background image

GEOLOGICA CARPATHICA

, FEBRUARY 2019, 70, 1, 62–74

doi: 10.2478/geoca-2019-0004

www.geologicacarpathica.com

Organic petrological and geochemical properties of jet 

from the middle Triassic Mogila Formation, West Bulgaria

ALEXANDER ZDRAVKOV

1, 

, GEORGE AJDANLIJSKY

2

, DORIS GROSS

3

 and ACHIM BECHTEL

3

1

Department of Geology and Exploration of Mineral Resources, University of Mining and Geology “St. Ivan Rilski”, 1700 Sofia, Bulgaria; 

 

alex_zdravkov@mgu.bg 

2

Department of Geology and Geoinformatics, University of Mining and Geology “St. Ivan Rilski”, 1700 Sofia, Bulgaria; g.ajdanlijsky@mgu.bg 

3

Department Angewandte Geowissenschaften und Geophysik, Montanuniversität Leoben, Peter-Tunner-Str. 5, A-8700 Leoben, Austria;  

doris.gross@unileoben.ac.at; achim.bechtel@unileoben.ac.at 

(Manuscript received November 27, 2018; accepted in revised form January 23, 2019)

Abstract: The paper presents the results of the petrographic and organic geochemical studies of a jet sample recovered 

from a Mid-Triassic carbonate succession from the West Balkan tectonic zone in Bulgaria. Total organic carbon contents 

(TOC = 92 % daf) and high vitrinite reflectance (Ro = 1.9 %) indicate semi-anthracite coalification rank. Very high T

max

 

(577 ºC) and low HI (~10 mg HC/g TOC) further support the overmature organic matter. Extractable organic matter is 

characterized by high portions of NSO compounds and asphaltenes (> 75 %). Hydrocarbons constitute about 20 % and 

are characterized by the predominance of the saturated hydrocarbons over the aromatics. n-Alkanes distribution, dominated 

by short-chain compounds (n-C

17–18

),  is  consistent  with  the  woody  origin  of  the  jet  and  the  thermal  maturity  of   

the organic matter. The predominance of PAHs with condensed structure over their alkylated isomers is considered to be 

a result of the complex reaction occurring within the organic matrix during the catagenesis, rather than to the presence of 

combustion-derived organic matter. Based on the distribution of the diterpenoids, a tentative identification of a possible 

Voltziales conifer family source is identified. Low Pr / Ph ratio (0.88) and aryl isoprenoids outline anoxic conditions of jet 

formation, whereas the presence of organic sulfur compounds and tri-MTTchroman suggest marine depositional 

 environment with normal salinity. 

Keywords: Bulgaria, Anisian jet, depositional environment, organic geochemistry, biomarkers.

Introduction

Recently, the occurrence of jet within the early Middle  Triassic 

(Anisian)  carbonate  succession  of  Mogila  Fm.  from  

the western part of the Balkan tectonic zone in Bulgaria, was 

noted (Ajdanlijsky et al. 2018). Jet occurs rarely in the sedi-

mentary record and is entirely constrained to Jurassic and 

 Cretaceous rocks (Howarth 1962; Minčev 1978, 1980, 1982; 

Minčev  &  Nikolov  1979;  Minčev  &  Šiškov  1986;  Suárez-

Ruiz et al. 1994 a, b; Bechtel et al. 2001b; Helfik et al. 2001; 

Marynowski et al. 2011b; Markova et al. 2017). Its formation 

proceeds through the deposition of drift wood fragments under 

anoxic conditions. Subsequent impregnation with bituminous 

substances, generated either within the resin-impregnated 

woods during maturation (Suárez-Ruiz et al. 1994b; Bechtel et 

al. 2001b), or derived externally from the surrounding envi-

ronment  (Suárez-Ruiz  et  al.  1994a)  aids  organic  matter 

 preservation. The molecular composition of the jet could be 

further influenced by bituminous matter, originating from  

the bacterial degradation of the organic matter (Bechtel et al. 

2001b). 

In this paper, the petrographic, bulk geochemical data and 

molecular composition of the non-polar extract fractions of 

the jet within the Anisian carbonate succession of Mogila Fm., 

is reported and used in order to infer its origin and depositional 

settings.

Geological settings

The Late Alpine Balkan Tectonic zone (Fig. 1a) is characte-

rized by widely distributed Triassic successions, which are 

part of the main regional tectonic structures and cover lower 

Palaeozoic high-grade metamorphosed and upper Palaeozoic 

sedimentary, igneous and volcanic rocks. Stratigraphically,  

the Triassic rocks are subdivided into three units: i) the Petrohan 

Terrigenous  Group  (Tronkov  1981)  consisting  mainly  of 

 fluvial  and  rare  alluvial  siliciclastic  deposits;  ii)  the  Iskar 

Carbonate  Group  (Tronkov  1981)  composed  of  shallow- 

marine carbonates and mixed siliciclastic-carbonate rocks; and 

iii) the Moesian Group (Chemberski et al. 1974) represented 

by terrigenous-carbonate and carbonate rocks. Tronkov (1983) 

defined four regional stratigraphic levels in the lower part of 

the Iskar Carbonate Group, i.e. the Tenuis, the Zhitolub,  

the Sfrazhen and the Sedmochislenitzi Beds (Fig. 1b). The log-

ged interval belongs to the Tenuis Bed from the base of  

the Iskar Carbonate Group and is situated about 10 m above 

the base of the Mogila Formation (Opletnya Member; Assereto 

et al. 1983). Recently, the rocks were chronostratigraphically 

constrained to the earliest Middle Triassic (Anisian; 

Ajdanlijsky  et  al.  2018).  Cyclic  sedimentary  succession  of 

mainly allochemic and mictritic limestone, alternating with 

dolomitized limestone and dolomite, is the most prominent 

feature of the Opletnya Member. 

background image

63

ORGANIC PETROLOGICAL AND GEOCHEMICAL PROPERTIES OF JET FROM THE MOGILA FORMATION

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

High-resolution lithostratigraphic description of the studied 

section is presented in Figure 1c. The sequence- and cyclo-

stratigraphic interpretation followed the concepts proposed by 

Strasser  et  al.  (1999).  The  sequence  boundaries  (SB)  were 

positioned into dolomitic beds with massive to poorly lamina-

ted structure, which are considered to represent the shallowest 

facies deposits. The transgressive surfaces (TS) are erosive, with 

amplitude in the range 8–10 cm, and mark the beginning of  

a deepening-up trend. The transgressive deposits (TSd) are com-

posed of massive to trough/planar cross-bedded allochemic to 

bioclastic limestones and dolo-limestones. Angular to round 

peb bly to cobble dolo- and limestone intraclasts, forming small 

lags or occurring along the foresets, are also observed. Well 

preserved bivalves, gastro- and cephalopod shales can also be 

recognized. Upwards, nodular to laminated and massive wacke- 

and mudstones are interbedded by centimeter-thick mainly 

fine grained beds of mixed carbonate– terrigenous deposits. 

The maximum-flooding surfaces (mfs) represent the deepest 

facies and are manifested by increasing bioturbation. 

The  shallowing-up  highstand  deposits  (HSd)  are  repre-

sented by massive and horizontally laminated, rarely nodular, 

mud- and wackestones with increasing upwards dolomite con-

tent. Up to 5 centimeter-thick (< 12 cm) beds with sigmoidal 

structure and concave up top surface, having lateral extension 

of < 30–40 m, are also common in this part of the elementary 

sequences. 

The thickness of the individual sequences from the Opletnya 

Member vary from 1.6 to over 4.3 m. Deposition within  

an unrimmed carbonate platform under peritidal settings is 

considered by Chatalov (2000).

Tectonically, the area belongs to the eastern part of the West 

Balkan tectonic zone (Fig. 1a). As part of the Alpine thrust belt 

the latter suffered multiple collisional and compressional 

events (in Late Triassic, Mid-Late Jurassic, Mid-Cretaceous, 

Late  Cretaceous  and  Mid-Eocene)  due  to  the  accretion  of 

proximal and exotic continental fragments to Eurasia during 

the closure of the Tethys ocean (Dabovski et al. 2002). 

Material and methods

The studied sedimentary succession is located about 35 km 

north of the capital city of Sofia, in the vicinity of the Lakatnik 

Railway Station within the Iskar gorge (43°05’18.2”N, 

23°23’01.6”E).  Two  small  jet  fragments  were  noted  within  

a relatively thin (~ 1.6 m) elementary sequence from the base 

of the Tenuis Bed (Fig. 1c). The first one represents ~ 4 cm 

long and ~ 0.7 cm thick extraclast within the lower transgres-

sive part of the cycle (Fig. 1c). Because of the small size, how-

ever, this jet fragment was not sampled as it would not yield 

enough material for the analytical procedu res. The second jet 

fragment is consistently larger (15×3 cm) and was recovered 

from a massive micritic limestone bed just above the maxi-

mum flooding surface (Fig. 1c). 

For microscopic investigation, three individual fragments 

(0.5 – 0.8 mm) of this jet were mounted in epoxy resin, ground 

and polished. Semi-quantitative maceral analysis was per-

formed on Leica DM 2500P microscope using reflected white 

and blue excitation light under oil immersion following stan-

dard  procedures  (Taylor  et  al.  1998).  Maceral  identification 

was done after ICCP (1998, 2001). Vitrinite reflectance was 

measured on 60 points using MIDAS MSP 200 spectrometer, 

attached to the same microscope. The calibration was done 

using  Spinel  (R = 0.421 %),  YAG  (R = 0.902 %)  and  GGG  

(R = 1.716 %) reflectance standards. EDS analyses were perfor-

med on JEOL JSM-6010 PLUS/LA scanning electron micro-

scope in order to study the composition of the established ore 

minerals. SEM was operated at reduced vacuum, back- 

scattered electron detector and 15 kV accelerating voltage.  

The total organic carbon (TOC) and sulphur (S) contents were 

determined with an Eltra Helios C/S analyzer. Moisture and 

ash yield were measured following standard procedures (ISO 

17246:2010). Rock-Eval pyrolysis was performed using a Rock- 

Eval 6 instrument. The value

 

of S

2

 (mg HC/g rock) was used to 

calculate  the  hydrogen  index  (HI = 100 × S

2

 / TOC  [mg  HC/g 

TOC];  Espitalié  et  al.  1977). The  temperature  of  maximum 

hydrocarbon generation (T

max

)  was  recorded  as  a  maturity 

parameter. 

For GC-MS analysis about 5 g of the jet sample were mixed 

with inert diatomaceous earth and homogenized. Extraction 

was performed by a Dionex ASE 200 equipment using 

dichloromethane for 1 hour at 75 ºC and 75 bar. The extract 

was concentrated using a Zymark Turbo Vap 500 device. 

Extract was dissolved in a solvent mixture of hexane:dichloro-

methane (80:1) and asphaltenes were subsequently separated 

by centrifugation. The hexane-soluble organic compounds 

(maltenes) were subdivided into saturated and aromatic hydro-

carbons and NSO components using a Köhnen-Willsch MPLC 

(medium pressure liquid chromatography) instrument (Radke 

et al. 1980).

The fractions of saturated and aromatic hydrocarbons were 

analyzed by Thermo-Fisher Trace GC Ultra analyzer, equipped 

with a 60 m silica capillary column (DB-5MS). Oven tempera-

ture was programmed from 40 –310 ºC with steps of 4 ºC/min, 

followed by isothermal period of 30 min. Helium was used as 

carrier gas. The device was set in electron impact mode with  

a scan rate of 50 – 600 Daltons (0.5 sec/scan). The results were 

processed with the software Thermo-Fisher Xcalibur v.1.4. 

Identification of biomarkers is based on retention time and 

comparison of mass spectra with published data. The determi-

nation of absolute concentrations of biomarkers was done using 

internal standards (deuterated n-tetracosane for the aliphatic 

fraction and 1,1’-binaphthyl for the aromatic fraction) and values 

were normalized against the total organic carbon contents.

Results and discussion

Petrography and ash yield

The studied jet represent single highly compressed fossil 

wood fragment, about 15 cm long and about 3 cm thick  

background image

64

ZDRAVKOV, AJDANLIJSKY, GROΒ and BECHTEL

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

Fig. 1. A — Schematic diagram of the main tectonic units in Bulgaria (simplified after Dabovski et al. 2002) with location of the studied profile. 

B — Lithostratigraphic column of early Mid-Triassic sediments: Te  — Tenuis Bed; Zt — Zhitolub Bed; Sf — Sfrazen Bed;  

Se — Sedmochislenitzi Bed; C — high-resolution litho- and cyclostratigraphic log of the studied carbonate succession, showing the position 

of the jet fragments.

background image

65

ORGANIC PETROLOGICAL AND GEOCHEMICAL PROPERTIES OF JET FROM THE MOGILA FORMATION

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

(Fig. 2a). It lies parallel to the bedding, within a thin (20 – 25 cm) 

laminated micritic limestone bed, constrained between allo-

chemic limestone at the base and allochemic dolomitic lime-

stone at the top (Fig. 1). Multiple millimeter sized (< 3 mm) 

carbonate-filled fractures indicate postdepositional deforma-

tions of the wood. Micropetrographic investigations revealed 

strongly gelified organic tissues (V; Fig. 2 b–e) significantly 

fragmented by at least two fracture systems. Neither liptinite 

nor inertinite macerals were detected. The fractures are mostly 

filled by carbonate minerals, but some host also pyrite (± pyr-

rhotite) – sphalerite  hydrothermal  mineralization  (Fig.  2 b, c). 

The latter is typically constrained to the fractures and only 

rarely  is  found  within  the  organic  matter  (Fig.  2e).  Hydro-

thermal pyrite always contain certain amount of zinc (pro-

bably in the form of micrometer-sized sphalerite inclusions; 

Fig. 2h), whereas sedimentary pyrite was found to be free of 

admixtures (Fig. 2i). Based on this observation sedimentary 

pyrite was tentatively differentiated from the hydrothermal 

one. Sedimentary pyrite is mostly represented by scarce 

micrometer-sized euhedral crystals with octahedral and penta-

gondodecahedral habit (Fig. 2f, i), scattered within the organic 

matter  (Fig.  2d).  Only  in  one  of  the  studied  jet  fragments,  

the presence of several clusters of framboidal and euhedral 

pyrite was established close to one of the sides of the fragment 

(Fig. 2f, g), which is assumed to be the outer side of the jet. 

Following the criteria of Wiese & Fyfe (1986) and Kortenski 

& Kostova (1996), and considering the co-occurrence of both 

framboidal and euhedral aggregates, mostly inorganic forma-

tion of pyrite can be assumed, although some of the framboids 

can be also considered of bacterial origin (e.g., Fig. 2g). Fram-

boids are typically poorly mineralized, which together with 

the very low amount of sedimentary pyrite (~ 0.1 %; Table 1) 

argue for a limited pyrite formation. Considering the abun-

dance of sulfur-containing organic compounds, the negligible 

pyrite formation indicates iron-deficient marine sedimentary 

environment. 

Vitrinite reflectance measurements (avg. Ro = 1.9 %; Table 1) 

indicate low volatile bituminous to semi-anthracite coalifi-

cation stage. However, because of the bituminization, jet is 

typically characterized by reduced vitrinite reflectance in 

com parison to coals of same maturity degree (Minčev 1978, 

1980; Minčev & Nikolov 1979; Petrova et al. 1985; Minčev & 

Šiškov 1986; Suárez-Ruiz et al. 1994a, b; Bechtel et al. 2001b). 

Therefore, in the present case vitrinite reflectance is presumed 

to be suppressed by the bituminization of the drift wood,   

and therefore, the actual coalification degree is expected to  

be higher. 

Because of the enhanced postdepositional mineralization, 

the ash yield of the studied jet is quite high (19 wt. %; Table 1). 

However, only a small fraction of it is expected to be contri-

buted from the living wood. Major elements, like Ca, Mg, Si, 

Al etc., which play a crucial role in growth or form structural 

supporting elements in the living tissues, are typically estab-

lished in living woods, but generally the total amount of ash do 

not exceed 0.5 wt. % (Rowell et al. 2005). Although the amount 

of inorganic matter in the woods vary due to taxonomy and 

environmental conditions of growth, it is highly unlike that  

the amount of wood ash in the studied jet sample will be higher 

than about 2 wt. %. Therefore, it is here assumed that the post-

depositional contribution to the jet’s ash yield is about 17 %, 

and all the geochemical parameters were recalculated in order 

to exclude this contribution.  

Bulk geochemical parameters

Bulk organic geochemical parameters, together with the nor-

malized yield of extractable organic matter (EOM) and pro-

portions of saturated, aromatic and polar compounds and 

asphaltenes, are summarized in Table 1. 

The  amount  of  total  organic  carbon  (TOC)  is  quite  high 

(92.4  wt. %; Table  1),  arguing  for  a  bit  higher  coalification 

rank (anthracite) than evidenced from the vitrinite reflectance. 

Very high T

max

 value of 577 °C (Table 1) suggests overmature 

organic matter. However, since the amount of hydrocarbons 

generated during the pyrolysis is quite low (S

2

 = 9.6 mg HC/g 

rock; Table 1) the high T

max

 might be erroneous. Nevertheless, 

the  very  low  hydrogen  index  (HI = 10.4  mg  HC/g  TOC)  is 

 consistent with the presence of inert organic matter. Since  

no inertinite was detected, the low HI unequivocally points 

towards overmature organic matter outside of the main  

hydrocarbon generation window. It is therefore possible to 

assume that the jet was subjected to temperatures over 150 °C. 

The pre sence of hydrothermal ore mineralization within  

the fractured jet suggest that organic matter might have been 

thermally  influenced by the hydrothermal activity. Over 40 

small-scale copper and lead-zinc deposits (mostly strata- 

bound)  of  presumably  Upper  Cretaceous  age  (Campanian–

Maastrichtian),  exerting  strong  tectonic  control,  have  been 

established within the western Balkan orogen (Mincheva-

Stefanova  1988).  How ever,  the  temperatures  of  main  ore 

 

formation in one of the big 

gest Pb–Zn deposits, i.e. 

“Sedmochislenitsi” deposit (located ~10 km NE of the studied 

section), most probably did not exceeded 80 –100 °C. In addi-

tion, no thermal changes can be established within the studied 

sedimentary sequence, thus suggesting that the temperature  

of the hydrothermal fluid might have been even lower.  

On the other hand, Botoucharov (2014) provide data for over-

mature  (Ro > 2 %)  organic  matter  within  the  Late  Triassic 

rocks at the front of the Balkan thrust zone (southern Moessian 

Platform  margin)  in  central  north  Bulgaria.  Furthermore, 

based on subsidence and thermal modelling, the same author 

indicates that the maximum degree of maturity of the organic 

matter was achieved by the end of late Aptian, i.e. significantly 

earlier than the presumed hydrothermal activity. Therefore, 

considering the complex geodynamic evolution of the Balkan 

orogen  (Dabovski  et  al.  2002),  burial  of  the  sedimentary 

sequence is here presumed to have exerted the main control on 

the organic matter maturity.

Total  sulfur  content  was  recorded  (Table  1),  but  due  to  

the epi 

genetic sulfide hydrothermal mineralization and 

 

the impos sibility to assess its share, the data cannot be inter-

preted in terms of environmental settings.

background image

66

ZDRAVKOV, AJDANLIJSKY, GROΒ and BECHTEL

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

Fig. 2. Petrography of the jet: a — macrophotograph showing the position of the jet within the sedimentary sequence; b, c — microphotographs 

showing the complex fragmentation patterns of the jet and the associated hydrothermal ore vein, V = vitrinite, Sph = sphalerite, C = carbonate 

minerals, Q = quartz, Py

m

 = massive hydrothermal pyrite;  d — a close-up microphotograph showing the typical highly gelified organic matter 

(V) with scarse micrometer sized euhedral pyrite crystals (Py

e

), oil immersion; e — an example of rare hydrothermal mineralization within  

the organic matter,  Py

m

 = massive pyrite, Pyr = pyrrhotite, FeO/OH = iron oxides/hydrooxides formed as weathering products of the sulfide 

minerals; f, g — framboidal (Py

f

) and euhedral (Py

e

) pyrite aggregates, oil immersion; h — SEM microphotograph of the hydrothermal vein 

mineralization with the typical pyrite and sphalerite composition; i — SEM microphotograph with composition of euhedral pyrite within  

the vitrinite. All microphotographs are taken under polarized white light. 

background image

67

ORGANIC PETROLOGICAL AND GEOCHEMICAL PROPERTIES OF JET FROM THE MOGILA FORMATION

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

Molecular composition of organic matter

The contents of extractable organic matter (EOM) in the stu-

died sample is low (0.4 mg/g TOC; Table 1) and is consistent 

with  the  (semi-)anthracite  rank  of  the  jet.  The  EOM  is 

 dominated  by  polar  compounds  and  asphaltenes  (Table  1). 

Predominance  of  the  saturated  (15.1 %)  over  the  aromatic 

hydrocarbons (7.3 %; Table 1) is established. 

n-Alkanes and isoprenoids

The total ion current (TIC) chromatogram of the saturated 

hydrocarbon fraction is shown in Figure 3. The concentrations 

and proportions of short-, mid- and long-chain n-alkanes are 

presented in Table 2.

The sample is characterized by the presence of n-alkanes in 

the range n-C

15

 to n-C

33

 (Figs. 3, 4). The distribution is uni-

modal and is characterized by the predominance of short-chain 

hydrocarbons with maximum concentrations at n-C

17-18

  

(Table 2, Fig. 4a). Prominent peaks at n-C

29

 and n-C

31

 (Fig. 4a) 

point toward the terrestrial origin of the organic matter. 

Long-chain n-alkanes (> n-C

25

) are common components of 

the  epicuticular  waxes  of  the  vascular  plants  (Eglinton  & 

Hamilton  1967)  and  are  therefore  widely  used  to  infer 

ter restrial origin of organic matter. On contrary, short-chain 

homologs  (< n-C

20

)  mostly  originate  from  algae  and  micro-

organisms (Blumer et al. 1971; Cranwell 1977), whereas mid-

chain  n-alkanes (n-C

21-25

)  are  related  to  submerged  aquatic 

plants  and  mosses  (Cranwell  1977;  Ficken  et  al.  2000), 

although bacterial lipids from sulfate-reducing bacteria also 

contain mid-chain homologs (Han & Calvin 1970). However, 

n-alkane distributions are known to be influenced by bacterial 

degradation (Allen et al. 1971; Johnson & Calder 1973) and 

thermal maturation (Tissot & Welte 1984; Peters et al. 2005) 

of the organic matter. The former results in the progressive 

removal of the n-alkanes, starting with the lighter ones, 

whereas the latter results in progressive thermal cracking of 

the long-chain homologs and the formation of shorter-chain 

n-alkanes during maturation. Considering the abundance of 

short-chain homologs in the studied sample, bacterial degra-

dation of the organic matter is unlike. Furthermore, the trace 

amounts of hopanes in the studied jet sample do not support 

the hypothesis of extensive biodegradation. Considering 

 

the over mature organic matter, however, thermal cracking of 

the long-chain hydrocarbons is expected. On the other hand, it 

has been shown that jet, similarly to other fossil palaeowoods 

(Fabiańska  &  Kurkiewicz  2013),  is  in  most  cases  characte-

rized by short-chain n-alkane distribution patterns (Bechtel et 

Parameter

Value

Units

Ash

19.0

wt. %

Total organic carbon (TOC)

92.4

wt. %

Sulfur

  2.7

wt. %

S

2

  9.6

mg HC/g rock

Hydrogen Index (HI)

10.4

mg HC/g TOC

T

max

 

  577.0

°C

Vitrinite

 100.0

vol. %, mmf

Pyrite (sedimentary)

  < 0.1

vol. %

Ro

 1.9

%

Extractable organic matter

 0.4

mg/ g TOC

Ʃ Saturated HC

15.1

%

Ʃ Aromatic HC

 7.3

%

Ʃ Polar

40.6

%

Ʃ Asph.

37.0

%

Table 1: Petrographic and organic geochemical characteristics of  

the studied jet.

Saturated hydr

ocarbons

n-Alkanes

Sum

  2.21

μg/g TOC

n-C

15-19

63.72

 %

n-C

21-25

15.55

 %

n-C

27-31

  8.85

 %

Isoprenoids

Sum

  0.42

μg/g TOC

Pr/Ph

  0.88

Pr/n-C

17

  0.41

Ph/n-C

18

  0.38

Diterpenoids

Sum

  0.50

μg/g TOC

Abietane-type

  9.54

%

Pimarane-type

36.85

%

Phyllocladane-type

53.61

%

Regular Steranes

Sum

  0.02

μg/g TOC

Hopanes

Sum

  0.02

μg/g TOC

Table 2: Molecular composition of the aliphatic fraction of the jet 

extract.

Fig. 3. Gas chromatogram of the saturated hydrocarbon fraction of the jet sample with partial chromatogram of diterpenoids distribution (a). 

Std. = standard.

background image

68

ZDRAVKOV, AJDANLIJSKY, GROΒ and BECHTEL

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

al.  2001b;  Marynowski  et  al.  2014;  Markova  et  al.  2017), 

because of the absence (or the minor amounts) of long-chain 

fatty acids within the woods, as these are present mainly in  

the leaf waxes. Considering these facts, the observed unusual 

n-alkane distribution pattern can be attributed mostly to  

the woody origin of the jet, overprinted by the effects of  

the thermal maturation of the organic matter. 

The studied jet sample is characterized by the presence of 

low amounts of isoprenoids in the range i-C

16–20

, among which 

pristane (Pr) and phytane (Ph) are the most abundant (Fig. 3; 

Table 2). Both commonly originate from the phytyl side chain 

of chlorophyll-α in photosynthesizing organisms (Peters et al. 

2005), whose degradation is strongly controlled by the Eh set-

tings of the depositional environment (Didyk et al. 1978; 

Volkman  &  Maxwell  1986).  Oxygenated  environments  are 

known to favor the formation of pristane, whereas anoxic set-

tings lead to preferential formation of phytane. Following this 

interpretation and considering the predominance of phytane 

over pristane and the low Pr/Ph ratio (< 1; Table 2) of the stu-

died jet sample, deposition of the drift wood fragments under 

reducing environment can be suggested. This hypothesis is 

also supported by the low Pr/n-C

17 

and Ph/n-C

18 

ratios (Fig. 4b; 

Table  2).  However,  the  amount  of  isoprenoids  might  be 

influen ced by the thermal maturity (Goossens et al. 1984; 

Volkman & Maxwell 1986; Koopmans et al. 1999; Peters et al. 

2005), as well as by the isoprenoid precursors (e.g., pristane 

formation from tocopherols (Goossens et al. 1984; ten Haven 

et  al.  1987),  and  phytane  formation  from  bacterial  lipids 

(Volkman & Maxwell 1986; Peters et al. 2005). Considering 

the high maturity degree of the organic matter, significant ren-

dering of the pristane concentrations due to the thermal evolu-

tion of the jet can be expected. In addition, isoprenoid 

contribution from additional sources cannot be excluded. 

Nevertheless, the presumed organic matter deposition under 

anoxic conditions is in agreement with the generally accepted 

mode of jet formation (e.g., Suárez-Ruiz et al. 1994a; Bechtel 

et al. 2001b). 

Diterpenoids

The studied jet sample is characterized by the presence of 

low amounts of tri- and tetracyclic saturated diterpenoids  

(0.5 μg/g TOC; Table 2) with abietane (fichtelite, abietane), 

pimarane (norisopimarane, norpimarane, iso-pimarane),  and 

phyllocladane  (α-  and  β-phyllocladane)  skeleton,  among 

which α-phyllocladane and pimarane are dominant (Fig. 3a). 

Their  aromatic  counterparts  are  more  abundant  (1.20  μg/g 

TOC;  Table 3)  and  include  abietane  type  hydrocarbons 

(nor(19)abieta-3,8,11,13-tetraene, nor(19)abieta-8,11,13-triene, 

dehydroabietane, simonellite, tetrahydro(1,2,3,4)retene, diaro-

matic totarane, retene), among which simonellite and retene 

are predominant. Trace amounts of diaromatic totarane-type 

diterpenoid is also established (Fig. 5a). 

Diterpenoid biomarkers are widely distributed in various 

sedimentary rocks, coal, fossil resins and amber (Simoneit 

1977, 1986; Noble et al. 1985, 1986; Sukh Dev 1989; Otto & 

Fig. 4. Histogram, showing the distribution of the n-alkanes (a)  

and cross-plot of Phytane/n-C

18

 versus Pristane/n-C

17

 (b)  marking  

the disoxic/anoxic depositional environment.

Ar

omatic hydr

ocarbons

Unsubstituted 

PAHs

Sum

17.57

μg/g TOC

Fl/(Fl+Py)

  0.44

BaA/(BaA+Tri+Ch)

  0.15

IP/(IP+BgP)

  0.41

Alkyl-

Naphthalenes

MN

  0.00

μg/g TOC

DMN

  0.01

μg/g TOC

TMN

  0.09

μg/g TOC

TeMN

  0.44

μg/g TOC

Alkyl-

Phenanthrenes

MP

  1.41

μg/g TOC

DMP

  0.15

μg/g TOC

MPI-1

  0.19

Rc

  2.19

%

Organic S 

compounds

DBT

  0.85

μg/g TOC

MDBT

  0.66

μg/g TOC

DMDBT

  0.21

μg/g TOC

BNT

  2.28

μg/g TOC

MDR

12.47

Furans

  0.05

μg/g TOC

Diterpenoids

  1.20

μg/g TOC

Aryl isoprenoids

  0.46

μg/g TOC

Chromans

  0.01

μg/g TOC

Table 3: Molecular composition of the aromatic fraction of the jet 

extract.

background image

69

ORGANIC PETROLOGICAL AND GEOCHEMICAL PROPERTIES OF JET FROM THE MOGILA FORMATION

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

Wilde 2001), and are common constituents of jets (Bechtel et 

al.  2001b;  Marynowski  et  al.  2014;  Markova  et  al.  2017). 

Diterpenoids are considered indicators of gymnosperm plants, 

because of their presence in essential oils and resins. Based on 

the distribution of the diterpenoid compounds, a general taxo-

nomic differentiation of the different conifer species may be 

possible (e.g., Simoneit 1977; Wakeham et al. 1980; Noble et 

al.  1985;  Otto  &  Simoneit  2001;  Otto  &  Wilde  2001).  

The presence of aromatic abietane type diterpenoids (i.e. simo -

nelite and retene) in recent sediment, however, indicate that aro-

matization proceeds early in the diagenetic process (Simoneit 

1998). Cyclization and aromatization of diterpenoids is known 

to be mediated by clay mineral catalysts or microbial activity 

(Wakeham  et  al.  1980).  Furthermore,  diterpenoids  are  ther-

mally unstable and commonly aromatize to retene and even-

tually to alkyl-phenanthrenes and phenanthrene with increasing 

rank  (Hayatsu  et  al.  1978).  Therefore,  considering  the  high 

level of thermal maturity of the studied jet, the presence of 

diterpenoids is surprising. Although the reasons for the occur-

rence of diterpenoids in this study are currently not fully 

understood, a combination of a very limited bacterial activity, 

which correspond to the negligible hopane concentrations,  

as well as a limited availability of clay mineral catalysts in  

the carbonate-dominated host rocks, can be considered.

Based on the available geochemical data for the studied jet 

sample, hardly any specific chemotaxonomic assignment 

could be done, because of the presence of diterpenoid com-

pounds, characteristic for the whole gymnosperm group. 

However, it should be noted that the radiation of the modern 

gymnosperm families (except Podocarpaceae;  Cleal  1993) 

from their common ancestors (e.g., Voltziales conifers) did not 

started earlier than the Late Triassic (Cleal 1993; Taylor et al. 

2009). Therefore, considering the early Anisian age of the sedi-

mentary succession, it can be assumed that the drift wood 

most likely originated from the Voltziales conifer family. 

Indirectly, this hypotheses is supported by the establishment  

of dominant Voltziales (e.g., Voltzia,  Albertia,  Yuccites

Aethophyllum) species in the early Anisian red beds of the Grès 

à Voltzia Formation from the western margin of the German 

Basin. The latter is believed to represent one of the first loca-

lities, where the resurgence of the gymnosperms (and particu-

larly the conifers), which survived the harsh arid- to semi-arid 

conditions following the end-Permian life crisis, have occurred 

(Grauvogel-Stamm & Ash 2005). On the other hand, Taylor et 

al. (2009) noted that many of the early conifers share common 

botanical features with other families, and it can therefore be 

suggested that this poor differentiation is also expressed in 

their chemical characteristics.

Steranes and hopanes

Regular steranes and hopanes were detected in equal trace 

amounts (0.02 μg/g TOC; Table 2), but because of the difficult 

identification of the individual compounds due to the poor 

 signal-to-noise ratio, their distribution is not discussed here. 

The very low concentrations of hopanes might indicate a very 

limited bacterial activity. This is consistent with the presumed 

coniferous origin of the drift wood, in which resins typically 

act as anti-bacterial agent. However, both hopanes and stera-

nes are thermally unstable at very high maturity degrees (e.g., 

Stout  &  Emsbo-Mattingly  2008)  and  considering  the  semi- 

anthracite rank of the studied jet this might be the main reason 

for their general absence. 

No rearranged steranes were detected, which is consistent 

with the dominant carbonate sedimentation during the drift-

wood deposition. Surprisingly, however, the aromatic fraction 

does not provide evidence for the presence of neither aromatic 

steranes, nor hopanes, although their aromatization is known 

to occur quite early during the diagenesis (Hussler et al. 1984; 

Peters et al. 2005). However, since the cyclization and aroma-

tization reactions are favored by the catalytic activity of clay 

minerals, the carbonate-dominated depositional environment 

might have predetermined their absence.

 

Polyaromatic hydrocarbons hydrocarbons (PAHs)

The analysis of the aromatics revealed complex composition, 

characterized by the occurrence and dominance of the unsub-

stituted three to six rings PAHs, over their alkyl- and 

 phenyl- substituted  derivatives  (Fig.  5;  Table  3). Apart  from 

the abundant phenanthrene, the C

0

 PAHs are characterized by 

predominance of three and four-ring compounds and notable 

bell-shaped distribution, maximizing at chrysene (Fig. 6).

Predominance of condensed over alkyl-substituted PAHs is 

often considered to represent contribution from combustion- 

derived organic matter (e.g., Laflamme & Hites 1978; Wakeham 

et al. 1980; Yunker et al. 2002). However, such pyrogenic pat-

terns are not uncommon for high-rank coals (semi-anthracite 

and  anthracite)  for  which  rearrangement,  fragmentation  and 

condensation reactions during coalification result in progres-

sive loss of alkyl side chains and thus produce significant pre-

dominance of the unsubstituted PAHs (e.g., Radke et al. 1982; 

Wang  et  al.  2017).  The  latter  is  especially  pronounced  for  

the two- and three-ring compounds (Chen et al. 2004; Stout & 

Emsbo-Mattingly 2008). In addition, phenanthrene concentra-

tions in high rank coal can also be increased by demethylation 

reactions  of    aromatic  diterpenoids  (Hayatsu  et  al.  1987).  

On  the  other  hand,  Bechtel  et  al.  (2001a)  report  significant 

increase of phenanthrene concentrations relative to the sum of 

the methylphenanthrenes, together with abundant sulfur and 

oxygen heterocyclic hydrocarbons, as a result of demethyla-

tion processes related to oxidative hydrothermal alteration of 

Kupferschiefer Formation near the Rote Fäule zone. Although 

such explanation for the significant predominance of phenan-

threne in the present study seems unlikely given the detected 

very low concentration of furans (Table 3), the possibility of 

triphenylene,  benz[b]fluoranthene,  benz[e]pyrene formation 

from dehydrocyclization of alkyl- or phenyl-substituted PAHs 

(Grafka et al. 2015) cannot be completely excluded as a reason 

for the decreased methylphenanthrene concentrations. Never-

theless, considering the above discussion, as well as the (semi-) 

anthracite rank of the studied jet, the established PAHs 

background image

70

ZDRAVKOV, AJDANLIJSKY, GROΒ and BECHTEL

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

distribution can mostly be attributed to the thermal transfor-

mation reactions within the organic matrix. However, because 

of the presence of unsubstituted PAHs, which have no appa-

rent biological precursors and are mostly considered to form 

during burning of organic matter (e.g., pyrene, chrysene, 

benz[a]pyrene, etc.; Fig. 5; Chen et al. 2004; Keiluweit et al. 

2012), the possibility of contribution from combustion- derived 

hydrocarbons should not be completely excluded. In order to 

evaluate  such  contribution  the  ratios  Fl/(Fl+Py),  BaA/

(BaA+Tri+Ch) and In/(In+BgP) (Yunker et al. 2002) were cal-

culated (Table 3). Following the criteria of Yunker et al. (2002) 

for separation of diagenetically-derived from combustion- 

deriver PAHs, however, somewhat contradictory conclusions 

can be drawn. While the ratios Fl/(Fl+Py) and In/(In+BgP) are 

above the boundaries for diagenetically-derived PAHs (0.4 and 

0.2  respectively; Yunker  et  al.  2002)  and  thus  indicate  that 

minor contribution from pyrolytic organic matter is possible, 

the  low  value  of  the  ratio  of  BaA / (BaA + Tri + Ch)  (< 0.2;  

Table 3) is consistent with the absence of such contribution. 

The results suggest there might be separate sources of  

the individual PAHs. As discussed above, the established 

PAHs distribution is probably mostly a result of the diagenetic 

and catagenetic transformation of the organic matter. 

Nevertheless, minor anthropogenic contribution, especially to 

the five- and six-ring PAHs, should also be taken into account, 

considering the fact that the studied outcrop is adjacent to  

a major road and a roadside restaurant.

The alkyl substituted naphthalenes and phenanthrenes are 

widely used to reflect the thermal maturity of the organic 

 matter. Numerous alkyl-naphthalene and alkyl-phenanthrene 

ratios have been developed for that purpose (Radke & Welte 

1981; Radke et al. 1982, 1986; van Aarssen 1999; Stojanović 

et al. 2007). Because the observed distribution of the alkyl-

naphthalenes is clearly modified by secondary processes  

(i.e. water washing, oxidation or biodegradation; Volkman et 

al. 1984; Marynowski et al. 2011a), the methylphenanthrene 

index (MPI-1; Radke & Welte 1981), which is calibrated and 

widely used as maturity indicator of Type-III kerogen, was 

used to further support the maturity assessment of the jet. 

Following the established empirical relation between MPI-1 

and the vitrinite reflectance (Rc = −0.6*MPI-1 + 2.3; Radke et 

al.  1984),  the  calculated  methylphenanthrene  index  (0.19; 

Table 3) can be correlated to equivalent vitrinite reflectance 

(Rc) of 2.19 % (Table 3). The value is close to the measured 

vitrinite  reflectance  (Ro = 1.9;  Table  1)  and  suggest  that  

the sup pressing effect that the bituminization play on Ro  

(e.g.,  Suárez-Ruiz  et  al.  1994a, b)  might  be  significantly 

reduced during maturation, possibly as a result of the transfor-

mation of the impregnating resins.

The aromatic fraction is further characterized by the occur-

rence  of  sulfur  heterocyclic  compounds  (Fig.  5;  Table  3). 

These are represented by dibenzothiophene and its alkyl-sub-

stituted  isomers,  as  well  as  by  benzo[b]naphthothiophenes 

(Fig. 5; Table 3). These compounds have no obvious biolo-

gical precursors and are therefore considered to originate from 

Fig. 5. Gas chromatograms of the aromatic hydrocarbon fraction of the jet sample with partial chromatograms of aromatic diterpenoids (a), 

aryl isoprenoids (b)  and  methyl-  and  dimethyl-dibenzothiophenes  (c).  TMN = trimethylnaphthalenes;  TeMN = tetramethylnaphthalenes;  

P = phenanthrene; MP = methylphenanthrenes; DMP = dimethylphenanthrenes; DBT = dibenzothiophenes; MDBT = methyldibenzothiophenes; 

DMDBT = dimethyldibenzothiophenes;  DBF = dibenzofurans;  BNT = benzonaphthothiophenes;  MFl = methylfluoranthene;  PhN = phenyl-

naphthalene; PhP = phenylphenanthrenes; PhA = phenylanthracenes; Std. = standard.

Fig. 6. Distribution of the unsubstituted PAHs.

background image

71

ORGANIC PETROLOGICAL AND GEOCHEMICAL PROPERTIES OF JET FROM THE MOGILA FORMATION

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

the reaction of the organic matrix with sulfur during early 

diagenesis (White et al. 1988). Although the mechanisms of 

sulfur  incorporation  are  still  debatable  (Wang  et  al.  2017),  

it is generally accepted that the origin and distribution of  

the organic sulfur compounds are controlled by the deposi-

tional  conditions  (Hughes  1984;  Hughes  et  al.  1995)  and 

maturity (Radke et al. 1986). Sediments, deposited in marine 

environments typically contain increased concentrations of 

sulfur aromatics, although hypersaline settings in lacustrine 

basins could also produce such compounds (Hughes et al. 

1995;  Radke  et  al.  2000).  Owning  to  the  carbonate  deposi-

tional environment, the presence of sulfur aromatics in 

 

the studied jet is not surprising. On the other hand, Radke et al. 

(1986) pointed out that methyl shifts in the alkyl-dibenzothio-

phenes follow the same pathway as in the alkyl substituted 

naphthalenes and phenanthrenes, and based on this proposed 

the methyldibenzothiophene ratio (MDR) as additional matu-

rity  parameter  (MDR = 4-MDBT/1-MDBT). The  distribution 

of the methyldibenzothiophenes in the examined aromatic 

extract is clearly dominated by the thermodynamically more 

stable  4-MDBT  (MDR = 12.47),  which  is  another  proof  of  

the enhanced maturity of the jet’s organic matter. 

Aryl isoprenoids and chromans

Small  amounts  (0.46  μg/g TOC; Table  3)  of  C

17

– C

22

 aryl 

isoprenoids with 2,3,6-trimethyl substitution pattern for the aro-

matic ring and a tail-to-tail isoprenoid chain, were tentatively 

identified in the aromatic extract fraction (Fig. 5b). The aryl 

isoprenoids are mostly considered to derive from isorenie-

ratene, which is known to be synthesized by photosynthetic 

green sulfur bacteria. Since these organisms are phototrophic 

anaerobes and require both light and H

2

S for growth (Pfennig 

1977;  Summons  1993),  the  presence  of  aryl  isoprenoids  is 

often interpreted as indication for photic zone anoxia (e.g., 

Summons  &  Powell  1987;  Sinninghe  Damsté  et  al.  2001). 

However, Koopmans et al. (1996) report aryl isoprenoids with 

a 2,3,6-trimethyl substitution as diagenetic transformation 

products of β-carotene, thus indicating that isorenieratene is 

not their sole precursor. Considering the very advanced 

organic matter maturity of the studied jet, the failure to detect 

isorenieratene, which is typically reported from immature 

sedi ments (e.g., Grice et al. 1996; Sinninghe Damsté et al. 

2001), is not surprising. Because of this and the absence of 

carbon isotopic data, however, the detected aryl isoprenoids 

cannot be equivocally assigned to green sulfur bacteria. 

Nevertheless, considering the abundance of organic sulfur 

compounds, arguing for H

2

S-rich environment, as well as  

the anoxic settings, that are prerequisite for jet formation and 

further evidenced from the low Pr/Ph ratio, at least partial 

 origin of aryl isoprenoids from green sulfur bacteria can tenta-

tively be suggested.

In a series of mono-, di- and trimethylated 2-methyl- 2-

trimethyltridecylchromans (MTTC; Sinninghe Damsté et al. 

1987)  only  trace  amount  of  the  tri-MTTC  was  tentatively 

identified  in  the  present  study  (0.01  μg/gTOC;  Table  3). 

Despite of the fact that the origin of the methyl substituted 

MTTCs is still controversial (Sinninghe Damsté et al. 1987, 

1993) they are widely used as a palaeosalinity indicator (e.g., 

Schwark & Püttmann 1990; Grice et al. 1998; Schwark 1998; 

Bechtel  et  al.  2013).  Empirical  evidences  (e.g.,  Sinninghe 

Damsté et al. 1987, 1993) suggest that formation and preser-

vation of mono-, di- and trimethyl substituted chromans is 

strongly controlled by the salinity of the depositional environ-

ment. Thus non-hypersaline environment favor the synthesis 

of tri-MTTC, whereas mono-MTTCs are preferentially formed 

under hypersaline settings. Considering the presence of only 

tri-MTTC in the studied jet, hypersaline settings seem rather 

unlike. It is therefore possible to suggest that the deposition 

and further transformation of the drift wood occurred under 

normal salinity marine environment. The latter is also sup-

ported by the complete absence of gammacerane, which is 

also widely used to infer hypersalinity (e.g., Jiamo et al. 1986; 

Sinninghe Damsté et al. 1995; Peters et al. 2005), and is in 

agreement  with  the  conclusions  of  Chatalov  &  Stanimirova 

(2001) based on the conditions of early diagenetic dolomiti-

zation of carbonate mud from Mogila Fm.

Conclusions

Early Mid-Triassic jet, occurring within the cyclic carbonate 

succession from the lower part of Mogila Formation from  

the Western Balkan tectonic zone in Bulgaria, was characte-

rized by petrographic analysis, Rock Eval pyrolysis, and 

organic geochemistry proxies. The results indicate that the jet 

was formed from wood of vascular plant. Тhe Voltziales coni-

fer family seem the most probable source based on the estab-

lished poor differentiation of the diterpenoid hydrocarbons. 

High  total  organic  carbon  content  (TOC = 92 %),  vitrinite 

reflectance  value  (Ro = 1.9 %)  and  T

max

  (577  °C)  argue  for 

overmature  organic  matter  at  (semi-)anthracite  coalification 

rank. The latter is also well expressed in the molecular compo-

sition of the jet. The dominance of short-chain n-alkane homo-

logs (n-C

17–18

) and the low amounts of the long-chain n-alkanes 

is considered to reflect the woody origin of the jet, overprinted 

by the enhanced thermal maturity. Furthermore, the predomi-

nance of PAHs with condensed structure over their alkylated 

isomers most probably reflect the complex fragmentation and 

condensation reactions within the organic matrix during  

the coalification, rather than the input of combustion- 

derived organic matter. The calculated equivalent reflectance 

(Rc = 2.14 %), based on the distribution of phenanthrene and 

methylphenanthrenes, is close to the measured one and further 

proofs the (semi-)anthracite rank. Because of the timing and 

the presumed low temperature hydrothermal fluids, the estab-

lished epigenetic hydrothermal activity is assumed as of negli-

gible influence on the thermal maturity of the organic matter. 

Therefore, the enhanced maturity of the jet is considered to be 

a result of the complex geodynamic evolution of the West 

Balkan tectonic zone and the deep burial of the sediments.  

The  low  Pr/Ph  ratio  (< 1),  as  well  as  the  presence  of  aryl 

background image

72

ZDRAVKOV, AJDANLIJSKY, GROΒ and BECHTEL

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

isoprenoids, are consistent with drift wood deposition under 

anoxic environmental settings. Furthermore, the presence of 

tri-MTTC and the absence of gammacerane argue for normal 

salinity marine environment of deposition.

Aknowledgments:  We would like to express our sincere 

grati tude to Prof. Dr. I. Kostova and D. Apostolova (Sofia 

 University “St. Kl. Ohridski”) for their support with vitrinite 

reflectance measurements. Additional gratitude are expressed 

to Assist. Prof. S. Dobrev and Assist. Prof. G. Lyutov (Univer-

sity of Mining and Geology “St. Ivan Rilski”, Sofia) for their 

helpful comments regarding the hydrothermal mineralization, 

and to Prof. Dr. J. Kortenski for the valuable comments on  

the origin of sedimentary pyrite. The critical reviews from 

three anonymous reviewers are also greatly acknowledged.  

References

Ajdanlijsky G., Götz A.E. & Strasser A. 2018: The Early to Middle 

Triassic continental–marine transition of NW Bulgaria: sedi-

mentology, palynology and sequence stratigraphy. Geol. Carpath. 

69, 129–148.

Allen J.E., Forney F.W. & Markovetz A.J. 1971: Microbial subtermi-

nal oxidation of alkanes and alk-1-enes. Lipids 6, 448–452.

Assereto  R.,  Tronkov  D.  &  Ćatalov  G.  1983:  Mogila  Formation 

(Lower–Middle Triassic) — a new Formation in west Bulgaria. 

Geol. Balc. 13, 25–27 (in Russian).

Bechtel  A.,  Gratzer  R.,  Püttmann  W.  &  Oszczepalski  S.  2001a: 

 Variable alteration of organic matter in relation to metal zoning 

at the Rote Fäule front (Lubin–Sieroszowice mining district, SW 

Poland). Org. Geochem. 32, 377–395.

Bechtel A., Gratzer R. & Sachsenhofer R.F. 2001b: Chemical charac-

teristics of Upper Cretaceous (Turonian) jet of the Gosau Group 

of Gams/Hieflau (Styria, Austria). Int. J. Coal Geol. 46, 27–49.

Bechtel A., Movsumova U., Strobl S.A.I., Sachsenhofer R.F., 

 

 

Soliman A., Gratzer R. & Püttmann W. 2013: Organofacies and 

paleoenvironment of the Oligocene Maikop series of Angeharan 

(eastern Azerbaijan). Org. Geochem. 56, 51–67.

Blumer  M.,  Guillard  R.R.L.  &  Chase  T.  1971:  Hydrocarbons  of 

 marine  phytoplankton.  Mar. Biol. 8, 183–189.

Botoucharov N. 2014: Vitrinite reflectance, limitations and modelling 

of oil and gas windows in the central southern part of the Moe-

sian Platform, In: Proceedings of the IV International Scientific 

and Technical Conference “Geology and Hydrocarbon Potential 

of the Balkan-Black Sea Region” 11–15. 09. 2013, Varna, Bul-

garia. 214–221.

Chatalov A.G.  2000: The  Mogila  Formation  (Spathian–Anisian)  in 

the Western Balkanides of Bulgaria — ancient counterpart of an 

arid peritidal complex. Zbl. Geol. Paleont. Teil-I, 1123–1135.

Chatalov  A.G.  &  Stanimirova  T.  2001:  Diagenesis  of  Spathian– 

Anisian dolomites from the western Balkanides, Bulgaria. N. Jb. 

Geol. Palaont. Mh. 301–320.

Chemberski G., Vaptsarova A. & Monahov I. 1974: Lithostratigraphy 

of the Triassic variegated terrigenous-carbonate and carbonate 

sediments studied with deep drilling in Northwestern and  Central 

North Bulgaria. Ann. DSO Geol. Res 20, 327–341.

Chen Y., Bi X., Mai B., Sheng G., & Fu J. 2004: Emission characte-

rization of particulate/gaseous phases and size association for 

polycyclic aromatic hydrocarbons from residential coal combus-

tion. Fuel 83, 781–790.

Cleal C.J. 1993: Gymnospermophyta. In: Benton M.J. (Ed.): The Fos-

sil Record 2. Chapman and Hall, London, 795–808.

Cranwell P.A. 1977: Organic geochemistry of Cam Loch (Sutherland) 

sediments. Chem. Geol. 20, 205–221.

Dabovski C., Boyanov I., Khrischev K., Nikolov T., Sapunov I., 

 Yanev Y. & Zagorchev I. 2002: Structure and Alpine evolution of 

Bulgaria. Geol. Balc. 32, 9–15.

Didyk  B.M.,  Simoneit  B.R.T.,  Brassell  S.C.  &  Eglinton  G.  1978: 

 Organic geochemical indicators of palaeoenvironmental condi-

tions of sedimentation. Nature 272, 216–222.

Eglinton G. & Hamilton R.J. 1967: Leaf epicuticular waxes. Science  

156, 1322–1335.

Espitalié J., Laporte J.L., Madec M., Marquis F., Leplat P., Paulet J. & 

Boutefeu A. 1977: Méthode rapide de caractérisation des roches 

mères, de leur potentiel pétrolier et de leur degré d’évolution. 

Rev. l’Institut Français du Pétrole 32, 23–42.

Fabiańska M.J. & Kurkiewicz S. 2013: Biomarkers, aromatic hydro-

carbons and polar compounds in the Neogene lignites and 

gangue sediments of the Konin and Turoszow Brown Coal Ba-

sins (Poland). Int. J. Coal Geol. 107, 24–44.

Ficken  K.J.,  Li  B.,  Swain  D.L.  &  Eglinton  G.  2000: An  n-alkane 

proxy for the sedimentary input of submerged/floating freshwa-

ter aquatic macrophytes. Org. Geochem. 31, 745–749.

Goossens  H.,  de  Leeuw  J.W.,  Schenck  P.A.  &  Brassell  S.C.  1984: 

Tocopherols as likely precursors of pristane in ancient sediments 

and crude oils. Nature 312, 440–442.

Grafka  O.,  Marynowski  L.  &  Simoneit  B.R.T.  2015:  Phenyl 

 derivatives of polycyclic aromatic compounds as indicators  

of hydrothermal activity in the Silurian black siliceous shales  

of the Bardzkie Mountains, Poland. Int. J. Coal Geol. 139,  

142–151.

Grauvogel-Stamm L. & Ash S.R. 2005: Recovery of the Triassic land 

flora from the end-Permian life crisis. Comptes Rendus Palevol 

4, 593–608.

Grice K., Schaeffer P., Schwark L., & Maxwell J.R. 1996: Molecular 

indicators of palaeoenvironmental conditions in an immature 

Permian shale (Kupferschiefer, Lower Rhine Basin, north-west 

Germany)  from  free  and  S-bound  lipids.  Org. Geochem. 25, 

131–147.

Grice K., Schouten S., Peters K.E., & Sinninghe Damsté J.S. 1998: 

Molecular isotopic characterisation of hydrocarbon biomarkers 

in Palaeocene–Eocene evaporitic, lacustrine source rocks from 

the Jianghan Basin, China. Org. Geochem. 29, 1745–1764.

Han J. & Calvin M. 1970: Branched alkanes from blue–green algae. 

J. Chem. Soc. D. 22, 1490–1491.

Hayatsu R., Botto R., Scott R., Mcbeth R. & Winans R. 1987: Ther-

mal catalytic transformation of pentacyclic triterpenoids: Alte-

ration of geochemical fossils during coalification. Org. Geochem. 

11, 245–250.

Hayatsu  R.,  Winans  R.,  Scott  R.,  Moore  L.  &  Studier  M.  1978: 

Trapped organic compounds and aromatic units in coals. Fuel 

57, 541–548.

Helfik W., Kwieciǹska B., & Zmudzka A. 2001: The occurence of 

gagate  in  Soltikow  (The  Holy  Cross  Mts.).  Mineral. Pol. 32, 

47–54.

Howarth B.Y.M.K. 1962: The Jet Rock Series and the Alum Shale 

Series of the Yorkshire coast. Proc. Yorksh. Geol. Soc. 33, 

 381–422.

Hughes W.B. 1984: The use of thiophenic organosulfur compounds in 

characterizing crude oils derived from carbonate versus silici-

clastic  source  rocks,  In:  Palacas  J.G.  (Ed.):  Petroleum  Geo-

chemistry and Source Rock Potential of Carbonate Rocks.  

AAPG Studies in Geology 18, 181–196.

Hughes  W.B.,  Holba  A.G.  &  Dzou  L.I.P.  1995:  The  ratios  of 

 dibenzothiophene to phenanthrene and pristane to phytane  

as indicators of depositional environment and lithology of  

petroleum source rocks. Geochim. Cosmochim. Acta 59, 

 

3581–3598.

background image

73

ORGANIC PETROLOGICAL AND GEOCHEMICAL PROPERTIES OF JET FROM THE MOGILA FORMATION

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

Hussler  G.,  Albrecht  P.,  Ourisson  G.,  Cesario  M.,  Guilhem  J.  & 

 Pascard C. 1984: Benzohopanes, a novel family of hexacyclic 

geomarkers in sediments and petroleums. Tetrahedron Lett. 25, 

1179–1182.

ICCP  1998:  The  new  vitrinite  classification  (ICCP  System  1994). 

Fuel 77, 349–358.

ICCP 2001: The new inertinite classification (ICCP System 1994). 

Fuel 80, 459–470.

ISO 17246:2010: Coal — Proximate analysis.

Jiamo F., Guoying S., Pingan P., Brassell S.C., Eglinton G. & Jigang J. 

1986: Peculiarities of salt lake sediments as potential source 

rocks in China. Org. Geochem. 10, 119–126.

Johnson R.W. & Calder J.A. 1973: Early diagenesis of fatty acids and 

hydrocarbons in a salt marsh environment. Geochim. Cosmo-

chim. Acta 37, 1943–1955.

Keiluweit  M.,  Kleber  M.,  Sparrow  M.A.,  Simoneit  B.R.T.  &  

Prahl F.G. 2012: Solvent-Extractable Polycyclic Aromatic 

 Hydrocarbons in Biochar: Influence of Pyrolysis Temperature 

and Feedstock. Environ. Sci. Technol. 46, 9333–9341.

Koopmans  M.P.,  Schouten  S.,  Kohnen  M.E.L.  &  Sinninghe  

Damsté J.S. 1996: Restricted utility of aryl isoprenoids as  

indicators for photic zone anoxia. Geochim. Cosmochim. Acta 

60, 4873–4876.

Koopmans M.P., Rijpstra W.I.C., Klapwijk M.M., de Leeuw J.W., 

 Lewan  M.D.  &  Sinninghe  Damsté  J.S.  1999:  A  thermal  and 

chemical degradation approach to decipher pristane and phytane 

precursors in sedimentary organic matter. Org. Geochem. 30, 

1089–1104.

Kortenski  J.  &  Kostova  I.  1996:  Occurrence  and  morphology  of 

 pyrite in Bulgarian coals. Int. J. Coal Geol. 29, 273–290.

Laflamme R.E. & Hites R.A. 1978: The global distribution of poly-

cyclic aromatic hydrocarbons in recent sediments. Geochim. 

Cosmochim. Acta 42, 289–303.

Markova K., Zdravkov A., Bechtel A. & Stefanova M. 2017: Organic 

Geochemical characteristics of Bulgarian jet. Int. J. Coal Geol. 

181, 1–10.

Marynowski L., Kurkiewicz S., Rakociński M., & Simoneit B.R.T. 

2011a: Effects of weathering on organic matter: I. Changes in 

molecular composition of extractable organic compounds caused 

by  paleoweathering  of  a  Lower  Carboniferous  (Tournaisian) 

 marine black shale. Chem. Geol. 285, 144–156.

Marynowski L., Szełęg E., Jędrysek M.O. & Simoneit B.R.T. 2011b: 

Effects of weathering on organic matter. Part II: Fossil wood 

weathering and implications for organic geochemical and petro-

graphic studies. Org. Geochem. 42, 1076–1088.

Marynowski L., Kubik R., Uhl D. & Simoneit B.R.T. 2014: Molecu-

lar composition of fossil charcoal and relationship with incom-

plete combustion of wood. Org. Geochem. 77, 22–31.

Minčev D. 1978: Neue daten über die Hypogenese des Saprovitains 

(des Gagats) in Zentral-Nordbulgarien. Annu. l’Universite Sofia 

“St. Kliment Ohridski”, Fac. Geol. Geogr. livre 1 — Geol. 70, 

193–205 (in Bulgarian).

Minčev  D.  1980:  Der  Gagat  in  der  Moesicschen  Platte  (Bezirk 

 Plewen). Annu. l’Universite Sofia “St. Kliment Ohridski”, Fac. 

Geol. Geogr. livre 1 — Geol. 72, 247–268 (in Bulgarian).

Minčev D. 1982: Autochthonity and allochthonity in gagath genesis. 

Annu. l’Universite Sofia “St. Kliment Ohridski”, Fac. Geol. 

Geogr. livre 1 — Geol. 76, 69–78 (in Bulgarian).

Minčev D. & Nikolov Z. 1979: New deposits of saprovitrain (gagate) 

in Balkan coal Basin. Annu. l’Universite Sofia “St. Kliment 

Ohridski”, Fac. Geol. Geogr. livre 1 — Geol. 71, 303–315 (in 

Bulgarian).

Minčev D. & Šiškov G. 1986: Gagate phytoclast of superhigh micro-

hardness from Low Jurassic sediments in NW Bulgaria. Annu. 

l’Universite Sofia “St. Kliment Ohridski”, Fac. Geol. Geogr., 1, 

80, 98–113 (in Bulgarian).

Mincheva-Stefanova J. 1988: Strata-bound polymetalic formation, In: 

Dimitrov R. (Ed.): Lead and Zinc Deposits in Bulgaria. Tehnika, 

Sofia, 258.

Noble R.A., Alexander R., Kagi R.I. & Knox J. 1985: Tetracyclic di-

terpenoid hydrocarbons in some Australian coals, sediments and 

crude oils. Geochim. Cosmochim. Acta 49, 2141–2147.

Noble R.A., Alexander R., Kagi R. & Nox J. 1986: Identification of 

some diterpenoid hydrocarbons in petroleum. Org. Geochem. 

10, 825–829.

Otto A. & Simoneit B.R.T. 2001: Chemosystematics and diagenesis 

of terpenoids in fossil conifer species and sediment from the 

 Eocene Zeitz formation, Saxony, Germany. Geochim. Cosmo-

chim. Acta 65, 3505–3527.

Otto A. & Wilde V. 2001: Sesqui-, di-, and triterpenoids as chemosys-

tematic markers in extant conifers — A review. Bot. Rev. 67, 

141–238.

Peters K.E., Walters C.C. & Moldowan J.M. 2005: The Biomarker 

Guide, Volume 2: Biomarkers and Isotopes in the Petroleum 

 Exploration and Earth History. Cambridge University Press,  

1–702.

Petrova R., Minčev D. & Nikolov Z. 1985: Comparative investiga-

tions on Gagate and Vitrain from the Balkan coal basin. Int. J. 

Coal Geol. 5, 275–280.

Pfennig N. 1977: Phototrophic green and purple bacteria: a compara-

tive, systematic survey. Annu. Rev. Microbiol. 31, 275–290.

Radke M. & Welte D.H. 1981: The Methylphenanthrene Index (MPI): 

A maturity parameter based on aromatic hydrocarbons. In: 

 Advances in Organic Geochemistry, 504–512.

Radke M., Willsch H. & Welte D.H. 1980: Preparative hydrocarbon 

group type determination by automated medium pressure liquid 

chromatography. Anal. Chem. 52, 406–411.

Radke M., Willsch H., Leythaeuser D. & Teichmüller M. 1982: Aro-

matic components of coal: relation of distribution pattern to 

rank. Geochim. Cosmochim. Acta 46, 1831–1848.

Radke  M.,  Leythaeuser  D.  &  Teichmüller  M.  1984:  Relationship 

 between rank and composition of aromatic hydrocarbons for 

coals of different origins. Org. Geochem. 6, 423–430.

Radke M., Welte D.H. & Willsch H. 1986: Maturity parameters based 

on aromatic hydrocarbons: Influence of the organic matter type. 

Org. Geochem. 10, 51–63.

Radke M., Vriend S.P. & Ramanampisoa L.R. 2000: Alkyldibenzofu-

rans in terrestrial rocks: Influence of organic facies and matura-

tion. Geochim. Cosmochim. Acta 64, 275–286.

Rowell R.M., Pettersen R., Han J.S., Rowell J.S. & Tshabalala M.A. 

2005: Cell Wall Chemistry. In: Rowell R. (Ed.): Handbook of 

Wood Chemistry and Wood Composites. CRC Press, 35–72.

Schwark L. 1998: Geochemical characterization of Malm Zeta lami-

nated  carbonates  from  the  Franconian Alb,  SW-Germany  (II). 

Org. Geochem. 29, 1921–1952.

Schwark L. & Püttmann W. 1990: Aromatic hydrocarbon composi-

tion of the Permian Kupferschiefer in the Lower Rhine Basin, 

N.W. Germany. Org. Geochem. 16, 749–761.

Simoneit B.R.T. 1977: Diterpenoid compounds and other lipids in 

deep-sea sediments and their geochemical significance. Geo-

chim. Cosmochim. Acta 41, 463–476.

Simoneit B.R.T. 1986: Cyclic terpenoids of the geosphere, In: Johns 

R.B.  (Ed.):  Biological  Markers  in  the  Sedimentary  Record. 

 Elsevier, Amsterdam, 43–99.

Simoneit B.R.T. 1998: Biomarker PAHs in the Environment. In: 

 Neilson A.H. (Ed.): PAHs and Related Compounds. Springer- 

Verlag, Berlin, 175–221.

Sinninghe Damsté J.S., Kock-Van Dalen A.C., De Leeuw J.W., 

Schenck P.A., Guoying S. & Brassell S.C. 1987: The identifica-

tion of mono-, di- and trimethyl 2-methyl-2-(4,8,12-trimethyltri-

decyl)chromans and their occurrence in the geosphere. Geochim. 

Cosmochim. Acta 51, 2393–2400.

background image

74

ZDRAVKOV, AJDANLIJSKY, GROΒ and BECHTEL

GEOLOGICA CARPATHICA

, 2019, 70, 1, 62–74

Sinninghe Damsté J.S., Keely B.J., Betts S.E., Baas M., 

 

Maxwell J.R. & de Leeuw J.W. 1993: Variations in abundances 

and distributions of isoprenoid chromans and long-chain 

 

alkylbenzenes in sediments of the Mulhouse Basin: a molecular 

sedimentary  

record of palaeosalinity. Org. Geochem. 20, 

 

1201–1215.

Sinninghe  Damsté  J.S.,  Frewin  N.L.,  Kenig  F.,  &  De  Leeuw  J.W. 

1995: Molecular indicators for palaeoenvironmental change in  

a  Messinian  evaporitic  sequence  (Vena  del  Gesso,  Italy).  In: 

Variations in extractable organic matter of ten cyclically deposi-

ted marl beds. Org. Geochem. 23, 471–483.

Sinninghe Damsté J.S., Schouten S. & van Duin A.C.. 2001: Isore-

nieratene derivatives in sediments: possible controls on their dis-

tribution. Geochim. Cosmochim. Acta 65, 1557–1571.

Stojanović K., Jovančićević B., Vitorovi D., Pevneva G.S., Golovko 

J.A. & Golovko  A.K. 2007: New maturation parameters based 

on naphthalene and phenanthrene isomerization and dealkyla-

tion processes aimed at improved classification of crude oils 

(Southeastern Pannonian Basin, Serbia). Geochemistry Int. 45, 

781–797.

Stout  S.A.  &  Emsbo-Mattingly  S.D.  2008:  Concentration  and  

character of PAHs and other hydrocarbons in coals of varying 

rank — Implications for environmental studies of soils and 

 

sediments containing particulate coal. Org. Geochem. 39, 

 

801–819.

Strasser A., Pittet B., Hillgärtner H. & Pasquier J.-B. 1999: Deposi-

tional sequences in shallow carbonate-dominated sedimentary 

systems: concepts for a high-resolution analysis. Sediment. 

Geol. 128, 201–221.

Suárez-Ruiz  I.,  Iglesias  M.J.,  Jiménez A.,  Laggoun-Défarge  F.,  & 

 Prado J.G. 1994a: Petrographic and Geochemical Anomalies 

Detected in Spanish Jurassic Jet. In: Vitrinite Reflectance as 

 Maturity  Parameter.  American Chemical Society, 76–92.

Suárez-Ruiz  I.,  Jimenez  A.,  Iglesias  M.J.,  Laggoun-Defarge  F.  & 

 Prado J.G. 1994b: Influence of Resinite on Huminite Properties. 

Energy & Fuels 8, 1417–1424.

Sukh Dev 1989: Terpenoids, In: Rowe J.W. (Ed.): Natural Products of 

Woody Plants. Springer, Berlin, 691–807.

Summons R.E. 1993: Biogeochemical cycles: a review of fundamen-

tal aspects of organic matter formation, preservation, and com-

position. In: Engel M.H., & Macko S.A. (Eds.): Organic Geo-

chemistry — Principles and Applications. Plenum Press, 

NewYork, 3–21.

Summons R.E. & Powell T.G. 1987: Identification of aryl isoprenoids 

in source rocks and crude oils: Biological markers for the green 

sulphur bacteria. Geochim. Cosmochim. Acta 51, 557–566.

Taylor G.H., Teichmuller M., Davis A., Diessel C.F.K., Littke R. & 

Robert P. 1998: Organic petrology. Gebruder Borntraeger, 

 Berlin, Stuttgart, 1–704.

Taylor T.N., Taylor E.L. & Krings M. 2009: Paleobotany: The Biolo-

gy and Evolution of Fossil Plants. Second. Ed., Elsevier Aca-

demic Press, 1–1253.

ten Haven H.L., de Leeuw J.W., Rullkötter J. & Damsté J.S.S. 1987: 

Restricted utility of the pristane/phytane ratio as a palaeoenvi-

ronmental indicator. Nature 330, 641–643.

Tissot  B.P.  &  Welte  D.H.  1984:  Petroleum  Formation  and  Occur-

rence. Second Revised and Enlarged Edition. Springer-Verlag

1–720.

Tronkov D. 1981: Stratigraphy of the Triassic System in part of the 

Western Srednogorie (West Bulgaria). Geol. Balc. 11, 3–20 (in 

Russian).

Tronkov D. 1983: Stratigraphical problems of Iskar Carbonate Group 

(Triassic)  in  South-Western  Bulgaria.  Geol. Balc. 13, 91–100  

(in Russian).

van Aarssen B.G.K. 1999: Distributions of methylated naphthalenes 

in crude oils: indicators of maturity, biodegradation and mixing. 

Org. Geochem. 30, 1213–1227.

Volkman J.K. & Maxwell J.R. 1986: Acyclic isoprenoids as biolo-

gical markers. In: Johns R.B. (Ed.): Biological Markers in the 

Sedimentary Record. Elsevier, Amsterdam, 1–42.

Volkman  J.K., Alexander  R.,  Kagi  R.,  Rowland  S.  &  Sheppard  P. 

1984: Biodegradation of aromatic hydrocarbons in crude oils 

from the Barrow Sub-basin of Western Australia. Org. Geochem. 

6, 619–632.

Wakeham S.G., Schaffner C. & Giger W. 1980: Polycyclic aromatic 

hydrocarbons in Recent lake sediments — II. Compounds 

 

derived from biogenic precursors during early diagenesis. 

 Geochim. Cosmochim. Acta 44, 415–429.

Wang R., Sun R., Liu G., Yousaf B., Wu D., Chen J. & Zhang H. 

2017: A review of the biogeochemical controls on the occurrence 

and  distribution  of  polycyclic  aromatic  compounds  (PACs)  in 

coals. Earth-Sci. Rev. 171, 400–418.

White C.M., Douglas L.J., Schmidt C.E. & Hackett M. 1988: Forma-

tion of polycyclic thiophenes from reaction of selected polycy-

clic aromatic hydrocarbons with elemental sulfur and/or pyrite 

under mild conditions. Energy and Fuels 2, 220–223.

Wiese J.R.G. & Fyfe W.S. 1986: Occurrence of iron sulfides in Ohio 

coals. Int. J. Coal Geol. 6, 251–276.

Yunker M.B., Macdonald R.W., Vingarzan R., Mitchell R.H., 

 

Goyette D. & Sylvestre S. 2002: PAHs in the Fraser River basin: 

A critical appraisal of PAH ratios as indicators of PAH source 

and composition. Org. Geochem. 33, 489–515.