 

samples, Sr content is higher in the Leitha limestones litho-

facies with an average of 571 ppm (n=23). The Sr/Ca ratio in 

this locality also differs significantly from Mannersdorf and 

Wöllersdorf,  with  average  value  of  1/716.  The  average  

value  in  the  intercalated  marly  facies  is  360  ppm.  The  Sr  

content  at  Fertőrákos  averages  629  ppm  and  Sr/Ca  ratio  of 

1/613. 

The Rosenberg samples show high Sr contents compared to 

the other localities not only in the Leitha limestones but also in 

surrounding siliciclastics. The average content in limestones 

(both pure and impure, n=4) equals 1271 ppm with Sr/Ca ratio 

average of 1/301, while Sr in the tuff layer has a value of 

1074.1 ppm. 

Rock type

Wöllersdorf impure 

limestones with 

clays

n=3

Wöllersdorf impure 

limestones with 

dolomite

n=3

Kummer pure 

limestones

n= 23

Kummer marls

n= 4

Kummer 

siliciclastics

n=3

Fertőrákos pure 

limestones

n= 3

SiO

2

2.38 ± 1.46

2.96 ± 1.45

0.61 ± 0.2

8.7 ± 1.53

26.73 ± 9.02

1.78 ± 0.44

Al

2

O

3

0.75 ± 0.41

0.52 ± 0.13

0.14 ± 0.06

3.18 ± 0.48

9.12 ± 2.59

0.27 ± 0.08

Fe

2

O

3

0.21 ± 0.1

0.22 ± 0.05

0.28 ± 0.17

0.9 ± 0.2

3.26 ± 2.06

0.09 ± 0.01

MgO

0.64 ± 0.01

4.7 ± 2.9

0.61 ± 0.06

0.75 ± 0.12

1.69 ± 0.3

0.57 ± 0.03

CaO

53.21 ± 1.31

47.26  ± 4.38

54.35 ± 0.45

46.11 ± 1.18

27.81 ± 9.11

53.97 ± 0.31

Na

2

O

0.01

0.01

0.03 ± 0.01

0.04 ± 0.01

0.1 ± 0.02

0.08 ± 0.03

K

2

O

0.13 ± 0.07

0.1 ± 0.03

0.03 ± 0.01

0.56 ± 0.07

1.61 ± 0.46

0.0.3 ± 0.01

MnO

0.06 ± 0.05

0.13 ± 0.14

0.02 ± 0.01

0.05 ± 0.02

LOI

42.47 ± 0.78

43.1 ± 0.21

43.75 ± 0.24

39.3 ± 1.12

28.8 ± 5.68

42.97 ± 0.23

As

1.43 ± 0.29

1.7 ± 0.7

1.7 ± 1.3

3 ± 2.76

159.8 ± 260.4

0.77 ± 0.06

Co

0.53 ± 0.25

0.4 ± 0.14

0.42 ± 0.25

0.98 ± 0.32

6.97 ± 5.58

0.25 ±0.15

Ni

2.93 ± 1.27

4.03 ± 0.32

1.99 ± 1.45

3.08 ± 1.08

46.87 ± 43.89

0.83 ± 0.4

Cr

25.66 ± 6.55

68.42 ±23.7

V

011 ± 0.03

0.07 ± 0.01

11.33 ± 5.17

30.75 ± 2.36

94.67 ± 40.51

15.33 ± 5.77

Sr

253.2 ± 15.5

141.5 ± 24.6

570. 7 ± 139

196.6 ± 53.4

359.8 ± 113.1

629 ± 45

Rb

5.97 ± 2.82

3.93 ± 1.03

1.31 ± 0.53

29.63 ± 3.18

91.33 ± 30.33

1.57 ± 0.35

Cs

0.43 ± 0.23

0.33 ± 0.16

0.21 ± 0.17

3.98 ± 0.62

14.13 ± 5.69

0.23 ± 0.06

Ba

18 ± 7.8

9 ± 1

20.57 ± 4.07

70.25 ± 3.3

170.3 ± 47.4

17.33 ± 0.58

Pb

0.9 ± 0.2

1.1 ± 0.26

1 ± 0.33

1.8 ± 0.39

9.1 ± 6.97

0.7

Zr

16.93 ± 7.03

22.63 ± 13.75

3.71 ± 2.65

50.8 ± 58.4

59.8 ± 17.15

6.4 ± 4.9

Y

2.6 ± 0.72

2.77 ± 0.59

6.15 ± 1.15

7.7 ± 1.2

18.57 ± 5.74

6.2 ± 0.4

Nb

0.8 ± 0.46

0.37 ± 0.21

0.25 ± 0.07

2.2 ± 0.59

6.6 ± 1.9

Hf

0.47 ± 0.21

0.6 ± 0.36

0.15 ± 0.08

1.4 ± 1.41

1.7 ± 0.52

Th

0.6 ± 0.26

0.67 ± 0.12

0.44 ± 0.16

2.5 ± 0.45

7.53 ± 2.48

0.43 ± 0.06

U

1.63 ± 0.78

1.7 ± 0.44

1.97 ± 0.47

2.18 ± 1.06

3.9 ± 1.51

1.3 ± 0.26

La

2.4 ± 1.13

2.7 ± 0.5

3.76 ± 0.73

8.33 ± 0.92

22.53 ± 7.19

3.33 ± 0.25

Ce

4.57 ± 2.63

4.8 ± 0.87

4.46 ± 1.01

14.43 ± 1.96

43.17 ± 15.74

4.1 ± 0.4

Pr

0.54 ± 0.27

0.59 ± 0.12

0.72 ± 0.14

1.78 ± 0.25

5.03 ± 1.59

0.69 ± 0.05

Nd

2.17 ± 1.16

2.37 ± 0.58

3.18 ± 0.63

7.08 ± 1.24

19.23 ± 6.19

3 ± 0.1

Sm

0.44 ± 0.24

0.5 ± 0.15

0.65 ± 0.14

1.35 ± 0.28

3.72 ± 1.07

0.62 ± 0.08

Eu

0.09 ± 0.06

0.1 ± 0.03

0.16 ± 0.3

0.31 ± 0.05

0.83 ± 0.29

0.17± 0.02

Gd

0.43 ± 0.16

0.54 ± 0.07

0.82 ± 0.14

1.33 ± 0.26

3.6 ± 1.3

0.83 ± 0.08

Tb

0.07 ± 0.03

0.08 ± 0.02

0.13 ± 0.02

0.21 ± 0.03

0.55 ± 0.19

0.12 ± 0.01

Dy

0.44 ± 0.19

0.45 ± 0.11

0.79 ± 0.15

1.2 ± 0.21

3.08 ± 1.03

0.81 ± 0.09

Ho

0.08 ± 0.03

0.09 ± 0.02

0.16 ± 0.02

0.22 ± 0.03

0.59 ± 0.22

0.16 ± 0.03

Er

0.23 ± 0.08

0.27 ± 0.06

0.47 ± 0.08

0.67 ± 0.12

1.67 ± 0.52

0.46 ± 0.05

Tm

0.03 ± 0.01

0.04 ± 0.01

0.06 ± 0.01

0.095 ± 0.01

0.23 ± 0.08

0.06 ± 0.01

Yb

0.2 ± 0.09

0.23 ± 0.08

0.38 ± 0.07

0.58 ± 0.07

1.51 ± 0.45

0.38 ± 0.09

Lu

0.04 ± 0.01

0.04 ± 0.01

0.06 ± 0.01

0.1 ± 0.01

0.22 ± 0.06

0.06 ± 0.01

Ce/Ce*

0.9 ± 0.08

0.87 ± 0.07

0.62 ± 0.05

0.86 ± 0.02

0.92 ± 0.04

0.62 ± 0.01

Eu/Eu*

0.88 ± 0.26

0.91 ± 0.19

1.02 ± 0.09

1.08 ± 0.1

1.05 ± 0.04

1.09 ± 0.17

ΣREEs

11.72 ± 6.07

12.8 ± 2.5

15.79 ± 2.91

37.67 ± 5.24

106 ± 35.9

14.8 ± 0.98

(La/Yb)

N

0.88 ± 0.04

0.9 ± 0.14

0.74 ± 0.1

1.06 ± 0.03

1.1 ± 0.04

0.66 ± 0.13

Y/Ho

33 ± 1.7

31 ± 2.2

38.56 ± 3.05

34.6 ± 2.8

31.7 ± 2

39 ± 4.7

* oxides in wt. %, elements in ppm and (–) indicates not detected

Table 2: Average chemical compositions* and standard deviations of Wöllersdorf, Kummer and Fertőrákos Leitha limestones facies types and 

Kummer siliciclastics.

background image

256

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

REEs patterns

In the Mannersdorf samples the overall concentration of the 

REEs in the carbonate facies in general is lower and varying 

due to the clay content compared to the intercalated clastics 

with  contents  higher  than  110  ppm.  The  average  ƩREEs  is  

5.7  ppm  (n=25)  in  the  absence  of  clays  (pure  limestones) 

while this average is enhanced due to the presence of clays to 

27.6  ppm  (n=6).  The  total  REEs  value  is  strongly  positive 

 correlated  to  Al

2

O

3

  wt. %  (r=0.95),  as  well  as  Zr  (r = 0.92);  

on  the  other  hand,  it  is  negatively  correlated  to  CaO  wt. % 

(r = − 0.84).  PAAS  normalized  REEs  patterns  of  the  inter-

calated clastics are flattened and show average shale patterns 

(Fig. 5a). PAAS (Taylor & McLennan 1985) normalized REEs 

patterns of limestones facies from Mannersdorf are almost flat 

with negative Ce anomaly in pure limestones (Fig. 5b, c)  

but this anomaly is less pronounced in impure limestones  

(Fig. 5d). The LREEs/HRREs ratio is marked by the (La/Yb)

N

 

ratio which has an average value around 1 (i.e. 0.97, n=30), 

apart from some samples which exhibit HREEs enrichment 

with an average value of 0.45 (n=9).

The total REEs content in Wöllersdorf samples averages 

12.3  ppm  (n=6).  The  PAAS  normalized  RREs  patterns  are 

nearly flat (Fig. 6a) with low LRREs/HRREs ratio indicated 

by (La/Yb)

N

 ratio ranging from 0.74 to 0.99 with an average  

of 0.86.

Table 3: Average chemical compositions* and standard deviations of Rosenberg Leitha limestones facies types and siliciclastics.

Rock type

Rosenberg lower 

siliciclastics

n=2

Rosenberg pure 

limestones

n=2

Rosenberg impure 

limestones

n=2

Rosenberg tuff layer

Rosenberg upper 

siliciclastics

n= 4

SiO

2

25 ± 0.23

0.25 ± 0.04

2.24 ± 0.22

16.93

41.42 ± 7.47

Al

2

O

3

4.78 ± 1.71

0.06 ± 0.04

0.75 ± 0.07

5.73

8.93 ± 4.08

Fe

2

O

3

3.06 ± 0.35

0.06 ± 0.03

0.35 ± 0.01

3.57

3.88 ± 1.34

MgO

2.42 ± 0.04

0.64 ± 0.07

1.81 ± 0.93

8.31

4.17 ± 1.32

CaO

32.96 ± 6.33

55.1 ± 0.54

51.4 ± 1.3

29.91

17.86 ± 9.05

Na

2

O

0.63 ± 0.23

0.02

0.07 ± 0.01

0.44

0.99 ± 0.18

K

2

O

1.17 ± 0.36

0.02 ± 0.01

0.16 ± 0.03

1.26

1.91 ± 0.86

MnO

0.04

0.01

0.02

0.09 ± 0.07

LOI

28.95 ± 4.45

43.6 ± 0.42

42.85 ± 0.07

32.7

19.85 ± 5.6

As

5.05 ± 1.63

1.3 ± 0.42

1.85 ± 0.5

68.8

6.25 ± 2.92

Co

7.1 ± 2.97

0.65 ± 0.07

1.15 ± 0.35

4.6

10.1 ± 5.52

Ni

20.95 ± 5.02

0.5-

3.2 ± 1.13

35

47.28 ± 26.64

Cr

75.26 ± 9.68

123.2

148.8 ± 22.6

V

39.5 ± 9.19

15 ± 1.4

99

78 ± 38

Sr

750 ± 110

1378 ± 229

1164 ± 325

1074.1

518 ± 360

Rb

43.1 ± 12.3

0.8 ± 0.28

7.1 ± 0.42

48.5

73.1 ± 38.5

Cs

2.5 ± 0.28

0.2

0.6

38

4.63 ± 2.59

Ba

187.5 ± 105.4

20.5 ± 2.12

43.5 ± 13.4

206

285.3 ± 113

Pb

6.75 ± 3.6

0.95 ± 0.07

1.3 ± 0.28

25.7

7.45 ± 4.26

Zr

68.4 ± 11.6

1.3 ± 0.28

8.85 ± 0.21

78.4

137.1 ± 56.1

Y

13.95 ± 6.01

2.65 ± 0.07

2.95 ± 0.21

19.8

20.9 ± 7.2

Nb

5.75 ± 2.62

0.65 ± 0.07

6.7

9.1 ± 4.24

Hf

1.75 ± 0.35

0.2

2

3.73 ± 1.43

Th

4.1 ± 1.9

0.7 ± 0.14

7.4

6.08 ± 3

U

2.6 ± 1.6

1.2 ± 0.71

2.25 ± 1.77

9.5

3.28 ± 1.23

La

14.95 ± 6.15

1.4 ± 0.28

3.05 ± 0.21

18.2

22 ±9.01

Ce

29.5 ± 12.73

1.85 ± 0.07

5.05 ± 0.07

37.3

43.38 ± 18.36

Pr

3.28 ± 1.36

0.27 ± 0.02

0.63

4.64

5.11 ± 2.29

Nd

12.75 ± 6.01

1.15 ± 0.07

2.45 ± 0.49

18.7

19.3 ± 9

Sm

2.58 ± 1.36

0.22 ± 0.01

0.48 ± 0.08

3.79

3.99 ± 1.83

Eu

0.62 ± 0.28

0.08

0.12 ± 0.02

0.86

0.86 ± 0.35

Gd

2.64 ± 1.3

0.26

0.49 ± 0.04

3.7

3.67 ± 1.51

Tb

0.41 ± 0.18

0.06 ± 0.01

0.08 ± 0.01

0.57

0.6 ± 0.28

Dy

2.34 ± 1.2

0.29 ± 0.01

0.45 ± 0.01

3.16

3.55 ± 1.31

Ho

0.47 ± 0.18

0.07 ± 0.01

0.1 ± 0.01

0.62

0.7 ± 0.24

Er

1.31 ± 0.64

0.22 ± 0.01

0.27 ± 0.02

1.63

2.09 ± 0.74

Tm

0.18 ± 0.08

0.03 ± 0.01

0.05 ± 0.01

0.21

0.3 ± 0.11

Yb

1.21 ± 0.57

0.19 ± 0.04

0.22 ± 0.01

1.41

1.96 ± 0.74

Lu

0.2 ± 0.08

0.02

0.04 ± 0.01

0.22

0.29 ± 0.1

Ce/Ce*

0.96 ± 0.02

0.7 ± 0.08

0.84 ± 0.04

0.93

0.94 ± 0.01

Eu/Eu*

1.13 ± 0.05

1.58 ± 0.03

1.13 ± 0.35

1.07

1.06 ± 0.04

ΣREEs

72.42 ± 32.14

6.09 ± 0.33

13.44 ± 0.29

95.01

107.8 ± 45.6

(La/Yb)

N

0.93 ± 0.13

0.54 ± 0.01

1.04 ± 0.04

0.95

0.83 ± 0.17

Y/Ho

29.4 ± 1.3

39 ± 8.8

31 ± 0.08

32

29.8 ± 1.99

* oxides in wt. %, elements in ppm and (–) indicates not detected 

background image

257

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

Pure limestone facies from Kummer have PAAS normalized 

REEs patterns which resemble seawater patterns with negative 

Ce anomaly and HREEs/LREEs enrichment (Fig. 6b). The total 

REEs content in this facies has an average of 15.8 ppm (n=23), 

this content has very weak negative correlation with Zr  

(r = − 0.05), weak positive correlation with Al

2

O

3

 with r =0.27 

and weak negative correlation with CaO with r = − 0.21. The 

(La/Yb)

N

 ratio average of this facies is 0.74. The pattern shape 

in the marly facies is characterized by MREEs enrichment 

(Fig. 6c) and the REEs content is higher than in the limestones 

facies with average values of 37.7 ppm, as well as the (La/Yb)

N

 

ratio average of 1.1. The clastic facies has a flat pattern  

(Fig. 6c) with the highest concentration among the facies in 

Kummer, with average REEs content of 106 ppm and (La/Yb)

N

 

ratio  average  of  1.1.  Fertőrákos  pure  limestones  exhibit  

a  simi lar  REEs  pattern  to  those  from  Kummer  (Fig.  6d)  

with average content of 14.8 ppm and (La/Yb)

N

 ratio average  

of 0.66.

The Rosenberg limestones show two REEs patterns: (1) for 

two samples in contact with siliciclastics there is a flat pattern 

with average 13.4 ppm and (La/Yb)

N

 ratio average of 1.04 and 

a seawater like REEs pattern with pronounced Ce anomaly 

and REEs average content of 6.09 ppm and HREEs enrich-

ment indicated by (La/Yb)

N

 ratio average of 0.54 (Fig. 7a);  

(2) the REEs patterns of the underlying and overlying layers 

and the intercalated tuff layer are flat with MREEs enrich-

ments and (La/Yb)

N

 ratio average of 0.9 (Fig. 7b).

Ce anomaly

In Mannersdorf the values of the anomaly in the various 

rock types are approximately the same, the average value in 

pure  and  impure  limestones  is  0.9  (n= 33).  Samples  from 

Wöllersdorf have Ce/Ce* values close to those of the Manners-

dorf average of 0.88. The Kummer pure limestones facies has 

an  average  Ce  anomaly  of  0.62,  while  the  marly  facies  has  

a  higher  average  of  0.86  and  the  highest  average  value  is 

 recorded in the intercalated clastics where it reaches 0.92.  

The average Ce anomaly in limestones from Fertőrákos is also 

0.62. Pure limestones from Rosenberg quarry have average Ce 

anomaly value of 0.7 (n= 2) while the average value of the 

impure facies is 0.84 (n=2).  

Fig. 5. PAAS-normalized REEs pattern of various rock types from Mannersdorf: a — clastic samples including lower siliciclastics and inter-

calated ones; b — pure limestone samples composed mainly of calcite; c — pure limestones composed of calcite and quartz; d — impure Leitha 

limestones with muscovite and clays.

background image

258

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

Eu anomaly

Both pure and impure limestones from Mannersdorf have an 

average Eu anomaly value of 1.12 (0.68–2.15) which does not 

differ from the average value for the siliciclastics from the 

same locality which equals 1.13. The Eu anomaly from the 

Wöllersdorf limestones ranges from 0.58 to 1.1 with average 

of 0.9. Kummer samples including pure limestones, marly 

limestones and siliciclastics all have nearly the same Eu 

 ano maly with an average value of 1.02, 1.03 and 1.05, respec-

tively.  Fertőrákos  limestones  have  an  Eu  anomaly  average  

of 1.09. The Eu anomaly of pure and impure limestones  

Fig. 6. PAAS-normalized REEs pattern of Leitha limestones: a — impure limestone samples from Wöllersdorf; b — pure limestone samples 

composed mainly of calcite at Kummer; c — impure samples from Kummer quarry where the lower ones are marly and the upper ones are 

siliciclastics; d — pure limestones from Fertőrákos locality.

Fig. 7. PAAS-normalized REEs pattern of Leitha limestones: a — limestones from Rosenberg quarry including pure limestones (the lower two 

samples) and impure limestones (the upper two samples); b — siliciclastics samples from Rosenberg quarry.

background image

259

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

from  the  Rosenberg  quarry  have  averages  of  1.6  and  1.13, 

 respectively. 

Y/Ho ratio

The average Y

mol

/Ho

mol

 ratio of pure and impure limestones 

from Mannersdorf are approximately the same 34 and 33, 

 respectively. In Wöllersdorf the limestone lithofacies contai-

ning dolomite has an average value of 31 which is lower than 

the average value in the lithofacies dominated by calcite with 

clays where it equals 33. In the Kummer pure limestones 

 facies the average value is 39 and in the marly facies the ratio’s 

average is 35. The limestones from Fertőrákos have exactly 

the similar ratios as the Kummer limestones. At Rosenberg  

the Y/Ho ratio averages 35 (31– 45) in both pure and impure 

limestone facies.  

Discussion

 Factors controlling mineralogy

The Leitha limestones studied here have slightly different 

mineralogical compositions according to the sampling locality 

and its regional substratum geology. Samples from Manners-

dorf and Wöllersdorf show the effect of intermittent silici-

clastic input during limestone deposition as indicated by the 

presence of quartz and clay minerals, as well as the effect of 

diagenesis which led to the formation of dolomite. Various 

types of pure to impure limestones can be recognized in all the 

studied localities. Major impurities are quartz, clay minerals 

and other phyllosilicates, and feldspar (e.g. microcline), 

 reworked and incorporated from adjacent basement complexes 

and derived from riverine input. The source of pyrite and 

 barite in some samples may be related to the degradation of 

organic matter during the early stages of diagenesis which 

 provides sulphur for the formation of such minerals in oxic to 

suboxic environments (Schieber 2011; Arndt et al. 2013). 

In pure limestones from Kummer and Fertőrákos, the pre-

dominance of calcite and the sole abundance of clastic mine-

rals (e.g., quartz and clay minerals) are also related to the role 

of erosion and reworking of pre-existing limestones in subaerial 

and shallow water conditions,

 

as a result of reduced marine 

conditions and/or lowered sea level. This scenario is based on 

the deposition of sands, gravels and hydrodynamically con-

trolled “detritic” Leitha limestones by Gilbert-type deltas in 

the Eisenstadt-Sopron Basin (Rostá 1993; Wiedl et al. 2014). 

Limestones from Rosenberg show detrital input signals espe-

cially in the impure limestone samples collected near the con-

tact with underlying and overlying siliciclastics. Magmatic 

activity certainly played a role in affecting carbonate sedimen-

tation in the studied area (Reuter et al. 2012). Mineralogically, 

dolomite formation in those limestones is related to the pre-

sence of intervening tuff layers, which are originally rich in 

Mg, acting as the primary source for dolomitization during 

early diagenesis.  

Geochemistry

Major elements

Generally,  CaO  wt. %  in  the  studied  limestones  varies  in 

 reverse relation to SiO

2

 and Al

2

O

which clearly indicates the 

effect of siliciclastic input and thus detrital dilution of carbo-

nates during deposition. Al

2

O

is used as a proxy for clay 

 content in limestones (Nothdurft et al. 2004), and the average 

contents of 1.2 wt. % in impure limestones from Mannersdorf 

confirms their classification as shale-contaminated (average 

value for siliciclastic-contaminated carbonates is 0.42 wt. %, 

Veizer 1983). The same feature applies at Wöllersdorf with 

Al

2

O

wt. % average of 0.63. On the other hand, pure lime-

stones from Kummer and Fertőrákos do not show any signifi-

cant siliciclastic contamination as indicated by Al

2

O

contents 

below 0.37 wt. %. At Rosenberg, the two samples in contact 

with the underlying and overlying siliciclastics have Al

2

O

contents of 0.7 and 0.8 wt. %, respectively, giving an indica-

tion of the contamination, while the two limestone samples 

further away from the contact and in the middle of the outcrop 

do not show a similar trend and can be classified as pure lime-

stones with an Al

2

O

wt. % average of 0.06. 

Minor elements 

The Leitha limestones exhibit a significant variation in Sr 

content, with a general increase towards the South of the 

 studied Austrian basins (Fig. 8). Although Sr values are lower 

than  the  average  global  carbonate  value  (610  ppm,  Mielke 

1979), the Sr content in the different limestone facies at 

 Mannersdorf is the highest among the various rock types in 

that area with an average of 231 ppm. The Sr/Ca ratio in 

 carbonates is used as a proxy to track the variations in sea-

water Sr/Ca ratio (e.g., Ullmann et al. 2016) and the diagenesis 

(e.g., Ando et al. 2006). The Sr/Ca ratio with an average of 

1/1845 in all the Leitha limestones samples is higher than the 

mean value given by Kulp et al. (1952) which should be less 

than 1/1000 in most types of limestone. The lowest content is 

recorded in Mannersdorf and Wöllersdorf with a maximum 

value of 352 ppm. The Sr is favoured to be incorporated into 

aragonite rather than calcite, but this Sr is lost during diage-

netic transformation of the metastable aragonite into the stable 

calcite (Morse & Mackenzie 1990). 

Two main causes of low Sr content in limestones from 

 Mannersdorf and Wöllersdorf are proposed here. First, the main 

constituents of these limestones are corallincean red algae and 

foraminifers which are, mineralogically, high-Mg calcites 

(Scholle & Ulmer-Scholle 2003), more prone to low Sr con-

tents. Second, the release of Sr to pore water during recrystalli-

zation of calcite (Baker & Bloomer 1988) reduces the Sr content 

of the deposits. Although, additional diagenetic studies should 

be carried out on these localities, we favour the second explana-

tion because Leitha limestones with largely similar composi-

tion, for instance, from the Kummer quarry have higher Sr con-

tents with slight upward increase due to diagenesis (Fig. 9). 

background image

260

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

The highest Sr content is recorded in samples from the Styr-

ian basin (i.e. Rosenberg samples) not only in the limestones, 

but also in the siliciclastics except for two samples. This basin 

suffered volcanism with a wide spectrum of magmatic rocks 

during the Neogene to Quaternary evolution of the Carpathi-

an-Pannonian region (Harangi et al. 2006; Seghedi & Downes 

2011, Ali & Ntaflos 2013). Volcanic activity may have played 

a role in a higher primary Sr content of regional Badenian wa-

ter masses, and thus may be preserved in limestones from the 

Rosenberg quarry. Scleractinian corals (e.g. Porites and Tar-

bellastraea, Riegl & Piller 2000) which are aragonitic (Schol-

le & Ulmer-Scholle 2003) could also account for parts of the 

higher Sr content.

REEs patterns

The total REEs contents in the pure and impure samples 

from Mannersdorf and impure samples from Wöllersdorf and 

their strong positive correlation with terrigenous proxies such 

as Al

2

O

3

 and Zr (Fig. 10) indicates the influence of the detrital 

input during sedimentation on the total REEs content. Zr is 

used as a terrigenous input proxy due to its resistance to 

weathering  and  alteration  processes  (Taylor  &  McLennan 

1985) besides its low abundance in seawater (Boswell & El-

derfield 1988) and the single valency state prevents the effect 

of the changing redox conditions during deposition (Calvert et 

al. 1996). The stratigraphic distribution of the REEs content at 

Mannersdorf does not show a clear relationship between the 

facies type and REEs concentration (Fig. 10). The PAAS nor-

malized REEs patterns of all the Leitha limestones facies are 

flat in shape and display a non-seawater like pattern due to the 

effect of clay admixture. These flat patterns are similar to the 

patterns of shales which represent the original source of the 

REEs in the ocean (Elderfield 1988). There is no strong cor-

relation between REEs content and detrital proxies such as 

Al

2

O

3

  and  Zr  (r=0.27  and  −0.05,  respectively)  in  Kummer 

pure limestones samples supporting the idea that these sam-

ples are not contaminated by siliciclastics and retain the (orig-

inal) seawater signal better. The REEs contents in the studied 

facies types do not vary and have a limited range (Fig. 11). 

Also, REEs patterns in these samples and at Fertőrákos show 

seawater-like shale normalized REEs patterns, characterized 

by HREEs enrichment shown by the low (La/Yb)

N

 ratio (ave-

rage 0.74 for 23 samples), a negative Ce anomaly (average = 

0.6) , and a high (Y/Ho)

mol

 ratio (average = 39). These patterns 

are interpreted as preserving the primary sedimentary environ-

mental signal of marine water under oxic conditions (Haley et 

al. 2004). 

Both REEs pattern types exist in the Rosenberg samples, 

where impure samples in contact with the underlying and 

overlying siliciclastic and contaminated with shale have flat 

non-seawater like patterns with (La/Yb)

of average 1.04 

(n=2), higher REEs content (average 13.4 ppm), lower (Y/Ho)

mol

 

ratio of average 31 and Ce anomaly of 0.84. On the other hand, 

the pure limestones samples are significantly different, 

 displaying seawater-like patterns with HREEs enrichment 

with (La/Yb)

of average 0.54 (n=2), lower REEs content of 

average 6.1 ppm, higher (Y/Ho)

mol

 ratio of average 39 and Ce 

anomaly of 0.7. The latter two samples suggest deposition of 

Fig. 8. Sr content of Leitha limestones from all localities, showing 

increasing values in Eisenstadt-Sopron (Kummer and Fertőrákos) and 

Styrian (Rosenberg) basins compared to the Vienna Basin (Manners-

dorf and Wöllersdorf).

Fig. 9. Stratigraphic distributions of Sr content in the Kummer quarry 

with an increase upward, while Y/Ho ratio is almost the same in all 

samples (colours of facies after Schmid et al. 2001).

background image

261

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

the Rosenberg limestones under oxic conditions whereas more 

impure samples indicate a masking of the primary environ-

mental (seawater) signal by siliciclastic contamination.

To test for the REEs elements pattern source three single- 

component samples were selected (cement, oyster shell, and 

pectinid shell) to be compared to the whole rock chemical 

analyses of the same parental samples. The cement fraction 

(sample M14/43) does not show a diverging total REEs con-

tent compared to the whole rock sample (5.28 vs. 5.73 ppm, 

respectively) but the Ce anomaly is different with a lower  value 

of 0.62 in the cement fraction compared to 0.86. The REEs 

pattern in the cement fraction is more similar to the seawater- 

like pattern than the sample itself, with HREEs enrichment 

(i.e. (La/Yb)

N

 value of 0.2), this ratio is 0.52 in the sample 

(Fig. 12a) and higher Y/Ho ratio of 27 and 23 for the whole 

sample. These relations give a hint of oxic marine conditions 

and cement precipitation from seawater trapped in pores 

during early diagenesis. 

In samples from Kummer the total REEs content is lower in 

the biogenic fractions compared to the whole rock sample 

(3.22 and 2.78 ppm vs. 12.96 and 13.8 ppm in the whole rock 

samples). The REEs patterns in the whole rock samples are 

better defined than in the biogenic material (Fig. 12b) which 

can be explained by the lack of some 

elements such as Dy and Tm and the 

low elements contents of these frac-

tions. The Ce anomaly is lower in the 

oyster shell (0.5 vs. 0.65 in the whole 

rock sample), a reverse situation to 

the pectinid shell where it is higher 

than the whole sample (0.59 vs. 0.52, 

respectively). Also, the Y/Ho ratio 

shows no clear trend from the shells 

versus the whole rock samples. 

Ce anomaly

The  Ce anomalies in limestones 

from Mannersdorf and Wöllersdorf 

are close to 1, and thus differ from 

seawater values (0.1– 0.4). This can be 

explained by (1) the minor presence 

of clay minerals (from detrital input) 

in the pure and impure samples which 

is identified in all samples based on 

XRD analysis but to some extent by 

using SEM; (2) the release of LREEs 

due to the degradation of the organic 

matter including Ce in the few centi-

metres depth of the sediment column 

below the sea floor. Haley et al (2004) 

showed that in oxic environments the 

remineralization of organic coatings 

results in a balancing of the Ce ano-

maly to the value of the source mate-

rial. Moreover, the presence of pyrite 

and barite supports such a conclusion by postulating that the 

organic matter was the source for sulphur (Schieber 2011; 

Arndt et al. 2013). The Ce anomaly record is different at 

 Kummer and Fertőrákos, displaying considerably lower values 

(average  0.62)  than  at  Mannersdorf  and  Wöllersdorf.  This 

 indicates primary seawater values and conforms to an inter-

pretation of an oxic depositional environment, characterized 

by the lack of detectable detrital input which may have been 

removed during reworking and redeposition of the predomi-

nating detrital limestones from these localities. At Rosenberg, 

two ranges of Ce anomalies exist which indicate significant 

detrital input in impure limestones resulting in higher values 

whereas the lower values in pure limestones record primary 

oxic conditions.   

Y/Ho ratio

The ratio of yttrium (Y) versus holmium (Ho) in the studied 

limestones samples is nearly similar in all the studied locali-

ties. The values are slightly higher in samples exhibiting low 

or  no  siliciclastic  input  (≥ 35  in  relatively  pure  limestones) 

compared  to  samples  with  siliciclastic  admixture  (29 –35  in 

impure limestones). The values indicate in general deposition 

Fig. 10. Stratigraphic distributions of Al

2

O

3

, REEs and Zr at Mannersdorf with strong correlation 

between the three parameters (colours of facies after Wiedl et al. 2012).

background image

262

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

under oxic conditions. Due to the simi-

larity in ionic radii of Y and Ho the ex-

pected oceanic distributions are closely 

similar (Zhang et al. 1994; Bau et al. 

1995). Nevertheless, their different geo-

chemical behaviour results in strong 

frac tionation between the two elements. 

The former authors deduced that the Y/Ho 

ratios in seawater are approximately 

two times higher than the chondritic and 

shale ratios. The ratio is also affected 

by redox conditions where it decreases 

from 102 in oxic to 67 in anoxic waters 

as a consequence of preferential sorp-

tion of Ho with respect to Y on Fe- and 

Mn-oxyhydroxide particles that even-

tually dissolve under anoxic conditions 

(Bau et al. 1997). The fractionation 

takes place predominately in the ocean 

(Nozaki et al. 1997). 

Interpretation of controlling factors

The evidence derived from the 

 

mineralogical compositions and the 

above mentioned geochemical proxies 

account for the influence of several 

factors controlling the sediment geo-

chemistry, notably the trace elements 

and RREs. In principle the following 

main factors govern the composition of 

the Miocene limestones: (1) detrital 

admixture from siliciclastics and volcaniclastics, (2) palaeo-

environmental signals from primary seawater composition 

and redox states, (3) diagenetic overprint. Detritally-contami-

nated limestones show clear devia tions from the normal sea-

water conditions while the purer limestones point more clearly 

to the primary depositional environments if not strongly affec-

ted by diagenesis and later alteration.

Diagenetic evolution 

There are significant differences in the diagenetic evolution 

between the studied localities. Chemical elements such as Sr, 

Na and Ba can be used as indices for the magnitude of diage-

netic overprint, because these elements are generally incorpo-

rated during deposition and redistributed during diagenesis 

Fig. 11. Stratigraphic distributions of Al

2

O

3

, REEs and Zr at Kummer with weak positive correla-

tion between REEs and Al

2

O

3

 and no correlation between REEs and Zr (colours of facies after 

Schmid et al. 2001).

Fig. 12. a — REEs patterns of sample M14/43 from Mannersdorf compared to cement fraction (M14/43Ce) of the same sample, b — REEs 

patterns of the biotic fractions (oyster, pectinid shell) compared to the whole samples from Kummer.

background image

263

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

(e.g., Land & Hoops 1973; Brand & Veizer 1980; Morrison & 

Brand 1986; Nothdurft et al. 2004; Webb et al. 2009). Here, 

the enrichment and/or depletion of these elements are used 

only to evaluate the relative degree of diagenesis in the studied 

limestones. 

Intensive meteoric diagenesis, as expected and demonstra-

ted for these carbonate platforms (Dullo 1983) leads to deple-

tion in elements such as Sr, Na and Ba. The overall Sr content 

varies considerably, with low amounts in all limestone sam-

ples from Mannersdorf and Wöllersdorf (averages 221 ppm in 

Mannersdorf and 197 ppm in Wöllersdorf), and increasing 

from Kummer and Fertőrákos (average 570 ppm) to Rosen-

berg (average 1271 ppm) as shown in Figure 8. This trend 

 indicates in general that Mannersdorf and Wöllersdorf lime-

stones suffered more intensive diagenesis compared to the 

other localities. A similar trend in Na and Ba supports this 

idea. Mannersdorf and Wöllersdorf show the lowest values of 

these elements (Na not detected in most of the samples, ave-

rage  Ba  of  15.6  ppm),  followed  Kummer  and  Fertőrákos 

 (average 212.9 ppm Na, average 20.6 ppm Ba) and with the 

highest values at Rosenberg (average 315.3 ppm Na, average 

43.5 ppm Ba). The above mentioned values for Na and Ba are 

derived only from pure limestone samples without visible con-

tamination by detrital input to avoid mixing with the signal 

coming from clay minerals. The minimal diagenetic overprint 

on limestones from the Rosenberg quarry was previously also 

reported on the basis of stable isotopes studies on shell mate-

rial (Bojar et al. 2004). REEs patterns are in accordance with 

this interpretation, indicating stronger diagenetic overprint 

 influencing the palaeoenvironmental signal considerably at 

Mannersdorf and Wöllersdorf, whereas at Kummer and Rosen-

berg localities, the original seawater signal is largely preserved. 

Provenance 

Geological setting and relations to basement rocks at the 

studied localities gives the opportunity to evaluate the applica-

bility of various elemental ratios to deduce the possible sources 

of siliciclastic input during shallow-water carbonate deposi-

tion. The distributions of immobile elements are used to inter-

pret sediment sources and hinterland geology, for example, La 

and Th are enriched in acidic rocks, while Sc, Cr and Co are 

enriched in basic rocks (McLennan et al. 1983, 1990; Cullers 

2000; Madhavaraju et al. 2010). The resulting elemental ratios 

are more affected by the composition of source rocks rather 

than the depositional environment (Cullers 1988, 1994).  

La/Sc, La/Co, Th/Sc, Th/Co and Th/Cr elemental ratios are 

used to evaluate and classify the detrital input sources. Ele-

mental ratios of limestone samples from Mannersdorf and 

Kummer containing considerable amounts of detrital materials 

and siliciclastic samples from the former localities and Rosen-

berg are shown in Table 4. The Mannersdorf samples exhibit  

a wide range of elemental ratios indicating intermediate to 

 silicic igneous rock sources, while the Kummer and Rosen-

berg samples plot in the range of silicic rocks (Figs. 13, 14). 

This conforms to the mainly silicic source material present in 

the Austroalpine basement units of the surrounding Alpine 

units (e.g., Schuster & Nowotny 2015), namely (meta) grani-

tes, gneisses and mica schists.

Palaeoenvironments

Considering pure (with respect to detrital contamination) 

and diagenetically relatively unaltered limestones, Ce and Eu 

anomalies can be used as proxies to infer palaeoenvironments, 

essentially indicating the redox state during limestone deposi-

tion.  In  addition,  the  Mn*  values  of  the  Leitha  limestone 

 samples are overall positive indicating deposition under oxic 

conditions. On the other hand, some siliciclastic samples from 

the three localities show negative Mn* values which point to 

(more) reducing depositional environments, in accordance 

with intermittently low-oxygen depositional sites and dysoxia 

events inferred for certain parts of the Leitha limestone 

 platforms (Schmid et al. 2001). 

Weathering 

Weathering indices, related to source rock weathering, were 

evaluated here to test the applicability of these methods and to 

evaluate the transfer of the weathering signal from the source 

to the carbonates. The relative proportions of mobile element 

oxides (CaO, Na

2

O and K

2

O) to immobile element oxides 

Table 4: Range of elemental ratios of Mannersdorf, Kummer and Rosenberg compared to felsic rocks, mafic rocks and PAAS.

Mannersdorf impure 

Leitha limestones 

n=4

Mannersdorf

siliciclastics 

n=8

Kummer 

marls

n=4

Kummer

siliciclastics 

n=3

Rosenberg

siliciclastics 

n=7

Felsic rocks

a

Mafic rocks

a

PAAS

b

La/Sc

0.49–4.4

0.62–3.21

2.87–3.85

2.93–2.95

2.13–3.86

2.5–16.3

0.43–0.86

2.4

La/Co

1.13–3.72

1.12–3.69

6.82–8.63

2.3–5.21

1.9–3.96

1.8–13.8

0.14–0.38

1.65

Th/Sc

0.07–1.2

0.35–1.82

0.87–1.1

0.86–1.03

0.53–1.23

0.84–20.5

0.05–0.22

0.90

Th/Co

0.17–0.94

0.27–1.91

2–4.2

0.78–1.82

0.54–1.61

0.67–19.4

0.04–1.4

0.63

Th/Cr

0.06–0.12

0.07–0.15

0.09–0.11

0.11

0.02–0.07

0.13–2.7

0.018–0.046

0.13

Cr/Th

8–17

7–186

9–11

9

15–47

4.0–15.0

25–500

7.53

Mn*

0.77–1.06

−0.72–0.21

0.92–1.47

−0.52–0.09

−0.32–0.84

CIA

71.4–72.9

74.7–91.1

83.1–84

83.2–83.6

64.2–72.8

CIW`

93.7–99.1

92.6–99.5

98.8–98.9

99

84.6–92.9

a

 Cullers (1994, 2000); Cullers & Podkovyrov (2000); Cullers et al. (1988)

b 

Taylor & McLennan (1985)

background image

264

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

(Al

2

O

3

) were used to determine the degree of chemical weathe-

ring, in our study related to weathering of the siliciclastic 

 admixture in impure limestones; pure limestones cannot be 

used for these methods because of the low contents of the 

 respective  elements. 

The Mannersdorf impure and Kummer marly limestones 

and siliciclastics show uniform, moderate to high degrees of 

chemical weathering with CIA values ranging from 71 to 91 

(Table 4). Using the A-CN-K (Al

2

O

3

– CaO*+ Na

2

O –K

2

O) 

 ternary plot (Fig. 15), these values approach the muscovite 

and illite fields which have CIA values of 75 and 75 to 85, 

 respectively  (Nesbitt  &  Young  1982).  Such  values  resulted 

from the transformation of feldspars (from felsic source rocks 

of the Austroalpine basement) into clay minerals along the 

transport path and finally deposited in limestones or as inter-

calated layers. In comparison, the Rosenberg samples show 

lower values (64–73, Fig. 15) with a low degree of chemical 

weathering indicating most probably the influence of the 

 primarily volcano-siliciclastic input in the Styrian basin with 

short transport distance und thus minor weathering. The CIW` 

shows a similar trend to CIA among the studied samples   

(Table 4). In contrast to impure limestones, the application of 

the CIA index in pure limestones (i.e. low Al

2

O

3

 content below 

0.42 wt. %, for example, in the Kummer limestones) proves 

impractical because the obtained values are attributed to low 

contents of both mobile and immobile oxides rather than a par-

ticular mineralogy.

Provenance versus palaeoenvironmental signals

Geochemical data from varying sites of Central Paratethys 

Middle Miocene carbonate platform limestones allow a testing 

of signals according to provenance and/or palaeo-environ-

mental significance. Samples investigated range from pure 

limestones (Al

2

O

3

 < 0.42 wt. %, CaCO

3 

>

99 wt. %) to impure 

limestones (Al

2

O

3

 > 0.42 wt. %, CaCO

3 

< 80 wt.

.

 %, impurities 

from quartz, pyrite and evaporites) to siliciclastics (sands, 

clays) present as interlayers and below and/or above the lime-

stones. Signals can thus be separated accordingly into those 

stemming from siliciclastic admixture, and so mainly influen-

ced by provenance signals carried by the (minor) siliciclastic 

fraction, and palaeoenvironmental signals representing origi-

nal sea-water composition — although this signal may also be 

indirectly influenced by the surrounding geology via disso-

lution and locally deviating water chemistry within the basins 

investigated.

As expected, the provenance signal is the main influence on 

various parameters as deduced from similar values going from 

impure to more pure limestones. Still, classical provenance 

indicators like immobile elemental ratios, CIA and CIW` are 

Fig. 13. Th/Co vs La/Sc plot of impure Leitha limestones samples 

containing detrital materials and siliciclastic samples from Manners-

dorf (intermediate to silicic), Kummer and Rosenberg (silicic) 

 quarries (after Cullers 2002).

Fig. 14. La–Th–Sc ternary plot of impure Leitha limestones samples 

containing detrital materials and siliciclastic samples from Manners-

dorf (intermediate to silicic), Kummer and Rosenberg (silicic) 

 quarries (after Cullers 2002).

Fig. 15. A-CN-K ternary plot with the chemical index of alteration 

(CIA) of impure limestones, marls and siliciclastics of Mannersdorf, 

Kummer and Rosenberg quarries, mineral fields are represented by 

kaolinite (Ka), chlorite (Chl), illite (Il), muscovite (Mu), smectite 

(Sm), plagioclase (Pl) and K-feldspars (Kfs).

background image

265

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

applicable and interpretable even in impure limestones such as 

those from Mannersdorf, which contain considerable amounts 

of Al

2

O

3

 up to 2.5 wt. % and marly samples from the Kummer 

quarry. The high values of the CIA indicate enhanced  chemical 

weathering in the hinterland due to warm and humid 

 palaeo climate, and thus compatible with increased detrital 

 input  into  the  basins  (Nesbitt  &  Young  1982;  Zhao  &  

Zheng 2014). 

Furthermore, the stratigraphic evolution at Mannersdorf 

with a significant gradual decline in the detrital content (e.g., 

Al

2

O

3

, Fig. 10) upwards coincides with the regional palaeo-

climatic changes in the Central Paratethys region in which  

a tropical (Wiedl et al. 2012) to warm-temperate climate 

(Mid-Miocene Climatic Optimum) was followed by gradual 

temperature decline (Ivanov et al. 2002; Hohenegger et al. 

2009; Kováčová et al. 2009). Provenance signals are derived 

clearly as a function of the presence of detrital-siliciclastic 

material which, by itself, is also a function of weathering, 

 palaeoclimate and palaeohydrology. The alternative explana-

tion of enhanced detrital input due to tectonic subsidence and 

hinterland uplift is unlikely because of the significant decrease 

in tectonic activity during the middle to late Badenian (Lee & 

Wagreich 2017).

In contrast, only pure limestones with Al

2

O

3

 below 0.23 wt. % 

give reliable information about their specific depositional 

 environment and environmental conditions in general. Thus, 

palaeo-environmental signals are relatively hard to decipher 

especially with increasing shale contamination into the lime-

stones. However, some indices may still work in impure lime-

stones with Mn*, Ce/Ce* and Eu/Eu* especially useful in the 

case of separating oxic from low-oxic/anoxic palaeoenviron-

ments. All of these indices generally indicate deposition of the 

Leitha limestones under oxic conditions; only exceptional 

 cases indicate low-oxic/anoxic conditions (e.g., negative Mn* 

values), also proven by faunal evidence such as the preserva-

tion of fish remains (Schmid et al. 2001). Other proxies show 

normal marine conditions as indicated by “normal” seawater 

values without further specifications. 

Conclusions 

The five studied Middle Miocene Leitha limestone occur-

rences exhibit various mineralogical, geochemical and environ-

mental features. These features can be summarized as follows:

•  The limestones were classified into pure limestones with 

Al

2

O

3

  content  below  0.42  wt. %  and  the  absence  of  clay 

minerals while impure limestones have higher Al

2

O

3

 and  

a detectable amount of clay minerals induced by the detrital 

input. Both limestone types exist in the studied localities 

with variable abundances.

•  Limestones from Mannersdorf and Wöllersdorf (southern 

Vienna Basin) are influenced by detrital input during depo-

sition which varies between intermediate to silicic sources 

in origin. This input appears in the mineralogy of the lime-

stones which contain quartz and clay minerals in most of the 

samples. The input decreased with respect to age, probably 

due to the palaeoclimatic change to a slightly cooler climate 

during the Middle Miocene. Geochemically, the marine 

 signal deduced by REEs patterns, Ce anomaly and Y/Ho 

 ratio is masked by the detrital input in most of the samples. 

No clear relationship between facies types and geochemical 

characteristics can be deduced due to both significant 

 detrital input and diagenetic influence. 

•  The  Leitha  limestones  from  Kummer  and  Fertőrákos 

 (Eisenstadt–Sopron Basin) are purer with regard to the 

 detrital input in which no clay minerals were detected. This 

criterion makes limestones from these localities excellent 

tracers for the primary depositional environments; at the 

same time it inhibits clear determination of provenance. 

REEs patterns, Ce anomaly and Y/Ho ratio indicate the 

deposition of limestones under oxic shallow marine condi-

tions. The origin of the siliciclastics in the marly facies and 

in clay layers in Kummer quarry is from silicic igneous 

rocks. The palaeoenvironments of these siliciclastics were 

oxygen-deficient.  

•  The limestones from the Rosenberg quarry (Styrian Basin) 

were influenced by volcano-siliciclastic events during the 

Badenian, displayed by quartz, clay minerals, feldspars and 

dolomite especially in samples near the contact with the 

 siliciclastics. The low degree of chemical weathering evi-

denced by low CIA values also assures the nearby and less 

altered volcano-siliciclastic source of the detrital input.  

The marine signal of these facies is not obvious due to this 

influence, while in the facies away from the contact with 

siliciclastics oxic shallow marine conditions represent the 

depositional environment. The source of the siliciclastics at 

Rosenberg is silicic igneous rocks for both underlying and 

overlaying ones. 

In general, provenance signals are detectable going from 

(siliciclastic) sediments to impure limestones which contain 

Al

2

O

3

 down to 2.5 wt. %. The source rock signal is thus the 

main influence on limestone geochemistry in impure lime-

stones, and classical provenance indicators like immobile 

 elemental ratios, CIA and CIW` are applicable and interpretable 

even in slightly impure limestones. Proxies for reconstructing 

environmental conditions during deposition are mainly indices 

that may separate oxic from low-oxic/anoxic palaeoenviron-

ments like Mn*, Ce/Ce* and Eu/Eu* which worked reliably 

going from impure to pure limestones. 

Acknowledgements:  We thank the quarry companies, 

 especially Lafarge Zementwerke GmbH (former Lafarge- 

Perlmoser) for making available sampling in the active quarry 

sites, and for hospitality and support for the work. Thanks go 

to the Egyptian Cultural Affairs & Missions Sector for provi-

ding a PhD scholarship for the first author. Field work was 

partly financed by UNESCO-IUGS IGCP 609 and the Austrian 

Academy of Sciences. We also thank Shahid Iqbal and 

 Wolfgang Knierzinger for assistance during field work, Sabine 

Hruby-Nichtenberger and Maria Meszar for sample prepara-

tion and lab work.

background image

266

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

References

Ali Sh. & Ntaflos Th. 2013: Petrogenesis and mantle source charac-

teristics of Quaternary alkaline mafic lavas in the western 

 Carpathian–Pannonian Region, Styria, Austria. Chem. Geol. 

337–338, 99–113. 

Anderson P.E., Benton M.J., Trueman C.N., Paterson B.A. & Cuny G. 

2007: Palaeoenvironments of vertebrates on the southern shore 

of Tethys: the nonmarine early cretaceous of Tunisia. Palaeo-

geogr. Palaeoclimatol. Palaeoecol. 243, 118–131. 

Ando A., Kawahata H. & Kakegawa T. 2006: Sr/Ca ratios as indicators 

of varying modes of pelagic carbonate diagenesis in the ooze, 

chalk and limestone realms. Sediment. Geol. 191, 37–53.

Arndt S., Jørgensen B.B., LaRowe D.E., Middelburg J.J., Pancost R.D. 

&  Regnier  P.  2013:  Quantifying  the  degradation  of  organic 

 matter in marine sediments: A review and synthesis. Earth Sci. 

Rev. 123, 53–86. 

Azmy K., Brand U., Sylvester P., Gleeson S.A., Logan A. & Bitner 

M.A. 2011: Biogenic and abiogenic low-Mg calcite (bLMC and 

aLMC): evaluation of seawater-REE composition, water masses 

and carbonate diagenesis. Chem. Geol. 280, 180–190.  


Baker P.A. & Bloomer S.H. 1988: The origin of celestite in deep-sea 

carbonate sediments. Geochim. Cosmochim. Acta 52, 335–340. 

Balogh K., Ebner F., Ravasz C., Herrmann P., Lobitzer H. & Solti G. 

1994: K/Ar-Alter tertiärer Vulkanite der südöstlichen Steiermark 

und des südlichen Burgenlandes. In: Lobitzer H., Császár G. & 

Daurer A. (Eds.): Jubiläumsschrift 20 Jahre Geologische Zusam-

menarbeit Österreich–Ungarn. Geol. Bundesanst. Vienna, 55–72.

Bau M. 1991: Rare-earth element mobility during hydrothermal and 

metamorphic fluid- rock interaction and the significance of the 

oxidation state of europium. Chem. Geol. 93, 3, 219–230. 

Bau  M.  &  Dulski  P.  1999:  Comparing  yttrium  and  rare  earths  in 

 hydrothermal fluids from the Mid-Atlantic Ridge: implications 

for Y and REE behaviour during near-vent mixing and for the  

Y/Ho ratio of Proterozoic seawater. Chem. Geol. 155, 1, 77–90. 

Bau M. & Möller P. 1992. Rare earth element fractionation in meta-

morphogenic hydrothermal calcite, magnesite and siderite. 

 Mineral. Petrol. 45, 3–4, 231–246. 

Bau M., Dulski P. & Möller P. 1995: Yttrium and holmium in South 

Pacific seawater: vertical distribution and possible fractionation 

mechanisms. Chem.  Erde  55, l–15.

Bau  M.,  Möller  P.  &  Dulski  P.  1997:  Yttrium  and  lanthanides  in 

 eastern Mediterranean seawater and their fractionation during 

redox-cycling. Mar. Chem. 56, 123–131. 

Bau  M.,  Balan  S.,  Schmidt  K.  &  Koschinsky A.  2010:  Rare  earth 

 elements in mussel shells of the Mytilidae family as tracers for 

hidden and fossil high-temperature hydrothermal systems. Earth 

Planet. Sci. Lett. 299, 310–316. 

Bojar A-V.,  Hiden  H.,  Fenninger A.  &  Neubauer  F.  2004:  Middle 

Miocene seasonal temperature changes in the Styrian basin, 

Austria, as recorded by the isotopic composition of pectinid and 

brachiopod shells. Palaeogeogr. Palaeoclimatol. Palaeoecol. 

203, 95–105.

Boswell S.M. & Elderfield H. 1988: The determination of zirconium 

and hafnium in natural waters by isotope dilution mass spectro-

metry. Mar. Chem. 25, 197–210. 

Brand U. & Veizer J. 1980: Chemical diagenesis of a multicomponent 

carbonate system-1: trace elements. J. Sediment. Petrol. 50, 

1219–1236.

Calvert  S.E.,  Burtin  R.M.  &  Ingall  E.D.  1996:  Influence  of  water 

 column anoxia and sediment supply on the burial and preserva-

tion of organic carbon in marine shales. Geochim.  Cosmochim. 

Acta 60, 1577–1593. 

Campbell A.C., Palmer M.R., Klinkhammer G.P., Bowers T.S., 

 

Edmond J.M., Lawrence J.R., Casey J.F., Thompson G., 

 Humphris S., Rona R. & Karson J.A. 1988: Chemistry of hot 

springs on the Mid- Atlantic Ridge. Nature (London) 335, 514-519. 

Chittleborough D.J. 1991: Indices of weathering for soils and palaeo-

sols formed on silicate rocks. Austr. J. Earth Sci. 38, 115–120. 

Craddock  P.R.,  Bach  W.,  Seewald  J.S.,  Rouxel  O.J.,  Reeves  E.  & 

Tivey M.K. 2010: Rare earth element abundances in hydrother-

mal fluids from the Manus Basin, Papua New Guinea: indicators 

of sub-seafloor hydrothermal processes in back-arc basins. 

 Geochim. Cosmochim. Acta 74, 5494–5513. 

Cullers R.L. 1988: Mineralogical and chemical changes of soil and 

stream sediment formed by intense weathering of the Danberg 

granite, Georgia, USA. Lithos 21, 301–314. 

Cullers R.L. 1994: The controls on the major and trace element 

 variation of shales, siltstones and sandstones of Pennsylvanian- 

Permian age from uplifted continental blocks in Colorado to 

platform sediment in Kansas, USA. Geochim. Cosmochim. Acta 

58, 4955–4972.

 Cullers R.L. 2000: The geochemistry of shales, siltstones and sand-

stones of Penn- sylvanianePermian age, Colorado, U.S.A.: 

 implications for provenance and metamorphic studies. Lithos 51, 

305–327. 

Cullers R.L. 2002: Implications of elemental concentrations for 

prove nance, redox conditions, and metamorphic studies of 

shales and limestones near Pueblo, CO, USA. Chem. Geol. 191, 

305–327.

Cullers R.L. & Podkovyrov V.N. 2000: Geochemistry of the Meso-

proterozoic Lakhanda shales in southeastern Yakutia, Russia: 

implications for mineralogical and provenance control, and 

 recycling.  Precambrian Res. 104, 77–93.

Cullers R.L., Basu A. & Suttner L. 1988: Geochemical signature of 

provenance in sand-size material in soils and stream sediments 

near the Tobacco Root batholith, Montana, USA. Chem. Geol. 

70, 335–348.

de Baar H.J.W., Bacon M.P. & Brewer P.G. 1983: Rare-earth distribu-

tions with a positive Ce anomaly in the western North Atlantic 

Ocean. Nature 301, 324–327. 

de Baar H.J.W., Bacon M.P., Brewer P.G. & Bruland K.W. 1985: Rare 

earth elements in the Pacific and Atlantic Oceans. Geochim. 

Cosmochim. Acta 49, 1943–1959.

Dubinin A.V. 2004: Geochemistry of rare earth elements in the ocean. 

Lithol. Min. Resour. 39, 289–307.  

Dullo W.C. 1983: Fossildiagenese im miozänen Leitha-Kalk der 

 Paratethys von Österreich: Ein Beispiel für Faunenverschiebung 

durch Diageneseunterschiede. Facies 8, 1–112. 

Ebner  F.  &  Sachsenhofer  R.F.  1995:  Palaeogeography,  subsidence 

and thermal history of the Neogene Styrian basin (Pannonian 

basin system, Austria). Tectonophysics 242, 133–150. 

Elderfield H. 1988: The oceanic chemistry of the rare-earth elements. 

Philos. Trans. R. Soc. London,325, 105–126.

Elderfield H. & Greaves M.J. 1982: The rare earth elements in sea-

water. Nature 296, 214–219. 

Fenninger A. & Hubmann B. 1997: Palichnologie an der Karpatium/

Badenium-Grenze des Steirischen Tertiärbeckens (Österreich). 

Geol.-Paläontol. Mitt. Innsbruck, 22, 71–83.

Haley  B.A.,  Klinkhammer  G.P.  &  McManus  J.  2004:  Rare  earth 

 elements in pore waters of marine sediments. Geochim. Cosmo-

chim. Acta 68, 1265–1279.

Garbelli C., Angiolini L., Brand U., Shen S.Z., Jadoul F., Posenato R., 

Azmy K. & Cao C.Q.  2016: Neotethys seawater chemistry and 

temperature at the dawn of the end Permian mass extinction. 

Gondwana Res. 35, 272–285. 

Handler R., Ebner F., Neubauer F., Hermann S., Bojar A.-V.,   

Hermann  S.  2006.  40Ar/39Ar  dating  of  Miocene  tuffs  from 

 Styrian part of the Pannonian Basin: an attempt to refine the 

 basin  stratigraphy.  Geol. Carpath. 57, 483–494. 

Haq B.U., Hardenbol J. & Vail P.R. 1987: Chronology of fluctuating 

sea levels since the Triassic. Science 235, 1156–1166. 

background image

267

GEOCHEMISTRY AND PROVENANCE STUDY OF THE LEITHA LIMESTONES, CENTRAL PARATETHYS

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

Harangi S., Downes H. & Seghedi I. 2006: Tertiary-Quaternary sub-

duction processes and related magmatism in the Alpine-Medi-

terranean region. In: Gee D.G. & Stephenson R.A. (Eds.): Euro-

pean Lithosphere Dynamics. Geol. Soc. London Memoir 32, 

167–190.

Harnois L. 1988: The CIW index: a new chemical index of weathe-

ring. Sediment. Geol. 55, 319–322. 

Harzhauser  M.  &  Piller  W.E.  2010:  Molluscs  as  a  major  part  of 

 

subtropical shallow-water carbonate production-an example 

from a Middle Miocene oolite shoal (Upper Serravallian, 

Austria). Spec. Publ. Int. Ass. Sed. 42, 185–200. 

Hedberg H.D. 1976: International Stratigraphic Guide. John Wiley, 

New York, 1–200.

Hinkley T.K. & Tatsumoto M. 1987: Metals and isotopes in Juan de 

Fuca Ridge hydrothermal fluids and their associated solid mate-

rials. J. Geophys. Res. 92, 1943–1959.  


Hohenegger J. & Wagreich M. 2012: Time calibration of sedimentary 

sections based on isolation cycles using combined cross-correla-

tion: dating the gone Badenian stratotype (Middle Miocene, 

 Paratethys, Vienna Basin, Austria) as an example. Int. J. Earth 

Sci. (Geol. Rundsch.) 101, 339–349.  

Hohenegger J., Ćorić S., Khatun M., Pervesler P., Rögl F., Rupp C., 

Selge A.,  Uchman A. & Wagreich M. 2009: Cyclostratigraphic 

dating in the Lower Badenian (Middle Miocene) of the Vienna 

Basin (Austria): the Baden-Sooss core: Int. J. Earth Sci. (Geol. 

Rundsch.) 98, 915–930. 

Hohenegger J., Ćorić S. & Wagreich M. 2014: Timing of the Middle 

Miocene Badenian Stage of the Central Paratethys. Geol. 

 Carpath. 65, 1, 55–66. 

Hongo Y.  &  Nozaki Y.  2001:  Rare  earth  element  geochemistry  of 

 hydrothermal deposits and Calyptogena shell from the Iheya 

Ridge vent field, Okinawa Trough. Geochem. J.  35,5,  347–354.   

Ivanov D., Ashraf A.R., Mosbrugger V. & Palamarev E. 2002:  Palyno- 

logical evidence for Miocene climate change in the Forecarpathian 

Basin (Central Paratethys NW Bulgaria). Palaeogeogr. Palaeo-

climatol. Palaeoecol. 178, 19–37.

Johannesson  K.H.,  Telfeyan  K.,  Chevis  D.A.,  Rosenheim  B.E.  & 

Leybourne M.I. 2014: Rare earth elements in stromatolites-1. 

Evidence that modern terrestrial stromatolites fractionate rare 

earth elements during incorporation from ambient waters.  In: 

Evolution of Archean Crust and Early Life. Springer Nether-

lands, 385–411. 

Kim J.H., Torres M.E., Haley B.A., Kastner M., Pohlman J.W., Riedel 

M. & Lee Y.J. 2012: The effect of diagenesis and fluid migration 

on rare earth element distribution in pore fluids of the northern 

Cascadia accretionary margin. Chem. Geol. 291, 152–165. 

Kocsis L., Trueman C.N. & Palmer M.R. 2010: Protracted diagenetic 

alteration of REE contents in fossil bioapatites: direct evidence 

from Lu–Hf isotope systematics. Geochim. Cosmochim. Acta 

74, 21, 6077–6092. 

Kováč M., Andreyeva-Grigorovich A., Bajraktarević Z., Brzobohatý 

R., Filipescu S., Fodor L., Harzhauser M., Nagymarosy A., 

Oszczypko  N.,  Pavelić  D.,  Rögl  F.,  Saftić,  B.,  Sliva  L.  & 

 Studencka B. 2007: Badenian evolution of the Central Para-

tethys Sea: paleogeography, climate and eustatic sea level 

changes. Geol. Carpath. 58, 579–606.

Kováčová  P.,  Emmanuel  L.,  Hudáčkova  N.  &  Renard  N.  2009: 

 Central Paratethys paleoenvironment during the Badenian (Mid-

dle Miocene): evidence from foraminifera and stable isotope 

(∂

13

C  and  ∂

18

O) study in the Vienna Basin (Slovakia). Int. J. 

Earth Sci. (Geol. Rundsch.) 98, 1109–1127. 

Kröll A. 1988: Reliefkarte des prätertiären/Untergrundes. In: Kröll 

A., Flügel H.W., Seiberl W., Weber F., Walach G. & Zych D. 

(Eds.):  Erläuterungen  zu  den  Karten  über  den  prätertiären 

 Untergrund  des  Steirischen  Beckens  und  der  Südburgen län-

dischen Schwelle. Geologische Bundesanstalt, Vienna,  16–20.

Kulp J.L., Turckian K. & Boyd D.W. 1952: Sr content of limestones 

and fossils. Geol. Soc. Am. Bull. 63, 701–716. 

Land L.S. & Hoops G.K. 1973: Sodium in carbonate sediments and 

rocks: a possible index to the salinity of diagenetic solutions.  

J. Sediment. Petrol. 43, 614–617.

Lankreijer  A.,  Kováč  M.,  Cloetingh  S.,  Pitoňák  P.,  Hlôška  M.  & 

 Biermann C. 1995: Quantitative subsidence analysis and forward 

modelling of the Vienna and Danube basins: thin-skinned versus 

thick skinned extension. Tectonophysics 252, 433–451. 

Laskarev V.N. 1924: Sur les equivalentes du Sarmatien supérieur en 

Serbie. Recueil de traveaux ofert a M. Jovan Cvijic par ses amis 

et collaborateurs, 73–85.

Lee E.Y. & Wagreich M. 2017: Polyphase tectonic subsidence evolution 

of the Vienna Basin inferred from quantitative subsidence analysis 

of the northern and central parts. Int. J. Earth Sci. 106, 687–705. 

Madhavaraju J., González-León C.M., Lee Yong Il., Armstrong- 

Altrin J.S. & Reyes-Campero L.M. 2010: Geochemistry of the 

Mural Formation (Aptian–Albian) of the Bisbee Group,  Northern 

Sonora, Mexico. Cretaceous Res. 31, 400–414. 

McLennan S.M., Taylor S.R. & Eriksson K.A. 1983: Geochemistry of 

Archean shales from the Pilbara Supergroup, Western Australia. 

Geochim. Cosmochim. Acta 47, 1211–1222. 

McLennan S.M., Taylor S.R., McCulloch M.T. & Maynard J.B. 1990: 

Geochemistry and Nd-Sr isotopic composition of deep-sea 

 turbidites: Crustal evolution and plate tectonic associations. 

Geochim. Cosmochim. Acta 54, 2014–2050.

McLennan S.M., Hemming S., McDaniel D.K. & Hanson G.M. 1993: 

Geochemical approaches to sedimentation, provenance, and tec-

tonics.  In:  Johnsson  M.J.  &  Basu  A.  (Eds.):  Processes  Con-

trolling the Composition of Clastic Sediments. Geol. Soc. Am., 

Spec. Pap. 284, 21–40.  

Michard A., Albarède  F.,  Michard  G.,  Minster  J.F.  &  Charlou  J.L. 

1983: Rare earth elements and uranium in high-temperature 

solutions from East Pacific Rise hydrolhermal vent field (13º N). 

Nature (London), 303, 795–797. 

Mielke J.E. 1979: Composition of the Earth’s crust and distribution of 

the elements. In: Siegel F.R. (Ed.), Review of Research on  Modern 

Problems in Geochemistry. UNESCO Report, Paris, 13–37.

Morrison  J.O.  &  Brand  U.  1986:  Geochemistry  of  recent  marine 

 invertebrates.  Geoscience Canada, 13, 237–254. 

Morse J.W. & Mackenzie F.T. 1990: Geochemistry of Sedimentary 

Carbonates. Elsevier, Amsterdam, 1–707.

Nesbitt H.W. & Young G.M. 1982: Early Proterozoic climates and 

plate motions inferred from major element chemistry of lutites. 

Nature 299, 715–717. 

Neuhuber S., Gier S., Hohenegger J., Wolfgring E., Spötl C., Strauss 

P.  &  Wagreich  M.  2016:  Palaeoenvironmental  changes  in  the 

northwestern Tethys during the Late Campanian Radotruncana 

calcarata Zone: Implications from stable isotopes and geoche-

mistry. Chem. Geol. 420, 280–296. 

Nothdurft L.D., Webb G.E. & Kamber B.S. 2004: Rare earth element 

geochemistry of Late Devonian reefal carbonates, Canning 

 Basin, Western Australia: confirmation of seawater REE proxy 

in ancient limestones. Geochim. Cosmochim. Acta 68, 263–283. 

Nozaki Y. 2001: Rare earth elements and their isotopes. Encycl. 

Ocean Sci. 4, 2354–2366. 

Nozaki Y., Zhang J. & Amakawa H 1997: The fractionation between 

Y and Ho in the marine environment. Earth Planet. Sci. Lett. 

148, 329–340.   

Olivarez A.M. & Owen R.M. 1991: The europium anomaly of sweater: 

implications for fluvial versus hydrothermal REE inputs to the 

oceans. Chem. Geol. 92, 317–328.

Papp  A.,  Cicha  I.,  Senes  J.  &  Steininger  F.  1978:  M4-Badenien 

(Moravien, Wielicien, Kosovien). Chronostratigraphie und 

Neostratotypen. Miozän der Zentralen Paratethys. Slowakische 

Akademie der Wissenschaften, Bratislava, 1–594.

background image

268

ALI and WAGREICH

GEOLOGICA CARPATHICA

, 2017, 68, 3, 248 – 268

Piepgras  D.J.  &  Jacobsen  S.B.  1992:  The  behavior  of  rare  earth 

 elements in seawater: Precise determination of variations in the 

North Pacific water column. Geochim. Cosmochim. Acta  56, 

1851–1862. 

Piller W.E. & Kleemann K. 1991: Middle Miocene Reefs and related 

facies in eastern Austria. 1) Vienna Basin: VI International 

 Symposium on Fossil Cnidaria including Archaeocyatha and 

 Porifera.  Excursion Guidebook, Excursion B4, 1–28.

Piller W.E., Decker K. & Haas M 1996: Sedimentologie und Becken-

dynamik des Wiener Beckens: Sediment 96, 11. Sedimentologen-

treffen, Exkursion guide, Geol. Bundesanst., Wien, 1–41.

Piller W.E., Summesberger H., Draxler I., Harzhauser M. & Mandic 

O. 1997: Meso- to Cenozoic tropical/subtropical climates — 

 Selected examples from the Northern Calcareous Alps and the 

Vienna Basin. In: Kollmann H.A. & Hubmann B. (Eds.) Excur-

sion Guides, Second European Paleontological Congress

 Climates: Past, Present and Future, Wien, 70–111.

Piller  W.E.,  Harzhauser  M.  &  Mandic  O.  2007:  Miocene  Central 

 Paratethys stratigraphy-current status and future directions. 

 Stratigraphy 4, 151–168. 

Piper D.Z. 1974: Rare earth elements in the sedimentary cycle: a sum-

mary. Chem. Geol. 14, 285–304. 

Reuter M., Piller W.E. & Erhart C. 2012: A Middle Miocene carbo-

nate platform under silici-volcaniclastic sedimentation stress 

(Leitha Limestone, Styrian Basin, Austria)- Depositional envi-

ronments, sedimentary evolution and palaeoecology: Paleo-

geogr. Paleoclimatol. Paleoecol. 350–352, 198–211. 

Riegl  B.  &  Piller W.E.  2000:  Biostromal  coral  facies  -  a  Miocene 

 example from the Leitha Limestone (Austria) and its actualistic 

interpretation. Palaios 15, 399–413.  

Rohatsch A. 2005: Neogene Bau- und Dekorgesteine Niederöster-

reichs und des Burgenlandes. Mitteilungen IAG BOKU,  9–56.

Rostá E. 1993: Gilbert-type delta in the Sarmatian-Pannonian sedi-

ments, Sopron, NW Hungary. Földtani Közlöny 123, 2, 167–193.

Royden L.H. 1985: The Vienna Basin: a thin-skinned pull-apart basin. 

In: Biddle K.T. & Christie-Blick N. (Eds) Strike-Slip deforma-

tion, basin formation, and sedimentation. SEPM Spec. Publ. 37, 

319–338 

Salvador A. 1994: International stratigraphic guide. International 

Union of Geological Sciences and The Geological Society of 

America, 1–214.

Schieber J. 2011: Marcasite in black shales- a mineral proxy for oxy-

genated bottom waters and intermittent oxidation of carbona-

ceous muds. J. Sediment. Res. 81, 447–458.

Schmid  H.P.,  Harzhauser  M.  &  Kroh A.  2001:  Hypoxic  events  on  

a Middle Miocene carbonate platform of the Central Paratethys 

(Austria, Badenian, 14 Ma). Ann. Naturhist. Mus. Wien 102, 1–50.

Schmidt K., Garbe-Schönberg D., Bau M. & Koschinsky A. 2010: 

Rare earth element distribution in >400 ºC hot hydrothermal 

 fluids from 5º S, MAR: the role of anhydrite in controlling 

 highly variable distribution patterns. Geochim. Cosmochim. 

Acta 74, 4058–4077. 

Scholle  P.A.  &  Ulmer-Scholle  D.S.  2003:  A  Color  Guide  to  the 

 Petrography of Carbonate Rocks. AAPG Memoir 77, 1–474. 

Schreilechner  M.G.  &  Sachsenhofer  R.F.  2007:  High  resolution 

 sequence stratigraphy in the eastern Styrian Basin (Miocene, 

Austria). Austrian J. Earth Sci. 100, 164–184.

Schuster  R.  &  Nowotny  A.  2015:  Die  Einheiten  des  Ostalpinen 

Kristallins auf den Kartenblättern GK50 Blatt 103 Kindberg und 

135 Birkfeld. Arbeitstagung der Geologischen Bundesanstalt 

Mitterdorf im Mürztal, 10–37.

Seghedi I. & Downes H. 2011: Geochemistry and tectonic develop-

ment of Cenozoic magmatism in the Carpathian-Pannonian 

 region.  Gondwana Res. 20, 655–672  

Slapansky  P.,  Belocky  R.,  Fröschl  H.,  Hradecký  P.  &  Spindler  P. 

1999: Petrography, Geochemie und geotektonische Einstufung 

des miozänen Vulkanismus im Steirischen Becken (Österreich). 

Abhandlungen der Geologischen Bundesanstalt 56, 419–434.

Steininger F.F. & Piller W.E. 1999: Empfehlungen (Richtlinien) zur 

Handhabung der stratigraphischen Nomenklatur. Courier For-

schungsinstitut Senckenberg 209, 1–19.

Strauss P., Harzhauser M., Hinsch R. & Wagreich M. 2006: Sequence 

stratigraphy in a classic pull-apart basin (Neogene, Vienna 

 Basin). A 3D seismic based integrated approach. Geol. Carpath. 

57, 185–197.

Suess E. 1860: Erhaltung von Fossilresten im Leithakalk. Verhand-

lungen der Geologischen Reichsanstalt, 1860, 1–9.

Taylor S.R. & McLennan S.M. 1985: The continental crust, its com-

position and evolution. Blackwell, 1–328.

Tollmann A. 1985: Geologie von Österreich. Band II. Außerzentral-

alpiner Anteil: Deuticke, Wien, 1–710.

Ullmann C.V., Campbell H.J., Frei R. & Korte C. 2016: Oxygen and 

carbon isotope and Sr/Ca signatures of high-latitude Permian to 

Jurassic calcite fossils from New Zealand and New Caledonia. 

Gondwana Res. 38, 60–73.

Veizer J. 1983: Trace elements and isotopes in sedimentary carbo-

nates. In: Reeder RJ (ed.): Carbonates: Mineralogy and 

 Chemistry.  Mineral. Soc. Am. 11, 265–299.

Wagreich M. & Schmid H.P. 2002: Backstripping dip-slip fault histo-

ries: apparent slip rates for the Miocene of the Vienna Basin. 

Terra Nova 14, 163–168.

Webb G.E., Nothdurft L.D., Kamber B.S., Kloprogge J.T. & Zhao J. 

2009: Rare earth element geochemistry of scleractinian coral 

skeleton during meteoric diagenesis: a sequence through neo-

morphism of aragonite to calcite. Sedimentology 56, 1433–1463.

 

Wedepohl K.H. 1978: Manganese: abundance in common sediments 

and sedimentary rocks. Handbook of Geochemistry. Springer

Berlin, 1-17.

Wessely  G.  1983:  Zur  Geologie  und  Hydrodynamik  im  südlichen-

Wiener Becken und seiner Randzone. Mitt. Geol. Ges. Wien 76, 

27–68.

Wessely  G.  2006:  Geologie  der  Österreichischen  Bundesländer  — 

Niederösterreich. Geologische Bundesanstalt, Wien, 1–416. 

Wiedl  T.,  Harzhauser  M.  &  Piller  W.E.  2012:  Facies  and  synsedi-

mentary tectonics on a Badenian carbonate platform in the 

southern Vienna Basin (Austria, Central Paratethys). Facies 58, 

523–548

Wiedl  T.,  Harzhauser  M.,  Kroh  A.,  Ćorić  S.  &  Piller  W.E.  2013: 

 Ecospace variability along a carbonate platform at the northern 

boundary of the Miocene reef belt (Upper Langhian, Austria). 

Palaeogeogr. Palaeoclimatol. Palaeoecol. 370, 232–246. 

Wiedl T., Harzhauser M., Kroh A., Ćorić S. & Piller W.E. 2014: From 

biologically to hydrodynamically controlled carbonate produc-

tion by tectonically induced paleogeographic rearrangement 

(Middle Miocene, Pannonian Basin). Facies 60, 865–881. 

Zhang J., Amakawa H. & Nozaki Y. 1994: The comparative behaviors 

of yttrium and lanthanides in the seawater of the North Pacific. 

Geophys. Res. Lett. 21, 2677–2680. 

Zhao  M.Y.  &  Zheng Y.F.  2014:  Marine  carbonate  records  of  terri-

genous input into Paleotethyan seawater: Geochemical con-

straints from Carboniferous limestones. Geochim. Cosmochim. 

Acta 141, 508–531.