22.11.2012 Views

aspects of organic geochemistry of nigerian coal university of ibadan

aspects of organic geochemistry of nigerian coal university of ibadan

aspects of organic geochemistry of nigerian coal university of ibadan

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

ASPECTS OF ORGANIC GEOCHEMISTRY OF NIGERIAN COAL<br />

AS A POTENTIAL SOURCE-ROCK OF PETROLEUM<br />

TAOFIK ADEWALE ADEDOSU<br />

B.Tech. Chemistry (Ogbomoso), M.Sc. Organic Chemistry (Ibadan)<br />

A Thesis in the Department <strong>of</strong> Chemistry<br />

Submitted to the Faculty <strong>of</strong> Science in Partial Fulfillment <strong>of</strong> the requirements for<br />

the<br />

Degree <strong>of</strong><br />

DOCTOR OF PHILOSOPHY<br />

<strong>of</strong> the<br />

UNIVERSITY OF IBADAN<br />

2009<br />

1


DEDICATION<br />

This thesis is dedicated firstly to Almighty Allah, the Omnipotent and Omnipresent who<br />

guided me both physically and spiritually throughout the duration <strong>of</strong> this programme and to my<br />

wife and my daughters; Aliyah and Atiyah.<br />

2


ACKNOWLEDGEMENTS<br />

All praises are due to Almighty Allah the Creator, the Cherisher and the Protector <strong>of</strong> the entire<br />

Universe. I thank Him immensely for sparing my life and giving me sound health throughout the<br />

duration <strong>of</strong> this research work. Certainly, He is the most glorious.<br />

My gratitude goes to the Department <strong>of</strong> chemistry for the opportunity given me to have sound<br />

training in the field <strong>of</strong> Organic <strong>geochemistry</strong>. My special gratitude goes to the Head <strong>of</strong> Department,<br />

Pr<strong>of</strong>. R.O. Oderinde and the entire members <strong>of</strong> staff (both academic and non-academic) for their<br />

various contributions toward the successful completion <strong>of</strong> this thesis. My pr<strong>of</strong>ound appreciation<br />

goes to my supervisors, Pr<strong>of</strong>. O. Ekundayo and Dr. O.O. Sonibare for their contributions, advice,<br />

support, prayers and thorough supervision <strong>of</strong> this project. May the Lord in His infinite mercies<br />

continue to be with them and their families. Once more, special thanks go to Dr. O.O. Sonibare for<br />

given me solid foundation in Organic <strong>geochemistry</strong>. Also your dual role as Teacher/Mentor is<br />

greatly acknowledged, may the good Lord continue to enrich you with knowledge and bless you.<br />

My appreciation also goes to the immediate past Head <strong>of</strong> Department, Pr<strong>of</strong>. O.O. Osibanjo for his<br />

fatherly advice and encouragement throughout the duration <strong>of</strong> this programme. I thank Dr (Mrs.)<br />

O.O. Aiyelagbe for her motherly support towards the successful completion <strong>of</strong> this research work.<br />

This research work has benefited immensely from the Chinese Academy <strong>of</strong> Science and Third<br />

World Academy <strong>of</strong> Science (CAS-TWAS) Postgraduate Fellowship awards. This afforded me<br />

opportunity to carry out my research work with the ‘State <strong>of</strong> Art’ equipments at the Institute <strong>of</strong><br />

Geology and Geophysics, Lanzhou, China between April 2007 and January 2008 under the able<br />

supervision <strong>of</strong> Pr<strong>of</strong>. Jincai Tuo. I thank Pr<strong>of</strong>. Jincai Tuo for the kindness and love extended to me<br />

during the period <strong>of</strong> the Fellowship.<br />

My appreciation goes to Pr<strong>of</strong>. Meng, Dr (Mrs.) Yong Li (Associate Pr<strong>of</strong>essor), all the<br />

Laboratory Technologists and some postgraduate students at the institute especially, Ma Wanyun,<br />

Li Zhoping, Zhang Mingfeng, Mr. Wang and Yu Wen Xio. I also thank Joan Q. Li and Ashley<br />

Wang for their assistance during my stay in China.<br />

The unforgettable assistance rendered to me by Dr Ehinola <strong>of</strong> Geology Department, University<br />

<strong>of</strong> Ibadan both in Nigeria and China is highly appreciated. The good Lord used him positively<br />

towards my achievement.<br />

The Almighty God will reward him abundantly. My sincere appreciation goes to Mr. Adabanija<br />

3


(LAUTECH) and Mr. Mathew (Itakpe) for their assistance in getting the samples used for this<br />

research work. I appreciate Nigerian Coal Corporation, Enugu, Nigeria for the supply <strong>of</strong> some <strong>of</strong><br />

the <strong>coal</strong> samples.<br />

Special thanks go to Pr<strong>of</strong>s. Olawore, Faboya, Odunola, Olajire and Ayodele for teaching me<br />

foundation in chemistry. The great work done by Pr<strong>of</strong>. C.M. Ekweozor, Pr<strong>of</strong>. S.O. Akande and<br />

Pr<strong>of</strong>. N. Obaje on Benue Trough, which really assisted me during this research work, is greatly<br />

acknowledged.<br />

The encouragement and prayers received from my mother, siblings, uncles, in-law (Mr. and<br />

Mrs. Ganiyu Obembe), Mr. Azeez, Dr. Yekeen, Dr Gbenga Bello, Dr. (Miss) Adekunle, Dr. (Mrs.)<br />

Oladipo, Dr. Lateef Agbaje, Dr Olaide Lawal, Dr. Niyi Afolabi, Mr. and Mrs. Jelil Bello (Canada),<br />

Dr. Ademola Idowu (USA), Mrs. A.O. Ibrahim and Dr. (Mrs.) Bello are highly commendable.<br />

My appreciation goes to LAUTECH management for granting me study leave to take up the<br />

CAS-TWAS Postgraduate Fellowship. Also both academic and non-academic members <strong>of</strong> staff <strong>of</strong><br />

Chemistry department, LAUTECH are highly appreciated for their various contributions towards<br />

the completion <strong>of</strong> this work.<br />

Lastly, I give a big thank to Queen Haleema, my darling wife for her love, care, support,<br />

encouragement and understanding and to my beautiful girls, Aliya and Atiya.I pray that Allah<br />

would grant us sound health, long life and prosperity to reap the fruit <strong>of</strong> our labour.<br />

4


CERTIFICATION<br />

This is to certify that Ta<strong>of</strong>ik Adewale ADEDOSU in the Department <strong>of</strong> Chemistry, University <strong>of</strong><br />

Ibadan, Ibadan carried out the work reported in this thesis under our supervision.<br />

…………………………………… ………………………….<br />

Supervisor Supervisor<br />

Olusegun Ekundayo Oluwadayo Olatunde Sonibare<br />

B.Sc (Lagos), Ph.D(London), FAS B.Sc (Ago-woye), PhD(Ibadan)<br />

Pr<strong>of</strong>essor <strong>of</strong> Chemistry Senior Lecturer<br />

Department <strong>of</strong> Chemistry, Department <strong>of</strong> Chemistry,<br />

University <strong>of</strong> Ibadan, University <strong>of</strong> Ibadan,<br />

Ibadan. Ibadan.<br />

5


ABSTRACT<br />

Several studies have suggested that <strong>coal</strong> can be a potential petroleum source rock in<br />

deltaic and other basins <strong>of</strong> the world. However despite its vast deposits in Nigeria,<br />

there is a depth <strong>of</strong> information on its use as a source-rock <strong>of</strong> petroleum. This study<br />

evaluated the source rock potential <strong>of</strong> Nigerian <strong>coal</strong> based on chemical and isotope<br />

composition <strong>of</strong> biomarkers and polycyclic aromatic hydrocarbons in the <strong>coal</strong>s.<br />

Twenty-two <strong>coal</strong> and carbonaceous shale samples were collected from four boreholes in Mamu<br />

and Awgu Formations <strong>of</strong> Lower and Middle Benue Trough, Nigeria. The samples were subjected<br />

to Elemental analysis using Carlo Erba 1108 CHNS-O Analyzer. The source rock potential <strong>of</strong> the<br />

samples were determined using Rock-Eval analysis. Biomarkers in the aliphatic and polar<br />

fractions and aromatic hydrocarbon distributions in the samples were studied using Gas<br />

Chromatography- Mass Spectrometry (GC-MS). The Carbon isotope analysis <strong>of</strong> individual n-<br />

alkanes in the aliphatic fraction was performed using Gas Chromatography-Combustion-Isotope<br />

Ratio Mass Spectrometer (GC-IRMS). The vitrinite reflectance measurements <strong>of</strong> the samples were<br />

taken using Zeiss standard universal reflected microscope.<br />

All the <strong>coal</strong>s analyzed contained the minimum <strong>of</strong> 0.5wt.% and 2mg/g <strong>of</strong> Total Organic Carbon<br />

(TOC) and Genetic Potential (GP) respectively <strong>of</strong> <strong>organic</strong> matter required to serve as good source<br />

rock for oil and gas. Several plots from the elemental and Rock-Eval pyrolysis classified the<br />

<strong>organic</strong> matter in Awgu and Mamu samples as type III and type II/III kerogen respectively. The<br />

abundance <strong>of</strong> hopanes, homohopanes (C31-C35), benzohopanes and C29 steranes and diasteranes in<br />

most <strong>of</strong> the samples indicate terrestrial plant, phytoplankton and cyanobacteria contributions to the<br />

<strong>organic</strong> matter that formed the <strong>coal</strong>. High (Pr/Ph) ratio (1.73-12.47), distributions <strong>of</strong> polyaromatic<br />

hydrocarbons and isotopic distribution <strong>of</strong> individual alkanes showed that Awgu samples consisted<br />

<strong>of</strong> mixed terrestrial/marine <strong>organic</strong> matter deposited under oxic condition in lacustrine-<br />

fluvial/deltaic depositional environment. Mamu samples consisted <strong>of</strong> terrestrial <strong>organic</strong> matter<br />

with marine incursion deposited under oxic/suboxic-oxic in lacustrine-fluvial/deltaic environment.<br />

The vitrinite reflectance values (0.48-1.15%Ro) and all the maturity parameters derived from the<br />

elemental, Rock-Eval analysis and biomarker distributions showed that Awgu samples are in the<br />

late oil window, and Mamu samples have low thermal maturity status.<br />

6


The distribution patterns <strong>of</strong> C32-C35 benzohopanes in Mamu samples confirmed the redox<br />

condition <strong>of</strong> <strong>organic</strong> matter deposition within the Formation. The presence <strong>of</strong> unsaturated<br />

oleanenes in Mamu samples further confirmed their thermal immaturity status.<br />

The kerogen types <strong>of</strong> the samples have capacity to generate both oil and gas. Benzohopanes<br />

(C32-C35) and three oleanene isomers; olean-12-ene, olean-13 (18)-ene and olean-18-ene were<br />

present in samples collected from Mamu Formation. These compounds were detected for the first<br />

time in Nigerian <strong>coal</strong>. Awgu Formation <strong>coal</strong>s might have generated gases into yet-to-be identified<br />

reservoir. Coals from Mamu Formation have potential to generate both oil and gas but are<br />

presently immature to have formed any significant hydrocarbons.<br />

Keywords: Biomarkers, Carbonaceous shale, Benue-Trough, Carbon Isotopes, Hydrocarbon<br />

potential.<br />

Word count: 460.<br />

7


TABLE OF CONTENTS<br />

Subject Pages<br />

Title ……………………………………………….………………………………i<br />

Dedication………………………………………….……………………………...ii<br />

Acknowledgements………………………………….……………………………iii<br />

Certification………………………………………….…………………………….v<br />

Abstract……………………………………………………………………………vi<br />

Table <strong>of</strong> contents………………………………………………………………….viii<br />

List <strong>of</strong> figures……………………………………………………………………...xii<br />

List <strong>of</strong> tables……………………………………………………………………….xv<br />

CHAPTER ONE INTRODUCTION<br />

1.1 General Introduction………………………………………………………….……1<br />

1.2 Coal Formation…………………………………………………………………….2<br />

1.2.1 Peatification……………………………………………………………………..2<br />

1.2.2 Coalification……………………………………………………………………..3<br />

1.3 Coal Structure……………………………………………………………………...3<br />

1.4 Petroleum potential <strong>of</strong> Coal………………………………………………………..4<br />

1.5 Coal Reserves in Nigeria…………………………………………………………..5<br />

1.6 Petroleum Source Rock Evaluation………………………………………………..9<br />

1.6.1 Quantity <strong>of</strong> Organic Matter……………………………………………………...9<br />

1.6.2 Quality <strong>of</strong> Organic Matter……………………………………………………...10<br />

1.6.2.1 Optical Methods……………………………………………………………...10<br />

1.6.2.2 Physico-chemical Methods…………………………………………………..10<br />

1.6.2.3 Chemical Methods……………………………………………………………11<br />

1.6.2.4 Carbon Isotopic Composition………………………………………………...12<br />

1.6.3 Maturity <strong>of</strong> Organic Matter…………………………………………………….12<br />

1.6.3.1 Rock-Eval Pyrolysis………………………………………………………….12<br />

8


1.6.3.2 Biomarker Analysis…………………………………………………………..13<br />

1.7 Aims and Objectives <strong>of</strong> this Study ………………………………………………14<br />

CHAPTER TWO LITERATURE REVIEW<br />

2.1 Biomarker Geochemistry………………………………………………………...15<br />

2.1.1 Normal and branched Alkanes…………………………………………………15<br />

2.1.2 Acyclic Isoprenoids…………………………………………………………….16<br />

2.1.3 n-Alkanol, Alkanoic acids and Fatty acids……………………………………..18<br />

2.1.4 Alkanones………………………………………………………………………19<br />

2.1.5 Sesquiterpenoids………………………………………………………………..20<br />

2.1.6 Diterpenoids……………………………………………………………………24<br />

2.1.7 Triterpenoids…………………………………………………………………...25<br />

2.1.7.1 Tricyclic and tetracyclic triterpanes………………………………………….25<br />

2.1.7.2 Pentacyclic triterpanes (Non-Hopanoids)…………………………………... 26<br />

2.1.7.3 Hopanoids……………………………………………………………………28<br />

2.1.8 Steroids/Steranes……………………………………………………………….29<br />

2.2 Non-Biomarker Compounds……………………………………………………..32<br />

2.2.1 Polynuclear aromatic hydrocarbons (PAHs)…………………………………...32<br />

2.2.2 Aromatic Sulphur Compounds…………………………………………………33<br />

2.3 Isotope Geochemistry…………………………………………………………….35<br />

2.3.1 Stable Carbon Isotope………………………………………………………….36<br />

2.3.2 Hydrogen Isotopes……………………………………………………………...39<br />

2.3.3 Sulphur Isotopes………………………………………………………………..41<br />

2.3.4 Oxygen Isotopes………………………………………………………………..41<br />

2.4 Analytical Methods Employed in Geochemical Analyses……………………….42<br />

2.4.1 Pyrolysis………………………………………………………………………..42<br />

2.4.1.1 Flash or Rapid temperature-Programmed Pyrolysis…………………………43<br />

9


2.4.1.2 Hydrous Pyrolysis……………………………………………………………43<br />

2.4.1.3 Rock-Eval Pyrolysis………………………………………………………….44<br />

2.4.2Gas Chromatography/Mass Spectrometry………………………………………47<br />

2.4.3Gas Chromatography-Isotope Ratio Mass Spectrometer (GC-IRMS)………….47<br />

2.5 Geology <strong>of</strong> Benue Trough………………………………………………………..50<br />

2.5.1 Geological Setting……………………………………………………………...50<br />

2.5.2 Lithological Description………………………………………………………..53<br />

CHAPTER THREE EXPERIMENTAL<br />

3.1 SAMPLING AND SAMPLE PREPARATION………………………………..55<br />

3.1.1 Sampling……………………………………………………………………….55<br />

3.1.2 Extraction <strong>of</strong> Soluble Organic Matter…………………………………………58<br />

3.1.3 Asphalthene Isolation………………………………………………………….58<br />

3.1.4 Fractionation…………………………………………………………………..58<br />

3.1.5 Derivatisation <strong>of</strong> Polar fraction………………………………………………..59<br />

3.2 ANALYTICAL METHODS……………………………………………………59<br />

3.2.1 Leco Analysis…………………………………………………………………59<br />

3.2.2 Elemental Analysis……………………………………………………………59<br />

3.2.3 Rock Eval Analysis……………………………………………………………60<br />

3.2.4Gas Chromatography-Mass Spectrometry Analysis (GC-MS)………………...60<br />

3.2.5Gas Chromatography-Isotope Ratio Mass Spectrometry (GC-IRMS)………...61<br />

3.2.6 Vitrinite Reflectance Measurement…………………………………………...61<br />

CHAPTER FOUR RESULTS AND DISCUSSIONS<br />

4.1 ELEMENTAL ANALYTICAL DATA……………………………………….…62<br />

4.1.1 Organic Matter Source and Depositional………………………………………62<br />

4.1.1.1 Awgu Formation……………………………………………………………..62<br />

10


4.1.1.2 Mamu Formation…………………………………………………………….63<br />

4.2 Source Rock Evaluation <strong>of</strong> Nigerian Coal……………………………………….66<br />

4.2.1 Organic Matter Concentration…………………………………………………66<br />

4.2.1.1 Awgu Formation……………………………………………………………..66<br />

4.2.1.2 Mamu Formation…………………………………………………………….68<br />

4.2.2 Organic Matter Quality………………………………………………………...68<br />

4.2.2.1 Awgu Formation……………………………………………………………..68<br />

4.2.2.2 Mamu Formation…………………………………………………………….70<br />

4.2.3 Thermal Maturity <strong>of</strong> Organic matter…………………………………………..70<br />

4.2.3.1 Awgu Formation……………………………………………………………..70<br />

4.2.3.2 Mamu Formation…………………………………………………………….73<br />

4.3 Biomarker Geochemistry <strong>of</strong> the Nigerian Coal………………………………….77<br />

4.3.1 n-Alkane and isoprenoid distributions…………………………………………77<br />

4.3.1.1 Awgu Formation……………………………………………………………..77<br />

4.3.1.2 Mamu Formation…………………………………………………………….87<br />

4.3.2 Fatty acids and alkanones……………………………………………………...87<br />

4.3.2.1 Awgu Formation……………………………………………………………..88<br />

4.3.2.2 Mamu Formation…………………………………………………………….98<br />

4.3.3 Tricyclic and C24 tetracyclic terpanes………………………………………….98<br />

4.3.3.1 Awgu Formation……………………………………………………………..98<br />

4.3.3.2 Mamu Formation……………………………………………………………104<br />

4.3.4 Hopanes and homohopanes…………………………………………………...104<br />

4.3.4.1 Awgu Formation……………………………………………………………104<br />

4.3.4.2 Mamu Formation……………………………………………………………113<br />

4.3.5 Steranes……………………………………………………………………….118<br />

4.3.5.1 Awgu Formation…………..………………………………………………..118<br />

4.3.5.2 Mamu Formation……………………………………………………………126<br />

4.4 Polycyclic aromatic hydrocarbons in Nigerian <strong>coal</strong>…………………………….128<br />

4.4.1 Naphthalene and alkylnaphthalenes…………………………………………..128<br />

4.4.1.1 Awgu Formation……………………………………………………………128<br />

11


4.4.1.2 Mamu Formation……………………………………………………………134<br />

4.4.2 Phenanthrene and alkylphenanthrenes………………………………………..134<br />

4.4.2.1 Awgu Formation……………………………………………………………134<br />

4.4.2.2 Mamu Formation……………………………………………………………142<br />

4.4.3 Dibenzothiophene and alkyldibenzothiophenes………………………………142<br />

4.4.3.1 Awgu Formation……………………………………………………………143<br />

4.4.3.2 Mamu Formation……………………………………………………………151<br />

4.5 Carbon Isotopic Composition <strong>of</strong> Nigerian Coal………………………………...152<br />

4.5.1 Origin and Depositional Environment <strong>of</strong> Organic matter…………………….152<br />

4.5.1.1 Awgu Formation……………………………………………………….…...152<br />

4.5.1.2 Mamu Formation……………………………………………………………155<br />

CHAPTER FIVE<br />

5.0 Summary and conclusion..……………………………………………………..159<br />

REFERENCES……………………………………………………………………..161<br />

APPENDIX…………………………………………………………………………182<br />

List <strong>of</strong> Figures<br />

Fig.1.1: Location <strong>of</strong> Anambra Basin……………………………………………….…7<br />

Fig.2.1: Chemical structures <strong>of</strong> some compounds cited in the literature…………….21<br />

Fig.2.2: Transformation <strong>of</strong> steranes at C-20 during thermal maturity……………….31<br />

Fig.2.3: Scheme <strong>of</strong> Rock-Eval analysis……………………………………………...45<br />

Fig.2.4: Geological map <strong>of</strong> Benue Trough, Nigeria………………………………….51<br />

Fig.2.5: Stratigraphic sequence <strong>of</strong> Benue Trough, Nigeria…………………………..52<br />

Fig.3.1: Geological map and Lithographic section <strong>of</strong> BH94<br />

and BH120 <strong>of</strong> Lafia-Obi <strong>coal</strong>, Awgu Formation…………………………….56<br />

Fig.3.2: Lithographic section <strong>of</strong> Onyeama and Okaba Mines in Mamu Formation…57<br />

Fig.4.1: Plot <strong>of</strong> Atomic H/C against O/C <strong>of</strong> Coal Samples from Benue Trough…….65<br />

Fig.4.2: Plot <strong>of</strong> HI vs. OI <strong>of</strong> Coal Samples from Benue Trough, Nigeria……………69<br />

12


Fig.4.3: Plots <strong>of</strong> S2 vs. TOC <strong>of</strong> Coal Samples from Benue Trough, Nigeria…….…..71<br />

Fig.4.4: Plots <strong>of</strong> Tmax vs. HI <strong>of</strong> Coal Samples from Benue Trough, Nigeria…….…72<br />

Fig.4.5: Plots <strong>of</strong> HI vs. Tmax <strong>of</strong> Coal Samples from Awgu Formation……..………74<br />

Fig.4.6: Plots <strong>of</strong> HI vs. Tmax <strong>of</strong> Coal Samples from Mamu Formation……..………75<br />

Fig.4.7: Plots <strong>of</strong> PI vs. Tmax <strong>of</strong> Coal Samples from Benue Trough, Nigeria………..76<br />

Fig.4.8: m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Awgu samples<br />

showing the distribution <strong>of</strong> n-Alkanes……………………………………...78<br />

Fig.4.9: m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu<br />

Formation sample (Okaba) showing the distribution <strong>of</strong> n-Alkanes……….....80<br />

Fig.4.10: m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu<br />

Formation sample (Onyeama) showing the distribution <strong>of</strong> n-Alkanes…..…..82<br />

Fig.4.11: Plots <strong>of</strong> Pr/nC17 against Ph/nC18 <strong>of</strong> Awgu and Mamu Formation samples..85<br />

Fig.4.12: Plots <strong>of</strong> CPI against OEP <strong>of</strong> Awgu and Mamu Formation samples…….…86<br />

Fig.4.13: m/z 74 Mass chromatograms showing the distribution <strong>of</strong> n-fatty acids<br />

in Awgu samples…………………………………………………………....89<br />

Fig.4.14: m/z 74 Mass chromatograms showing the distribution <strong>of</strong> n-fatty acids<br />

in Mamu samples (Okaba)…..……………………………………………...91<br />

Fig.4.15: m/z 74 Mass chromatograms showing the distribution <strong>of</strong> n-fatty acids<br />

in Mamu samples (Onyeama)…..…………………………………………..92<br />

Fig.4.16: m/z 58 Mass chromatograms showing the distribution <strong>of</strong> alkan-2-ones<br />

in Awgu samples…………………………………………………………….94<br />

Fig.4.17: m/z 58 Mass chromatograms showing the distribution <strong>of</strong> alkan-2-ones<br />

in Mamu samples (Okaba)..………………………………………………...95<br />

Fig.4.18: m/z 58 Mass chromatograms showing the distribution <strong>of</strong> alkan-2-ones<br />

in Mamu samples (Onyeama)..…………………………………………….96<br />

Fig.4.19: m/z 191 Mass chromatograms showing the distribution <strong>of</strong> tricyclic<br />

and tetracyclic terpanes in Awgu samples……………………………...100<br />

Fig.4.20: m/z 191 Mass chromatograms showing the distribution <strong>of</strong> tricyclic<br />

and tetracyclic terpanes in Mamu Formation samples (Okaba)…………..101<br />

13


Fig.4.21: m/z 191 Mass chromatograms showing the distribution <strong>of</strong> tricyclic<br />

and tetracyclic terpanes in Mamu Formation samples (Onyeama)……….102<br />

Fig.4.22: m/z 191 Mass chromatograms showing the distribution <strong>of</strong> hopanes<br />

in Awgu samples………………………………………………………….106<br />

Fig.4.23: m/z 191 Mass chromatograms showing the distribution <strong>of</strong> hopanes<br />

and benzohopanes in Mamu Formation samples (Okaba).………………107<br />

Fig.4.24: m/z 191 Mass chromatograms showing the distribution <strong>of</strong> hopanes<br />

and benzohopanes in Mamu Formation samples (Onyeama).……….……108<br />

Fig.4.25:Mass fragmentogram and spectra <strong>of</strong> Olean-18-ene in Okaba<br />

Mine samples………………………………………………………………114<br />

Fig.4.26:Mass fragmentogram and spectra <strong>of</strong> Olean-13(18)-ene in Okaba<br />

Mine samples………………………………………………………………115<br />

Fig.4.27:Mass fragmentogram and spectra <strong>of</strong> Olean-12-ene in Okaba<br />

Mine samples………………………………………………………………116<br />

Fig.4.28: m/z 217 Mass chromatograms showing the distribution <strong>of</strong> steranes<br />

and diasteranes in Awgu samples………………………………………….119<br />

Fig.4.29: m/z 217 Mass chromatograms showing the distribution <strong>of</strong> steranes<br />

and diasteranes in Mamu samples (Okaba)……………………………..…120<br />

Fig.4.30: m/z 217 Mass chromatograms showing the distribution <strong>of</strong> steranes<br />

and diasteranes in Mamu samples (Onyeama)……..……………………..121<br />

Fig.4.31:Ternary plots <strong>of</strong> C27, C28 and C29 steranes distributions in Nigerian Coal..123<br />

Fig.4.32: Ternary plots <strong>of</strong> C27, C28 and C29 diasteranes distributions in<br />

Nigerian Coal.…………………………………………………………..…124<br />

Fig.4.33: Plots <strong>of</strong> 22S/22S+22R C32 hopanes against C29αββ/αββ+ααα steranes….125<br />

Fig.4.34: m/z 156, 170 Mass chromatograms showing the distribution <strong>of</strong> naphthalene<br />

and alkylnaphthalenes in Awgu Samples………………………………….129<br />

Fig.4.35: m/z 156, 170 Mass chromatograms showing the distribution <strong>of</strong> naphthalene<br />

and alkylnaphthalenes in Mamu Samples (Okaba)………………………….131<br />

14


Fig.4.36: m/z 156, 170 Mass chromatograms showing the distribution <strong>of</strong> naphthalene<br />

and alkylnaphthalenes in Mamu Samples (Onyeama)………………………132<br />

Fig.4.37: m/z 178, 192, 206 Mass chromatograms showing the distribution<br />

<strong>of</strong> phenanthrene and alkylphenanthrenes in Awgu Samples………………136<br />

Fig.4.38: m/z 178, 192, 206 Mass chromatograms showing the distribution<br />

<strong>of</strong> phenanthrene and alkylphenanthrenes in Mamu Samples (Okaba)……..137<br />

Fig.4.39: m/z 178, 192, 206 Mass chromatograms showing the distribution<br />

<strong>of</strong> phenanthrene and alkylphenanthrenes in Mamu Samples (Onyeama)..138<br />

Fig.4.40: Organic matter source discrimination in the samples………………….…141<br />

Fig.4.41: m/z 184, 198, 212, 226 Mass chromatograms <strong>of</strong> dibenzothiophene<br />

and alkyldibenzothiophenes in Awgu Samples…………………………..144<br />

Fig.4.42: Cross plot <strong>of</strong> total Sulphur vs. dibenzothiophene/phenanthrene (DBT/P)<br />

for Nigerian Coal………………………..……………………………..…147<br />

Fig.4.43: Cross plot <strong>of</strong> total Sulphur vs. methyldibenzothiophene/methylphenanthrene<br />

(MDBT/MP) for Nigerian Coal……..….……………………………..…148<br />

Fig.4.44: Cross plot <strong>of</strong> dibenzothiophene/phenanthrene (DBT/P)<br />

vs. pristane/phytane for Nigerian Coal…….……………………………...149<br />

Fig.4.45: Cross plot <strong>of</strong> methyldibenzothiophene/methylphenanthrene (MDBT/MP)<br />

vs. pristane/phytane for Nigerian Coal…….……………………………...150<br />

Fig.4.46: Carbon Isotopic distribution <strong>of</strong> individual n-alkanes in Awgu Samples…154<br />

Fig.4.47: Carbon Isotopic distribution <strong>of</strong> individual n-alkanes<br />

in Okaba samples, Mamu…………………………………………………157<br />

Fig.4.48: Carbon Isotopic distribution <strong>of</strong> individual n-alkanes<br />

in Onyeama samples, Mamu……………….……………………………..158<br />

15


List <strong>of</strong> Tables<br />

Table 1.1: Coal and Lignite Reserves <strong>of</strong> Nigeria…………………………………..…8<br />

Table 2.1: Relevant Characteristics <strong>of</strong> the Light Stable Isotopes…………………....38<br />

Table 2.2: Rock-Eval Parameters………………………………………………….…46<br />

Table 4.1: Elemental Analysis and Vitrinite Reflectance Data <strong>of</strong> Nigerian Coal…....64<br />

Table 4.2: TOC and Rock Eval Pyrolysis Data………………………………………67<br />

Table 4.3: n-Alkanes and Isoprenoids Parameters…………………….……………..84<br />

Table 4.4: Parameters calculated from n-Fatty acids and alkanones composition<br />

<strong>of</strong> Nigerian Coal…………………………………………………………...97<br />

Table 4.5: Tri- and Tetracyclic terpanes source and depositional<br />

environment parameters…………………………………………………103<br />

Table 4.6: Peak identities on m/z 191 mass chromatograms………………………..109<br />

Table 4.7: Source and depositional environment parameters computed from<br />

hopane and sterane distributions in the <strong>coal</strong>s……………………………111<br />

Table 4.8: Maturity parameters computed from hopanes and steranes distributions<br />

in the <strong>coal</strong>s……………………………………………………………….112<br />

Table 4.9: Peak identities on m/z 217 mass chromatograms <strong>of</strong> Nigerian <strong>coal</strong>……..122<br />

Table 4.10: Peak identities on m/z 156 and 170 mass chromatograms……………133<br />

Table 4.11: Peak identities on m/z 178, 192, 206 mass chromatograms ……..……139<br />

Table 4.12: Source and Depositional Environment and maturity parameters<br />

derived from phenanthrene and dimthyldibenzothiophene and<br />

their alkyl derivatives…………………………………………………..140<br />

Table 4.13: Peak identities on m/z 184,198,212,226 mass chromatogram ………...146<br />

Table 4.14: Carbon Isotopic Composition <strong>of</strong> n-alkanes in Awgu Samples<br />

(δ 13 ‰PDB)……………………………………………………………….153<br />

Table 4.15: Carbon Isotopic Composition <strong>of</strong> n-alkanes in Mamu Samples<br />

(δ 13 ‰PDB)……………………………………………………………….156<br />

16


1.1 Introduction<br />

CHAPTER ONE<br />

INTRODUCTION<br />

Organic <strong>geochemistry</strong> deals with the process governing the origin and<br />

fate <strong>of</strong> <strong>organic</strong> materials such as petroleum (crude oil and natural gas), <strong>coal</strong>,<br />

oil shale and tar sands. Coals are remnants <strong>of</strong> terrestrial higher plants formed<br />

under non-marine and paralic conditions (Tissot and Welte, 1984) and found<br />

at its site <strong>of</strong> deposition as a solid and a relatively pure massive <strong>organic</strong><br />

substance. Many studies have suggested the possibility <strong>of</strong> non-marine <strong>coal</strong><br />

sediments having potential capacity to generate petroleum (Katz,<br />

1983;Betrand et al., 1986; Boreham and Powell, 1991; Hunt 1991, 1996;<br />

Hutton et al., 1994). Coal, the world’s most abundant, accessible and versatile<br />

source <strong>of</strong> fossil energy was brought to the forefront <strong>of</strong> the global energy scene<br />

by the industrial revolution <strong>of</strong> the 10 th century (Balogun et al., 2003). The<br />

largest reserves <strong>of</strong> <strong>coal</strong> were deposited in the late Carboniferous and Permian<br />

with lesser amounts in the Jurassic through the Tertiary (Kotarba et al., 2002).<br />

During the past 20 years, <strong>coal</strong>s and shales containing <strong>coal</strong>y <strong>organic</strong> matter<br />

(<strong>coal</strong>y shales) have received increasing attention as commercial quantity <strong>of</strong><br />

petroleum (oil and gas) have been discovered in many basins around the<br />

world containing these type <strong>of</strong> <strong>organic</strong> matter (Hendrix et al., 1995; Mario et<br />

al., 1997; Petersen et al., 2000; Bechtel et al., 2001; Kotarba et al., 2002;<br />

Boreham et al., 2003). Despite the huge reserves <strong>of</strong> <strong>coal</strong> in Nigeria, its<br />

petroleum potential has not been extensively studied.<br />

1.2 Coal Formation<br />

17


Coal is a brown to black combustible rock that originated by accumulation<br />

and subsequent physical and chemical alteration <strong>of</strong> plant materials over long<br />

period <strong>of</strong> time, and on moisture free basis contains not more than 50 % mineral<br />

matter.It is generally accepted that <strong>coal</strong> originated from plant debris including<br />

ferns, trees, bark, leaves, roots and seeds some <strong>of</strong> which accumulated and settled<br />

in swamps (Vlado and Valkovic, 1983). The result <strong>of</strong> accumulation <strong>of</strong> in-situ<br />

residues and important debris in swamps leads to the formation <strong>of</strong> peat.<br />

The <strong>coal</strong> is <strong>of</strong>ten interbedded with shale, sandstone and other sedimentary rocks.<br />

The two main phases involved in <strong>coal</strong> formation are briefly discussed as follows:<br />

1.2.1 Peatification<br />

Peat is being formed today in marshes and bogs. The plant debris<br />

accumulated in various wet environment, commonly called peat swamps, where<br />

dead plants were largely protected from decay by high water table and oxygen<br />

deficient water. The accumulating spongy, water saturated, plant derived <strong>organic</strong><br />

material known, as peat is the precursor <strong>of</strong> <strong>coal</strong>. Less than 10 % <strong>of</strong> plant debris<br />

deposited in a typical peat-forming environment accumulates as peat, the rest<br />

being recycled by the associated microbial community (bacteria, fungi) or lost<br />

from the system through mineralisation and leaching. Oxidizing conditions<br />

generally prevail at the surface <strong>of</strong> pit bogs but reducing conditions develop<br />

where there is cover <strong>of</strong> stagnant water or with increasing compaction and depth.<br />

Acidity increases with increasing burial depth, bacterial community changes<br />

and their activities decreases and eventually ceases. This stage corresponds to<br />

early diagenesis and the humic substances are the products <strong>of</strong> peatification.<br />

There are gradual biochemical transformations <strong>of</strong> lignin involving<br />

depolymerisation and defunctionalisation. Gases such as CH4, NH3, N2O and<br />

CO2 are produced (Killops and Killops, 1993).<br />

18


1.2.2 Coalification<br />

This is the process whereby Peat is progressively transformed into lignite or<br />

brown <strong>coal</strong>, sub-bituminous <strong>coal</strong>, bituminous <strong>coal</strong> and anthracite. Coalification<br />

can be subdivided into biochemical and geochemical stages. At the biochemical<br />

stage, there is further loss <strong>of</strong> oxygen containing compounds and condensation <strong>of</strong><br />

<strong>organic</strong> residue progresses to form brown <strong>coal</strong> containing about (50-70) %C and<br />

(5-7) %H.<br />

Temperature and pressure are the main agents during the geochemical stage and<br />

as the carbon content increases there is a corresponding decrease in oxygen. The<br />

hydrogen level remains constant initially but later it begins to decrease with<br />

increasing rate at the highest level <strong>of</strong> maturity. The nitrogen content ranges<br />

between (1-2) % while sulphur levels are variable but generally


As the original plant materials, consisting largely <strong>of</strong> well-ordered polymers<br />

(e.g. cellulose and lignin) were degraded lighter, hydrogen rich compounds were<br />

formed and trapped leaving macromolecular residue depleted in hydrogen and<br />

rearranged as a completely disordered macromolecular material. As <strong>coal</strong>ification<br />

progressed, the aromatic character <strong>of</strong> the <strong>coal</strong> increased with ring fusing and<br />

becoming crosslinked. Thus the degree <strong>of</strong> condensation increases in the order:<br />

lignite, bituminous <strong>coal</strong>, and anthracite. The hydroxyl group in <strong>coal</strong>s is mainly<br />

phenolic or acidic. Hydroxyl oxygen may be present in concentration <strong>of</strong> about 9 %<br />

in brown <strong>coal</strong>s, but it is usually about 8 % with 65 %C and 27 %O.<br />

As rank increases, the hydroxyl content decreases slightly untill, at 80 %C and<br />

12.5 %O, a more rapid decrease is initiated, resulting in a hydroxyl oxygen<br />

concentration <strong>of</strong> less that 1 % at 90 %C content.<br />

Carboxyl group are absent in most <strong>coal</strong>s above the lignite stage <strong>of</strong> development, but<br />

are present in brown <strong>coal</strong>s (lignite). At 83 %C content, all carboxyl group<br />

dissappears.<br />

The methoxyl group is lost early during metamorphism <strong>of</strong> <strong>coal</strong> but carbonyl<br />

group are found at all stages <strong>of</strong> <strong>coal</strong>ification. During the metamorphic process in<br />

<strong>coal</strong>s containing up to 70-80 %C, the methoxyl group are lost first, then the<br />

carboxyl and carbonyl groups decrease rapidly. At 81-89 %C content, the hydroxyl<br />

group decreases rapidly. At greater than 92 %C content, almost all oxygen is in<br />

non-reactive stable forms. Hydroxyl and carbonyl groups may account for 70-90 %<br />

<strong>of</strong> the oxygen in bituminous <strong>coal</strong>s. Nitrogen in <strong>coal</strong> is almost completely in cyclic<br />

structures.<br />

1.4 Petroleum Potential <strong>of</strong> Coal<br />

Petroleum generation from <strong>coal</strong> source rocks and its type are fundamentally<br />

dependent on the availability <strong>of</strong> hydrogen (Hunt, 1996; Petersen and Nyt<strong>of</strong>t,<br />

2006). The importance <strong>of</strong> paraffinicity <strong>of</strong> <strong>coal</strong> for generating and expelling liquid<br />

hydrocarbons has received great attention (Isasken et al., 1998; Killops et al.,<br />

1998).<br />

20


The longer the aliphatic chains the greater the potential for generating and<br />

expelling liquid hydrocarbons during thermal maturation (Hunt, 1996; Isasken et<br />

al., 1998; Petersen and Nyt<strong>of</strong>t, 2006).<br />

Isasken et al. (1998) have shown that the concentration <strong>of</strong> long-chained<br />

aliphatic hydrocarbons in the <strong>coal</strong> matrix is the main factor responsible for<br />

petroleum potential <strong>of</strong> humic <strong>coal</strong>s. The debate on hydrocarbon generation and<br />

migration within <strong>coal</strong> has intensified since the early work by Brooks and Smith<br />

(1967,1969) who proposed that certain oils originated from <strong>coal</strong>s. Today, it is<br />

widely accepted that some <strong>coal</strong>s can indeed generate oil (Betrand, 1986;<br />

Boreham and Powell, 1991; Hunt, 1991,1996; Hutton et al., 1994).<br />

1.5 Coal Reserves in Nigeria<br />

Although <strong>coal</strong> deposits exist in nearly every region <strong>of</strong> the world, but vast<br />

<strong>coal</strong> deposits occur only in Europe, Asia, Australia and North<br />

America.Commercial <strong>coal</strong> deposits occur in sedimentary rock basins, typically<br />

sand-witched as layers (beds or seams) between sandstone and shale. Nigeria has<br />

one <strong>of</strong> the largest <strong>coal</strong> deposits and largest lignite deposits in Africa.<br />

Coal was discovered in Nigeria in 1909 by the Minerals Survey <strong>of</strong> Southern<br />

Nigeria in the South Eastern part <strong>of</strong> the country near Udi. Between 1909 and<br />

1930,further <strong>coal</strong> outcrops were located in the viccinities <strong>of</strong> Enugu and Ezimo in<br />

Enugu state, Orukpa in Benue state and Okaba and Ogboyoga-Odukpani in Kogi<br />

state. Later, discoveries <strong>of</strong> the lignite fields in Delta state and bituminous <strong>coal</strong>s<br />

<strong>of</strong> Lafia-Obi in Nasarawa state were reported (Nigerian <strong>coal</strong> Corporation report,<br />

2005). Majority <strong>of</strong> the <strong>coal</strong> deposits are located between latitudes 6 o 39’ and 8 o<br />

35’N and longitudes 6 o 32’ and 9 o 21’E within the Cretaceous Mamu Formation<br />

in the Anambra basin (Fig. 1.1).<br />

21


The <strong>coal</strong> deposits <strong>of</strong> the Anambra basin located in southeastern Nigeria<br />

appear to contain the largest and most economically viable <strong>coal</strong> resources. This<br />

basin covers an area <strong>of</strong> approximately 1.5million hectares and is constrained by<br />

the Niger River on the west, the Benue River on the north and the Enugu<br />

Escarpment on the east. The <strong>coal</strong> is predominantly in one seam that outcrops<br />

along the eastern side <strong>of</strong> the basin at the base <strong>of</strong> the Enugu escarpment and dips<br />

gently towards the center <strong>of</strong> the basin. Coal outcroppings have also been<br />

reported at Idah and Dekina on the northwestern part <strong>of</strong> the basin, demonstrating<br />

that <strong>coal</strong> exists on the western side <strong>of</strong> the basin as well as the east. Detailed<br />

distribution <strong>of</strong> reserves is shown in Table 1.1.<br />

22


Fig. 1.1 : Location <strong>of</strong> Anambra Basin in Nigeria (boxed area <strong>of</strong> inset) showing the location <strong>of</strong> the <strong>coal</strong>, lignite<br />

and shale deposits. Numbers indicate Cretaceous and Tertiary formations as follows: 1. Asu River Group;<br />

2. Odukpani Formation; 3. Eze-Aku Shale; 4.Awgu Shale; 5. Enugu/Nkporo Shale; 6. Mamu Formation;<br />

7. Ajali Sandstone; 8. Nsukka Formation; 9. Imo Shale; 10. Ameki Formation and 11. Ogwashi-Asaba<br />

Formation (Akande et al., 2007).<br />

23


Table 1.1 : Coal and Lignite Reserves <strong>of</strong> Nigeria.<br />

Serial No Mine Location State Type <strong>of</strong> Coal<br />

24<br />

Estimated<br />

Reserves (Million<br />

tones)<br />

Proven<br />

Reserves<br />

(Million tones)<br />

Borehole<br />

Records<br />

Coal Outcrop and<br />

SeamThickness<br />

(m)<br />

1 Okpara Enugu Sub-bituminous 100 24 20 Many (1.5)<br />

2 Onyeama Enugu Sub-bituminous 150 40 Many Many (1.5)<br />

3 Ihioma Imo Lignite 40 Not Available Nil Many<br />

4 Ogboyoga Kogi Sub-bituminous 427 107 31 17(0.8-2.3)<br />

5 Ogwashi-Azagba/Obomkpa Delta Lignite 250 63 7 4(3.5)<br />

6 Ezimo Enugu Sub-bituminous 156 56 4 10(0.6-2.0)<br />

7 Inyi Enugu Sub-bituminous 50 20 4 0.9-2.0<br />

8 Lafia-Obi Nassarawa Bituminous 156 21.42 123 Nil (1.3)<br />

9 Oba/Nnewi Anambra Lignite 30 Not Available 2 14(0.3-4.5)<br />

10 Afikpo/Okigwe Ebonyi/Imo Sub-bituminous 50 Not Available Nil Not Available<br />

11 Amasiodo Enugu Bituminous 1000 Not Available 3 Not Available<br />

12 Okaba Kogi Sub-bituminous 250 3 Many 0.8-2.3<br />

13 Owukpa Benue Sub-bituminous 75 57 Many 0.8-2.3<br />

14 Ogugu/Awgu Enugu Sub-bituminous Not Available Not Available Nil Not Available<br />

15 Afuji Edo Sub-bituminous Not Available Not Available Nil Not Available<br />

16 Ute Ondo Sub-bituminous Not Available Not Available Nil Not Available<br />

17 Doho Bauchi Sub-bituminous Not Available Not Available Nil Not Available<br />

18 Kumuru-Pindosa Bauchi Sub-bituminous Not Available Not Available Nil Not Available<br />

19 Lamja Adamawa Sub-bituminous Not Available Not Available Nil Not Available<br />

20 Ganin Maigunga Bauchi Sub-bituminous Not Available Not Available Nil Not Available<br />

21 Gindi Akwati Plateau Sub-bituminous Not Available Not Available Nil Not Available<br />

22 Janata Kogi Kwara Sub-bituminous Not Available Not Available Nil Not Available<br />

Source: Nigerian Coal Corporation Report, 2005.


1.6 Petroleum Source Rock Evaluation<br />

A petroleum source rock may be defined as fine-grained sediment that<br />

has generated and released enough hydrocarbons to form an accumulation <strong>of</strong> oil<br />

and gas while potential source rock is one that is not mature to generate<br />

petroleum in its natural setting but will form significant quantities <strong>of</strong> petroleum<br />

when required thermal maturity is attained (Hunt, 1996; Hunt et al., 2002).<br />

Accurate characterization <strong>of</strong> the oil generation potential <strong>of</strong> source rocks is<br />

essential for hydrocarbon accumulation assesment in a petroleum system (Kwan-<br />

Hwa Su et al., 2006). The petroleum potential <strong>of</strong> any source rock is evaluated by<br />

determining the quantity, type and thermal maturity <strong>of</strong> <strong>organic</strong> matter contained<br />

in such rock. These parameters are discussed briefly below.<br />

1.6.1 Quantity <strong>of</strong> Organic matter<br />

The quantity <strong>of</strong> <strong>organic</strong> matter in source rocks is usually expressed as the total<br />

<strong>organic</strong> carbon (TOC). The minimum acceptable TOC values for various types<br />

<strong>of</strong> source rocks are 0.5 % for shales, 0.3 % for carbonates and 1.0 % for clastic-<br />

type rocks (Killops and Killops, 1993). A minimum <strong>of</strong> 1.5-2 % TOC has<br />

generally been accepted for defining good source rocks (Hunt, 1996).<br />

The amount <strong>of</strong> hydrocarbon isolated from the bitumen extracted from finely<br />

ground rock samples can also provide a useful indication <strong>of</strong> whether any oil<br />

potential exists. Oil source rocks are generally considered to require a minimum<br />

hydrocarbon content <strong>of</strong> 200-300 ppm (Killops and Killops, 1993).<br />

The Genetic Potential (GP) expressed in milligram hydrocarbon per gramme <strong>of</strong><br />

rock (mg/HC/g) can also be used to evaluate the maximum quantity <strong>of</strong><br />

hydrocarbon that a particular rock had already generated (S1) and would be<br />

generated (S2) if exposed to a sufficient prolonged thermal stress i.e. (S1 + S2).<br />

Both the S1 and S2 values can be obtained from the Rock-Eval pyrolysis <strong>of</strong> rocks.<br />

25


1.6.2 Quality <strong>of</strong> Organic matter<br />

The quality <strong>of</strong> <strong>organic</strong> matter contained in rocks can be determined by<br />

Optical and Physiochemical methods. These methods are briefly described<br />

below:<br />

1.6.2.1 Optical Methods<br />

These methods are based on maceral recognition techniques. Maceral<br />

examination can be carried out using reflected light microscopy <strong>of</strong> thin sections<br />

<strong>of</strong> the whole rock or <strong>of</strong> isolated <strong>organic</strong> particles. Transmitted light microscopy<br />

can also be used for isolated maceral concentrates. Shape and degree <strong>of</strong><br />

transmittance or reflectance, and also fluorescence under uv-illumination, can be<br />

used to identify broad maceral groups (liptinite, exinite, vitrinite and inertinite).<br />

1.6.2.2 Physico-chemical Methods<br />

Kerogen is the disseminated <strong>organic</strong> matter in sedimentary rock that is<br />

insoluble in non-oxidizing acids, bases and <strong>organic</strong> solvents. Elemental analysis<br />

<strong>of</strong> kerogen concentrate from rock is the most reliable method <strong>of</strong> characterizing<br />

the types or quality <strong>of</strong> <strong>organic</strong> matter. It is based on the major constituents (C, H,<br />

O), which have been used to define main types <strong>of</strong> kerogen based on the plot <strong>of</strong><br />

H/C versus O/C in Van Krevelen diagram. The plot <strong>of</strong> Hydrogen index (HI) vs.<br />

Oxygen index (OI) provides an analogue to the van Krevelen diagram. Both the<br />

HI and OI can be obtained from Rock-Eval pyrolysis. Based on the plot <strong>of</strong> H/C<br />

vs. O/C and HI vs. OI, kerogen can be classified into types I to IV which are<br />

broadly equivalent to the maceral groups, liptinite, exinite, vitrinite and inertinite<br />

respectively for <strong>coal</strong>s (Killops and Killops, 1993).<br />

Type I kerogen has a high H/C ratio and a low O/C ratio. It contains a<br />

significant contribution from lipid material. Algal material and bacteria remains<br />

are the main contributors <strong>of</strong> its <strong>organic</strong> matter. The kerogen is deposited under<br />

anoxic conditions in shallow water environment (e.g. Lagoon and Lake) with<br />

total 4.6% oxygen. The type I kerogen is relatively rare but has high oil potential.<br />

26


The type II kerogen has relatively high H/C and low O/C ratios. It is formed<br />

from mixed authochthonous phytoplankton, zookplankton and microbial (mainly<br />

bacteria) <strong>organic</strong> matter deposited under reducing condition in marine<br />

environment. The type II kerogen has lower yields <strong>of</strong> hydrocarbon upon<br />

pyrolysis compared to Type I kerogen. It is the most common type and has both<br />

oil and gas potential.<br />

The type III kerogen that is derived from vascular plants has low H/C and<br />

high O/C ratio. Type III kerogen is formed from. Vitrinite concentrations<br />

generally occur as <strong>coal</strong> or <strong>coal</strong>y shales, hence type III kerogen is comparable<br />

with <strong>coal</strong>s in terms <strong>of</strong> its composition and behaviour with increasing burial. Type<br />

III kerogen is much less likely to generate oil compared to type II and I but may<br />

be a source <strong>of</strong> gas (chiefly methane).<br />

The type IV kerogen is probably formed from higher plant matter that has<br />

been severely oxidized on land and then transported to its deposition site. It has<br />

low atomic H/C and low O/C ratio. It can be derived from other kerogen types<br />

that have been reworked and oxidized. It is sometimes not considered as true<br />

kerogen because it has no hydrocarbon generating potential.<br />

1.6.2.3 Chemical Methods<br />

Chemical methods based on bitumen extracts for assessing the <strong>organic</strong><br />

matter in source rocks are mainly based on biomarkers. Analysis <strong>of</strong> biomarkers<br />

in source rock extracts gives useful information on sources <strong>of</strong> <strong>organic</strong> material.<br />

Biomarkers allow the recognition <strong>of</strong> main input <strong>of</strong> <strong>organic</strong> matter (Hunt et al.,<br />

2002; Olvella et al., 2006). The application <strong>of</strong> biomarker in kerogen quality<br />

assessment is discussed further in section 2.1.<br />

27


1.6.2.4 Carbon Isotopic Composition<br />

Carbon isotope distributions in whole source rock or extracts give<br />

information on type <strong>of</strong> <strong>organic</strong> material. Kerogen from higher plant sources<br />

usually exhibit lower δ 13 C values than those sourced by marine organism by a<br />

difference <strong>of</strong> about 3-5 % (Killops and Killops, 1993; Revill et al., 1994; Meyers,<br />

2003). This topic is discussed further in section 2.3.<br />

1.6.3 Thermal maturity <strong>of</strong> Organic Matter<br />

Information on the level <strong>of</strong> maturity <strong>of</strong> <strong>organic</strong> matter in the source rock<br />

is needed to determine if the source rock has reached the stage <strong>of</strong> hydrocarbon<br />

generation or not. The maturity status <strong>of</strong> source rock can be determined from<br />

Rock-Eval pyrolysis, petrographic and biomarker analyses. These parameters are<br />

briefly discussed below:<br />

1.6.3.1 Rock-eval Pyrolysis<br />

The major maturity parameters from the Rock-eval pyrolysis are<br />

Production Index (PI) or Transformation Ratio (TR) and Tmax.These parameters<br />

increase with increasing maturation. The PI or TR expressed as the ratio <strong>of</strong><br />

S1 /S1+S2 measures the extent to which the genetic potential <strong>of</strong> the rock has been<br />

effectively utilized. It is expressed as the ratio <strong>of</strong> the hydrocarbon already<br />

generated to the total genetic potential. Generally, the threshold <strong>of</strong> the oil zone is<br />

fixed at about 0.1. The ratio reaches about 0.4 at the bottom <strong>of</strong> the oil-window<br />

and increases to 1.0 when the hydrocarbon generative capacity <strong>of</strong> the kerogen<br />

has been exhausted.<br />

The Tmax indicates the temperature <strong>of</strong> the maximum generation <strong>of</strong> S2<br />

peak. Generally, the threshold <strong>of</strong> the oil zone is fixed at Tmax <strong>of</strong> 430 o C - 435 o C<br />

for type II and III kerogen while gas zone ranges from 450 o C - 455 o C and<br />

465 o C - 470 o C for type II and type III respectively (Killops and Killops, 1993,<br />

2005).<br />

28


Organic petrographers also developed series <strong>of</strong> maturation indicators, the most<br />

reliable <strong>of</strong> which is vitrinite reflectance (Hunt et al., 2002).<br />

The use <strong>of</strong> vitrinite reflectance as a technique for determining the maturity<br />

<strong>of</strong> oil in sedimentary rocks was first described by Teichmuller (1958).<br />

Today, vitrinite reflectance is a widely used indicator <strong>of</strong> thermal stress because it<br />

extends over a longer maturity range than any other indicator (Hunt et al., 2002).<br />

Vitrinite reflectance can be used to asses thermal maturity in types II and III but<br />

cannot be used for type I kerogen due to absence <strong>of</strong> vitrinite. Vitrinite reflectance<br />

values for main phase <strong>of</strong> oil generation ranges from (0.65-1.3) %R0 and values<br />

greater than 2.0 %R0 indicate dry gas generation (Tissot and Welte, 1984).<br />

Macerals in rocks observed under microscope are affected by increasing<br />

maturity. There is change in colour <strong>of</strong> spores and pollens observed under<br />

transmitted light with increasing temperature. Colour changes from yellow in<br />

immature samples through orange/yellow-brown during diagenesis and brown<br />

during catagenesis to black in metagenesis.<br />

1.6.3.2 Biomarker analysis<br />

Maturity <strong>of</strong> source rocks can also be determined from the biomarker<br />

distribution in source rock extracts. Many maturity parameters have been<br />

developed from the distribution <strong>of</strong> terpanes, steranes and polyaromatic<br />

hydrocarbons in source rocks.<br />

29


1.7 Aims and Objectives <strong>of</strong> this Study<br />

The debate on hydrocarbon generation and migration within <strong>coal</strong> has<br />

intensified since the early work by Brooks and Smith (1967, 1969). Many<br />

studies suggest the possibility <strong>of</strong> non-marine <strong>coal</strong> sediments having potential<br />

capacity to generate petroleum by thermal transformation <strong>of</strong> <strong>organic</strong> matter<br />

compounds during natural diagenesis when buried in sedimentary basins<br />

(Betrand et al., 1986; Boreham and Powell, 1991; Hunt 1991, 1996; Hutton et<br />

al., 1994). Today, it is widely accepted that some <strong>coal</strong>s can indeed generate oil<br />

(Hendrix et al., 1995; Mario et al., 1997; Petersen et al., 2000; Bechtel et al.,<br />

2001; Kotarba et al., 2002; Boreham et al., 2003).<br />

Nigeria is one <strong>of</strong> the countries with the largest deposits <strong>of</strong> <strong>coal</strong> in Africa,<br />

however despite its vast deposits, there is a dearth <strong>of</strong> information on its use as<br />

a source-rock <strong>of</strong> petroleum. The present study therefore aimed at:<br />

(a) evaluating the source rock potential <strong>of</strong> some Nigerian <strong>coal</strong>,<br />

(b) assessing the <strong>organic</strong> source, depositional environment and maturity<br />

<strong>of</strong> the <strong>coal</strong> based on the distribution <strong>of</strong> biomarkers and polycyclic<br />

aromatic hydrocarbons (PAH), and<br />

(c) determining the carbon isotopic composition <strong>of</strong> individual n-<br />

alkanes in the <strong>coal</strong> and its implication on <strong>organic</strong> source and<br />

depositional environment.<br />

30


2.1 Biomarker Geochemistry<br />

CHAPTER TWO<br />

LITERATURE REVIEW<br />

Biomarker or geochemical fossils are <strong>organic</strong> compounds found in geosphere<br />

whose structure can be unambiguously linked to their biological origin, despite the<br />

possibility <strong>of</strong> some structural alteration due to diagenetic or other processes. Treibs<br />

(1934) was the first to develop the biomarker concept with his pioneering work on<br />

identification <strong>of</strong> porphyrins in crude oils and suggested that these porphyrins may<br />

have originated from the chlorophyll <strong>of</strong> plants. All biomarker molecules have<br />

definitive chemical structures, which can be related directly or indirectly through a set<br />

<strong>of</strong> diagenetic alterations to biogenic precursors, and cannot be synthesized by<br />

abiogenic processes (Simoneit, 2002). The use <strong>of</strong> biomarkers as indicators <strong>of</strong> biogenic,<br />

paleoenvironmental, and geochemical processes on Earth has been widely accepted<br />

(Mackenzie et al., 1982; Johns, 1986; Simoneit et al., 1986; Brassel, 1992; Imbus and<br />

Mckirdy, 1993; Mitterer, 1993; Simoneit, 1998). Biomarkers are widely used in<br />

petroleum geochemical studies in source rock evaluation, oil-oil or oil-source rock<br />

correlations, basin evaluation and reservoir management (Peters et al., 2005). Some <strong>of</strong><br />

the biomarkers widely used in geochemical studies are briefly described below.<br />

2.1.1 Normal and branched Alkanes<br />

n-Alkanes are widely distributed in various plants and other organisms and are<br />

probably the most exploited class <strong>of</strong> biomarkers (Philp, 1985).High proportions <strong>of</strong><br />

long chain C27-C31 members relative to the total n-alkanes especially, n-C27 and n-C29<br />

are typical <strong>of</strong> terrestrial higher plants, where they occur as main components <strong>of</strong> plant<br />

waxes i.e.leaf curticles, spores, pollens and resins (Bray and Evans 1961; Meinschein,<br />

1961; Kvenvolden, 1962; Eglinton and Hamilton, 1963; Caldicott and Eglinton, 1973;<br />

Miranda et al., 1999; Tissot and Welte, 1984; Barthlot et al., 1998).<br />

31


The short chain n-alkanes with odd-to-even predominance in the medium molecular<br />

weight region (C11-C17) maximizing at n-C16 are predominantly found in algae and<br />

microorganisms (Clark and Blummer, 1967; Fowler et al., 1986; Miranda et al., 1999;<br />

Ficken et al., 2000).<br />

The ratio <strong>of</strong> odd/even carbon numbered n-alkanes has been in use over a long<br />

period in estimating the thermal maturity <strong>of</strong> fossil fuels (Bray and Evans; 1961;<br />

Philippi, 1965; Scalan and Smith, 1970; Tissot et al., 1977). These ratios can be<br />

expressed as Carbon preference index, CPI (Bray and Evans, 1961) and improved<br />

Odd-to-even predominance, OEP (Scalan and Smith, 1970). The CPI and OEP values<br />

above or below 1.0 indicate low thermal maturity. Values <strong>of</strong> 1.0 suggest, but do not<br />

prove, that an oil or rock extract is thermally mature. The CPI or OEP values below<br />

1.0 are unusual and typify low maturity oils or bitumen from carbonate or hypersaline<br />

environments (Peters et al., 2005). Organic matter inputs do affect these ratios and<br />

therefore are mostly applied with caution.<br />

2.1.2 Acyclic isoprenoids<br />

Acyclic isoprenoids have been widely used in fossil fuel studies for<br />

characterization and correlation studies and to obtain information on depositional<br />

environment especially in lacustrine source rocks. The acyclic isoprenoids<br />

hydrocarbons; pristane, (Pr)(structure I) and phytane, (Ph)(structure II) are ubiquitous<br />

in sedimentary rocks, crude oils and <strong>coal</strong>s (Bechtel et al., 2003). The most abundant<br />

source <strong>of</strong> pristane and phytane is the phytol side chain <strong>of</strong> chlorophyll a and b (Powell<br />

and McKirdy, 1973). In the presence <strong>of</strong> Oxygen, phytol would be oxidized to phytenic<br />

acid, yielding pristane after decarboxylation. In the absence <strong>of</strong> Oxygen, phytol would<br />

be dehydrated, yielding phytadienes, which eventually hydrogenated to phytane. The<br />

Pr/Ph ratios have been proposed as an indicator for the oxicity <strong>of</strong> the depositional<br />

environment (Powell and McKirdy, 1973; Didyk et al., 1978). Pristane/phytane ratio<br />


However, the ratio is known to be affected by differences in the precursors <strong>of</strong> acyclic<br />

isoprenoids (Volkman and Maxwell, 1986; Ten Haven et al., 1987).<br />

A common precursor for pristane and phytane is inferred by the similarity in their<br />

δ 13 C values (Hayes et al., 1990).<br />

High Pr/Ph (>3.0) indicates terrigenous <strong>organic</strong> matter input under oxic<br />

conditions, while low values (


Abundant lycopanes have been reported in hypersaline euxinic settings (Grice et al.,<br />

1998) and in mesosaline carbonates, such as the Jurassic Malm stage carbonates<br />

(Schwark et al., 1998). Lycopane has been proposed to be a bacterial marker derived<br />

from reduction <strong>of</strong> lycopene (Killops and Killops, 2005).<br />

Squalene (structure V) serves as precursor to polycyclic terpenoids, steroids and<br />

carotenoids. Squalene is a major lipid produced by methanogenic, thermophillic and<br />

thermoacidophilic archaea. Abundant squalane, a saturated C30 isoprenoidal alkane,<br />

may represent a direct archaebacterial input (Matsumoto and Watanuki, 1990) or<br />

derive from diagenetic reduction <strong>of</strong> squalene, which occur in a variety <strong>of</strong> organisms.<br />

Squalane has been used as a biomarker for archaea and hypersaline depositional<br />

environments (Ten Haven and Rulkotter, 1988).<br />

2.1.3 n-Alkanol, Alkanoic acids and Fatty acids<br />

Aliphatic wax lipids (n-alkanols, n-alkanoic acids, fatty acids) in the range <strong>of</strong><br />

C22-C32 were identified as the major extractable components <strong>of</strong> angiosperm leaves,<br />

barks and roots (Huang et al., 1995; Lockheart et al., 2000; Kogel-Knaber 2002;<br />

Yuang and Huang, 2003). Similar distributions have been reported in sediments and<br />

macr<strong>of</strong>ossils (Logan and Eglinton, 1994; Huang et al., 1995, 1996; Lockheart et al.,<br />

2000). The functionalized lipids are partly degraded during diagenesis (Bechtel et al.,<br />

2001). The wax esters, their hydrolysis products, and the free alkanol and fatty acids<br />

are preferentially degraded to aldehyde, ketones and alkanes by microorganisms<br />

(Puttmann and Bracke, 1995). These degradation pathways favor short chain (


Generally, short chain n-fatty acids (C25) homologs with a high odd carbon<br />

number predominance were interpreted to originate from vegetation wax by oxidation<br />

while the lower homologues (


These isoprenoid ketones could be produced from free phytol by bacteria degradation<br />

and photosynthesized oxidation, photosynthesized oxidation <strong>of</strong> some isoprenoids<br />

hydrocarbons and during photo degradation <strong>of</strong> α-chlorophyll (Tuo and Li, 2005).<br />

The homologous series <strong>of</strong> alkan-2-ones is generally found with an odd carbon number<br />

predominance and its source is likely microbial (Leif and Simoneit, 1995; Bai et al.,<br />

2006). It has been proposed that n-alkan-2-ones are formed by microbially mediated<br />

β-oxidation <strong>of</strong> alkanes (Simoneit et al., 1979; Cranwell et al., 1987;<br />

Rieley et al., 1991) or from β-oxidation and decarboxylation <strong>of</strong> n-fatty acids<br />

(Volkman et al., 1983). n-Alkan-2-ones maximizing at C25 or C27 have been reported<br />

in higher plants,microalgae and phytoplankton(Gonzalez et al., 2003; Bai et al., 2006).<br />

2.1.5 Sesquiterpenoids<br />

Sesquiterpanes primarily those based on the cadinane skeleton, with their aromatic<br />

derivatives, are found in many crude oils from inferred deltaic source rock<br />

(Van Aarssen et al., 1990, 1992). Cadinanes (structure VI) and bisabolanes belong to<br />

the most common sesquiterpenoids in the plant kingdom and are thus non-specific<br />

(Otto and Wilde, 2001). Derivatives <strong>of</strong> longifolane, santalane, and himachalane were<br />

reported from conifer species <strong>of</strong> the Pinaceae, Cupressaceae, and Podocarpaceae<br />

(Otto and Wilde, 2001). Also, aromatic sesquiterpenoids, dihydro-ar-curcumene,<br />

cuparene, calamene, cadina-1(10),6,8-triene and cadalene have been reported from<br />

various origin including, higher plants, conifers, dammar resin, Dipterocarpacceae<br />

and Cupressaceae sensu lato (Haberer et al., 2006),thereby supporting its non-<br />

specificity. However, there are some specific sesquiterpenes like cuparene, which are<br />

characteristic biomarkers for the conifer family Cupressaceae sensu lato including the<br />

genera, Cupressus, Thuja and Juniperus (Grantham and Douglas, 1980; Otto and<br />

Wilde, 2001) or sativene, a specific marker for Pinaceae (Otto and Wilde, 2001).<br />

Sesquiterpenoids are valuable chemosystematic markers for extant conifers.<br />

Examples include sesquiterpenoid classes <strong>of</strong> the cedrane, cuparane, valencane and<br />

thujospane.<br />

36


Their derivatives were reported only from modern species <strong>of</strong> the Cupresaceae and<br />

Taxodiaceae (Stefanova et al., 2002). The biological precursors <strong>of</strong> cadalene type<br />

sesquiterpenoids are common constituents <strong>of</strong> the resin <strong>of</strong> conifers (Simoneit et al.,<br />

1986; Otto et al., 1997).<br />

Cadalene precursors are resistant to diagenesis and thermal maturation and their<br />

relative abundances in sedimentary <strong>organic</strong> matter have been used as proxies <strong>of</strong> higher<br />

plant input in previous reconstructions <strong>of</strong> palaeo-vegetation and palaeo-climate (van<br />

Aarssen et al., 2000).<br />

37


Fig. 2.1: Chemical structures <strong>of</strong> some compounds cited in the literature<br />

(Peters et al., 2005).<br />

38


Fig. 2.1 (contd.): Chemical structures <strong>of</strong> some compounds cited in the literature<br />

(Peters et al., 2005).<br />

39


2.1.6. Diterpenoids<br />

Diterpenoids form a large group <strong>of</strong> natural products, which are widely<br />

distributed in the plant and animal kingdom. Diterpenoids originate mainly from<br />

gymnosperms and can provide useful chemotaxonomic information on fossil and<br />

extant Coniferales families (Thomas, 1970; Simoneit, 1977; Wakeham et al., 1980;<br />

Simoneit et al., 1986; Otto et al., 1997; Otto and Simoneit, 2001; Otto and Wilde,<br />

2001; Otto et al., 2003).<br />

Large numbers <strong>of</strong> diterpenoids from various classes with the main groups being<br />

abietanes, pimaranes, kauranes and phyllocladanes have been identified in recent<br />

plant species (Sukh, 1989; Otto et al., 1997). Possible precursors <strong>of</strong> most <strong>of</strong> the<br />

common diterpenoid skeleton-types can be found in species <strong>of</strong> nearly all conifer<br />

family (Tuo and Philp, 2005). They can also be preserved in fossil plants and were<br />

previously reported from several Tertiary and Cretaceous conifers (Otto et al., 1997,<br />

2001, 2002; Otto and Simoneit, 2001), and are found in high concentration in the seed<br />

cone <strong>of</strong> conifer species (Otto et al., 2003, 2005).<br />

Diterpenoids containing the abietane, pimarane and kaurane skeletons have<br />

been found in fossil resins, <strong>coal</strong>s, crude oils, recent and older sediments (Otto et al.,<br />

1997; Miranda et al., 1999; Bastow et al., 2001). Simonellite and its associated<br />

aromatic abietanes are <strong>of</strong>ten referred to as biomarkers for Pinaceae, because abietic<br />

acid only occurs in Pinaceae species (Simoneit, 1977; Simoneit et al., 1986).<br />

Phenolic abietanes (ferruginol and derivatives) are characteristics for the conifer<br />

families Cupressaceae senso lato, Podocarpaceae and Araucariaceae, and can be<br />

used as their chemosystematic markers (Thomas, 1970; Otto and Wilde, 2001;Otto et<br />

al., 2005). Tricyclic diterpenoids such as 18-norpimarane, pimarane and<br />

dehydroabietane, simonellite and retene are thought to be derived from resins formed<br />

by higher plants primarily gymnosperms (conifer), but also by some angiosperms;<br />

pteridophytes, and bryophytes (Otto et al., 1997). For instance, simonellite and retene<br />

may originate from other abietane-type precursors; such as taxodone and ferruginol<br />

present in recent Taxodum species (Otto et al., 1997).<br />

40


Sponges are the only marine organisms known thus far to contain tricyclic<br />

diterpenoids (Wen et al., 2000).<br />

The presence <strong>of</strong> tetracyclic diterpenoids, 16α(H)-phyllocladane and ent-16β(H)-<br />

kaurane are usually considered to be indicative <strong>of</strong> Podocarpaceae, Cupressaceae and<br />

Araucariaceae conifer contributions (Thomas, 1970, 1988; Sukh, 1989; Otto et al.,<br />

1997), Taxodiaceae (Dehmer, 1995; Otto et al., 1997, Bechtel et al., 2002) and<br />

probably from pteridophytes (Cheng et al., 1997). Two isomeric phyllocladanes, β-<br />

and α- phyllocladane, have been identified in many sediments and <strong>coal</strong>s (Simoneit,<br />

1986). They have been reported as biomarkers for the Pordocarpaceae and<br />

Araucariaceae, because high amounts <strong>of</strong> phyllocladane precursors have been detected<br />

in recent species <strong>of</strong> these conifer families (Schulze and Michaelis, 1990).<br />

The saturated and aromatized abietanes, pimaranes and phyllocladanes, commonly<br />

identified in sediments and <strong>coal</strong>s, can be related only to the whole Coniferales group.<br />

This is because functionalized precursors <strong>of</strong> these diterpenoids occur in recent species<br />

<strong>of</strong> several conifer families (Otto et al., 1997).<br />

2.1.7 Triterpenoids<br />

All triterpenoids appear to be derived from the acyclic isoprenoid squalene,<br />

which is a ubiquitous component in organisms (Killops and Killops, 2005). Members<br />

<strong>of</strong> triterpenoids include tricyclic-, tetracyclic and pentacyclic. Many terpanes originate<br />

from prokaryotic bacteria membrane lipid (Ourisson et al., 1982). Some <strong>of</strong> the<br />

triterpenoids <strong>of</strong>ten used are discussed briefly below.<br />

2.1.7.1 Tricyclic and tetracyclic triterpanes<br />

Tricyclic terpanes are common in crude oils and sediments and were first<br />

observed in extract from the Green River Formation (Gallegos, 1971). The most<br />

prominent tricyclic terpanes are important components in the saturated hydrocarbon<br />

fractions <strong>of</strong> petroleum (Moldowan et al., 1983). C30 tricyclohexaprenol in bacterial<br />

membranes or malabaricatrienes from algae or bacteria could be intermediates in the<br />

biosynthetic pathway that account for tricyclic terpanes in petroleum (Peters, 2000).<br />

41


Higher homologs may originate from C40 tricyclooctaprenol (Azevedo et al., 1998) or<br />

larger precursors.<br />

High concentration <strong>of</strong> tricyclic terpanes and their aromatic analogs commonly<br />

correlate with high paleolatitude Tasmanite-rich rocks suggesting an origin from these<br />

algal (Aquino Neto et al., 1983; Azevedo et al., 1992). A higher C19-C21 tricyclic<br />

terpanes relative to C23 tricycloterpane can be interpreted as terrestrial <strong>organic</strong> matter<br />

(Ozcelik and Altunsoy, 2005). Tricyclic terpanes can be used to correlate crude oils<br />

and source rock extracts, predict source rock characteristics, and to evaluate the extent<br />

<strong>of</strong> thermal maturity and biodegradation (Peters and Moldowan, 1993). C22/C21 and<br />

C24/C23 tricyclic terpane ratios help to identify extracts and crude oils derived from<br />

carbonate source rocks while C26/C25 tricyclic terpane ratio is useful as a supporting<br />

method to distinguish lacustrine from marine oils (Peters et al., 2005).<br />

Tetracyclic terpanes generally in the range C24-C27 are also frequent constituents <strong>of</strong><br />

oils and bitumens. There is tentative evidence for homologs up to C35<br />

(Killops and Killops, 2005). They are structurally related to hopanes, for which they<br />

may be formed by thermal or bacterial cleavage <strong>of</strong> the 17(21) C-C bond in the ring,<br />

although a direct bacterial source cannot be discounted (Killops and Killops, 2005;<br />

Peters et al., 2005).<br />

C24-C27 tetracyclic terpanes are referred to as de-E-hopanes, or 17,21-secohopanes<br />

and are more resistant to biodegradation and maturation than hopanes. Therefore they<br />

are used in the correlation <strong>of</strong> biodegraded oils (Peters and Moldowan, 1993).<br />

Abundant C24 tetracyclic terpane in fossil fuel indicate carbonate and evaporite source<br />

rock settings (Connan et al., 1986) and are also common in most marine oils<br />

generated from mudstones to carbonate source rocks. C25-C27 tetracyclic terpanes<br />

have also been reported in carbonate and evaporite samples (Connan et al., 1986).<br />

2.1.7.2 Pentacyclic triterpanes (Non-Hopanoid triterpenoids)<br />

Today the biological origin <strong>of</strong> most pentacyclic triterpenoids is known (Borrego<br />

et al., 1997). Angiosperms contain triterpenoids <strong>of</strong> β-amyrin, which on diagenesis<br />

ultimately produced the C30 triterpane oleanane.<br />

42


The use <strong>of</strong> α-and β- amyrins as terrestrial plant markers relies on the generally<br />

accepted view that these triterpenoids are only synthesized in higher plant cells.<br />

Although, possibility <strong>of</strong> contaminations has been suggested (Volkman, 2003).<br />

Pentacyclic terpanes gives information on the <strong>organic</strong> matter type, maturation, and the<br />

lithology <strong>of</strong> source rocks (Waples and Machiara, 1990). Pentacyclic triterpenoids<br />

based on the oleanane, lupane and related skeletons have provided some <strong>of</strong> the most<br />

useful markers inputs <strong>of</strong> <strong>organic</strong> matter from terrestrial plants to marine sediments<br />

(Killops and Frewin, 1994; Volkman et al., 2000; Volkman, 2005). Oleanane<br />

frequently encountered in sediments younger than Late Cretaceous, is therefore useful<br />

as indicator <strong>of</strong> geological age (Nyt<strong>of</strong>t et al., 2002; Ozcelik and Altunsoy, 2005).<br />

Oleanane occurs in two isomers; 18α(H)-oleanane (structure VII) and 18β(H)-<br />

oleanane (structure VIII), the relative amount <strong>of</strong> which changes with the level <strong>of</strong><br />

thermal maturation, and thus can be used as indicators <strong>of</strong> maturity (Ekweozor and<br />

Telnǽs, 1990). Oleanane index (oleanane/C30 hopane) can also be used to indicate the<br />

terrestrial and marine input into the <strong>organic</strong> matter (Waples and Machiara, 1990).<br />

Gammacerane (structure IX) has been found in sediments from the Late<br />

Proterozoic, which contains some <strong>of</strong> the earliest examples <strong>of</strong> fossil protozoan<br />

(Summons et al., 1988) and bacteria in marine sediments (Kleemann et al., 1990). Its<br />

presumed precursor is tetrahymanol first isolated from the ciliate protozoan<br />

Tetrahymena pyriformis (Mallorry et al., 1963) and since then, it has been found in<br />

other eukaryotes i.e. other ciliates, ferns and fungi (Volkman, 2005). Diagenetic<br />

conversion <strong>of</strong> tetrahymanol to gammacerane most likely proceed by dehydration to<br />

form gammacer-2-ene, followed by hydrogenation or may also arise by sulphurization<br />

and subsequent cleavage <strong>of</strong> tetrahymanol (Sinninghe Damste et al., 1995). High<br />

gammacerane is <strong>of</strong>ten seen in fresh water lacustrine sediments (Peters and Moldowan,<br />

1993; Peters et al., 2005) and in certain marine crude oils from carbonate-evaporite<br />

source rocks (Peters et al., 2005). It was proposed that the gammacerane is in fact an<br />

indicator for water stratification during source deposition (Sinninghe-Damste et al.,<br />

1995).<br />

43


Lupanes (structure X) are probably derived mainly from angiosperms, and the<br />

occurrence <strong>of</strong> these compounds should thus be expected to follow that <strong>of</strong> angiosperm<br />

group (Nyt<strong>of</strong>t et al., 2002). Various higher plant lupanoids e.g. lupane-3β, 20,28-triol,<br />

lup-20 (29)-en-3β, 28-diol (betulin), lup-20 (29)-en-3β-ol (lupeol), and 3β-<br />

hydroxylup-20(29)-en-28-oic acid (betulinic acid), may be possible biological<br />

precursors. Lupane occurs frequently in <strong>coal</strong>s and lignites (Wang and Simoneit, 1990;<br />

Peters and Moldowan, 1993; Stefanova et al., 1995) and has been identified in<br />

evaporitic sediments (Poinsot et al., 1995). Lupane is reported rarely in crude oils, and<br />

this may be attributed to difficulties in its identification (Nyt<strong>of</strong>t et al., 2002). Lupane<br />

has similar mass spectrum and also co-elute with oleananes. Therefore lupane may be<br />

resolved partially from oleanane using gas chromatography capillary columns with<br />

polar stationary phases or via High Pressure Liquid Chromatography (Nyt<strong>of</strong>t et al.,<br />

2002) and best quantified using GC-MS/MS methods scanning the m/z 412.4 to 369.3<br />

transition (Peters et al., 2005).<br />

2.1.7.3 Hopanoids<br />

Hopanoids are considered biomarkers for bacteria and cyanobacteria. Most<br />

hopanes molecular fossils originate from polar constituents <strong>of</strong> prokaryotic organisms<br />

(Nyt<strong>of</strong>t et al., 2006). The most probable biological precursor <strong>of</strong> the hopane derivatives<br />

is bacteriohopanepolyol, which are present in the cell membranes <strong>of</strong> prokaryotic<br />

organisms (bacteria and blue algae) where they played the rigidifying role by steroids<br />

in eukaryotic organisms (Ourisson et al., 1982; Rohmer et al., 1992; Durand, 2003;<br />

Bechtel et al., 2007a&b). The C30 Hopanoids have also been found in some<br />

cryptogams; moss, fern (Bechtel et al., 2007a&b). While hopanes with 30 or few<br />

carbon atoms are <strong>of</strong>ten interpreted as diagenetic products <strong>of</strong> C30 hopanoids (e.g.<br />

diploptene and diplopterol), the extended hopanes have been related to C35 precursors,<br />

such as bacteriohopane polyols, aminopolyols and a number <strong>of</strong> composite hopanoids<br />

(Wang et al., 1996).<br />

44


High abundance <strong>of</strong> C35 hopane usually found in marine carbonates and evaporitic<br />

sediments has been attributed to highly reducing depositional environments<br />

(Yangming et al., 2005).<br />

V-shape pattern <strong>of</strong> homohopane, i.e. C31>C32≥C33≤C34


The predominance <strong>of</strong> C27 steranes in non-marine strata indicates a deep lake facies<br />

and source input <strong>of</strong> plankton (algae) while C29 sterane predominance shows a swamp<br />

shallow water environment and a terrestrial higher plant input (Volkman, 1986; Peters<br />

and Moldowan, 1993; Volkman et al., 1999; Otto et al., 2005; Jauro et al., 2007).<br />

C28 sterane is a unique biomarker signature <strong>of</strong> <strong>organic</strong> matter deposited in saline<br />

lacustrine facies. Sterane/hopane ratio is <strong>of</strong>ten used as a measure <strong>of</strong> the relative inputs<br />

<strong>of</strong> eukaryotic versus prokaryotic debris (Peters and Moldowan, 1993). Low<br />

sterane/triterpane ratio (Norgate et al., 1999) has been postulated to favour terrestrial<br />

paralic facies rather than peat swamp facies as <strong>organic</strong> matter source. Sterane<br />

isomerisation at C-20 have been found useful in assessing the level <strong>of</strong> thermal<br />

maturity <strong>of</strong> oil and sediments (Fig. 2.2). The 20S/20S+20R ratio (usually measured<br />

using the C29ααα steranes) is one <strong>of</strong> the most widely applied molecular maturity<br />

parameters in petroleum <strong>geochemistry</strong> (Farrimond et al., 1998). It is based on the<br />

relative enrichment <strong>of</strong> the 20S isomer compared with the biologically inherited 20R<br />

stereochemistry with increasing maturity.<br />

46


Fig. 2.2: Transformation <strong>of</strong> steranes at C-20 during thermal maturity<br />

47


Sterane nuclear isomerisation ratio, (αββ/ αββ+ααα) is widely applied owing to its<br />

operation beyond oil window (Farrimond et al., 1998). In relatively immature samples,<br />

coelution <strong>of</strong> the ααα isomer with the αββ doublet is a common problem that is<br />

responsible for the relatively high αββ values at shallow depth and apparent decreases<br />

in the parameter with increasing depth (Farrimond et al., 1998). In highly mature<br />

source rocks, an eventual decrease in this ratio occurs (Peters et al., 1990).<br />

2.2 Non-Biomarker Compounds<br />

2.2.1 Polynuclear aromatic hydrocarbons (PAHs)<br />

Polycyclic aromatic hydrocarbons are a suite <strong>of</strong> <strong>organic</strong> compounds that are<br />

ubiquitous in sediments. The possible source may be oil, <strong>coal</strong>, air-transported particles<br />

combustion <strong>of</strong> fossil fuel, plant material, algae and bacteria or diagenesis (Pereira et<br />

al., 1999).<br />

The widespread occurrence <strong>of</strong> aromatic hydrocarbons in sediments whose carbon<br />

skeletons are non-isoprenoidal (Polycyclic aromatic hydrocarbons) has led to<br />

suggestion that such compounds are the products <strong>of</strong> sedimentary reactions (Radke et<br />

al., 1982a; Smith et al., 1994). It is now widely accepted that assemblages <strong>of</strong><br />

polycyclic aromatic hydrocarbon (PAH) and their alkyl homologs found in recent<br />

sediments are partly derived from non-aromatic biogenic precursors (Wakeham et al.,<br />

1980; Radke et al., 1982a). These transformations occured in the subsurface, either<br />

through microbial processes in the initial stages <strong>of</strong> diagenesis or during subsequent<br />

burial in which such precursors experience the effects <strong>of</strong> temperature, pressure and<br />

catalytic action <strong>of</strong> mineral matrix (Albrecht and Ourisson, 1971; Radke, 1987; Ellis et<br />

al., 1997).<br />

Alkylated polycyclic aromatic hydrocarbons (PAHs) have received attention as<br />

indicators <strong>of</strong> thermal maturity (Budzinski et al., 1995). Throughout the oil window,<br />

aromatic maturity indicators are thought to be more sensitive to maturity effects than<br />

many biomarker maturity parameters (Alexander et al., 1986).<br />

48


A variety <strong>of</strong> aromatic maturity parameters have been poposed on the basis <strong>of</strong><br />

ratios <strong>of</strong> the relative concentration <strong>of</strong> more thermally stable isomers to less stable ones<br />

(Radke and Welte, 1983; Bastow et al., 2000). The compounds substituted in α-<br />

positions are less stable than related isomers with β-substitution patterns.<br />

Naphthalenes and its alkylated derivatives are common constituents <strong>of</strong> fossil fuels<br />

(Bastow et al., 1998, 2000;van Aarssen et al., 1999).<br />

Sesqui- and triterpenoids derived from microbial and land plant sources have been<br />

suggested as major precursors for methylated naphthalenes (Puttmann and Villar,<br />

1987; Strachan et al., 1988).<br />

The distributions <strong>of</strong> methylated naphthalenes are highly variable between samples as<br />

they are controlled by the effects <strong>of</strong> source, thermal stress and biodegradation (van<br />

Aarssen et al., 1999). Several maturity ratios involving methylated naphthalenes have<br />

been developed over the years and used to asses maturities <strong>of</strong> crude oils (Alexander et<br />

al., 1985; Radke et al., 1990,1994; Bastow et al., 1998).<br />

Phenanthrene maturity parameters are based on the greater stability <strong>of</strong> 3-methyl<br />

phenanthrene and 2-methylphenanthrene compared to 9-methylphenanthrene and 1-<br />

methylphenanthrene (Radke and Welte, 1983). The methyl phenanthrene index (MPI)<br />

proposed by Radke is a widely used molecular maturity parameter and depends on the<br />

relative stability <strong>of</strong> the isomers (Radke and Welte, 1983). 1,7-DMP can be suggested<br />

as a marker <strong>of</strong> terrestrial input <strong>of</strong> <strong>organic</strong> matter (Budzinski et al., 1995).<br />

2.2.2 Aromatic Sulphur Compounds<br />

Dibenzothiophenes, benzodibenzothiophenes and their alkylated derivatives are<br />

common constituents <strong>of</strong> aromatic fractions <strong>of</strong> fossil fuels. They are <strong>of</strong> geochemical<br />

importance in the determination <strong>of</strong> type <strong>of</strong> <strong>organic</strong> matter, depositional environment<br />

and thermal maturity <strong>of</strong> geological samples. Sulphurisation <strong>of</strong> <strong>organic</strong> matter is a<br />

process that occurs either at an intermolecular level, forming low molecular weight<br />

<strong>organic</strong> sulphur compounds, like, thiophenes, thilane or thiane moieties or the sulphur<br />

atoms form (poly)-sulphide bridges between individual compounds resulting in<br />

49


creation <strong>of</strong> macromolecular network (Schouten et al., 1995a; Kolonic et al., 2002).<br />

It is now widely accepted that incorporation <strong>of</strong> reduced in<strong>organic</strong> sulphur species into<br />

functionalised lipids at early stages <strong>of</strong> diagenesis give rise to organosulphur<br />

compounds which are readily preserved in sediments (Sinninghe Damste et al., 1989a;<br />

de Graaf et al., 1992).<br />

Low molecular weight organosulphur compounds and sulphur rich<br />

macromolecules bear an unambiguous link with natural precursors and hence source<br />

organisms and in turn provide valuable information on palaeoenvironments <strong>of</strong><br />

deposition (Grice et al., 1998). With increasing diagenesis, an increasing number <strong>of</strong><br />

C-S bonds are cleaved and compounds enter fractions <strong>of</strong> lower molecular weight (van<br />

kaam-Peters et al., 1998). This implies that the amount <strong>of</strong> lower molecular weight<br />

organosulphur compounds generated is directly linked to the number <strong>of</strong> sulphur<br />

crosslinks by which compounds are bound in macromolecular fractions (van kaam-<br />

Peters et al., 1998). Sulphur enrichment in <strong>coal</strong>s suggests high activity <strong>of</strong> anaerobic<br />

bacteria in a slightly acidic to neutral, sulphate bearing and methanogenic<br />

environment (Grice et al., 1998; Sachsenh<strong>of</strong>er et al., 2000a; Bechtel et al., 2001).<br />

High sulphur rich oils are indications <strong>of</strong> carbonate evaporite source rocks, while low<br />

sulphur oils are more typical <strong>of</strong> siliciclastic source rock (Younes and Philp, 2005).<br />

Sykes et al. (2004) adopted 0.5% sulphur (dry-ash-free) as the highest concentration<br />

for entirely non-marine <strong>coal</strong>.<br />

The low content <strong>of</strong> organosulphur compounds in the oils, as reflected in the low<br />

dibenzothiophene/phenanthrene ratio,


more stable β-substituted isomer (4-MDBT).<br />

A number <strong>of</strong> ratios are applicable for thermal maturity assessments on the basis <strong>of</strong><br />

aromatic sulphur compounds e.g.Lorgarithmic-scale cross plots <strong>of</strong> 4-MDBT/1-MDBT<br />

(MDR) versus three maturity parameters (the ratios (4,6/1,4)-DMDBT, (2,4/1,4)-<br />

DMDBT, and DBT/phenanthrene (Younes and Philp, 2005).<br />

2.3 Isotope <strong>geochemistry</strong><br />

Isotope measurements combined with the study <strong>of</strong> molecular structures,<br />

including those <strong>of</strong> biomarkers, are common and very efficient approach in modern<br />

<strong>organic</strong> <strong>geochemistry</strong>. Isotope analysis <strong>of</strong> individual compounds separated online by<br />

gas chromatography (Gas Chromatography-Isotope Ratio Mass Spectrometer) also<br />

known as compound specific δ 13 C analysis have been applied to oil-oil and oil-source<br />

rock correlations and the elucidation <strong>of</strong> petroleum generation mechanism (Hayes et al.,<br />

1990; Rooney et al., 1998). Many workers have shown that structures and isotope<br />

compositions <strong>of</strong> biomarkers provide valuable genetic information (Hayes et al., 1990;<br />

Macko, 1994; Popp et al., 1997; Huang et al., 2000; Grice et al., 2001; Lu et al., 2003;<br />

Wang et al., 2004b) and also useful for oil-source rock correlation. (Bogacheva and<br />

Galimov, 1979; Galimov, 2006). The isotopic composition <strong>of</strong> the molecule can<br />

indicate parent organisms and in turn, reveal the carbon source utilized by the<br />

producer hence its position within the ancient ecosystem (Hayes, 1993). Stable<br />

isotope compositions <strong>of</strong> <strong>organic</strong> matter and carbonates are widely used in<br />

palaeoenvironmental studies <strong>of</strong> ancient ecosystems and for the assessment <strong>of</strong><br />

palaeoclimate (Sabel et al., 2005; Lis et al., 2006). However interpretation <strong>of</strong> the<br />

stable isotope records is <strong>of</strong>ten complicated by the effect <strong>of</strong> facies dependent variations<br />

in the sedimentary environment (Eh, pH, salinity, temperature etc.) and post<br />

depositional events during and after early diagenesis that can substantially influence<br />

the isotopic composition <strong>of</strong> <strong>organic</strong> matter and carbonates (Benner et al., 1987).<br />

By convention, the relative abundance <strong>of</strong> stable isotopes is always referenced to the<br />

heavy isotopes; thus an increase in the 15 N/ 14 N ratio would be reported rather than a<br />

decrease in the 14 N/ 15 N.<br />

51


This convention is codified in the ubiquitous delta (δ) notation. The δ 13 Csample<br />

=[(Rsample – Rstandard)/Rstandard] x 1000‰, where R is the absolute 13 C/ 12 C ratio <strong>of</strong> the<br />

sample or an internationally accepted standard (summarized in Table 2). The delta<br />

value is the relative difference in isotope ratio between sample and standard, and is<br />

expressed in units <strong>of</strong> permil (‰) or parts per thousand. Equivalent delta notation is<br />

used for all <strong>of</strong> the stable isotopes i.e. δ 2 H (more commonly δD), δ 15 N, δ 18 O, δ 34 S, &<br />

δ 37 Cl.<br />

2.3.1 Stable Carbon Isotope<br />

Stable carbon isotopes have played an important role in many <strong>aspects</strong> <strong>of</strong><br />

petroleum <strong>geochemistry</strong>. Early applications <strong>of</strong> bulk carbon isotope values included a<br />

study by Silverman and Epstein (1958) <strong>of</strong> a number <strong>of</strong> Tertiary crude oils from<br />

different environment. Marine and non-marine oils have been differentiated on the<br />

basis <strong>of</strong> their carbon isotopic compositions (Revill et al., 1994; Hunt et al., 2002;<br />

Meyers, 2003). Other early applications noted a general trend <strong>of</strong> enrichment <strong>of</strong> the<br />

light 12 C isotope with increasing age (Welte et al., 1975; Stahl, 1977), possibly<br />

caused by variations in intensity <strong>of</strong> photosynthesis and changes in the isotopic<br />

composition <strong>of</strong> the atmospheric CO2. Kvenvolden and Squires (1967) successfully<br />

used isotopic age trend to relate and distinguish crude oils on regional basis in West<br />

Texas. The δ 13 C value may to a certain extent reflect the chemical composition <strong>of</strong><br />

<strong>organic</strong> material originating from the same vegetation type (Chabbi et al., 2007). The<br />

distribution <strong>of</strong> 13 C among natural products is a sensitive palaeoenvironmental<br />

indicator (Hayes et al., 1990) and can provide a great deal <strong>of</strong> information about<br />

ancient biogeochemical processes (Hayes, 1993). Carbon isotopic compositions <strong>of</strong><br />

low rank <strong>coal</strong>s have been suggested to be useful in reconstruction <strong>of</strong> changes in the<br />

global carbon cycle as well as climatic changes (Lücke et al., 1999; Arens et al.,<br />

2000).<br />

52


Also, it has been shown that n-alkane distributions in kerogen using conventional<br />

geochemical tools, without detailed stable carbon isotope data can sometimes lead to<br />

erroneous interpretations <strong>of</strong> the precursors contributing to the kerogen even with<br />

samples containing abundant biomarkers (Audino et al., 2002). Several studies have<br />

utilized the stable carbon isotopic compositions <strong>of</strong> individual n-alkanes to resolve<br />

source <strong>of</strong> n-alkanes or n-alkanes precursors in ancient sediments (Kennicult and<br />

Brooks, 1990; Collister et al., 1994; Boreham et al., 1994, Chabbi et al., 2007).<br />

53


Isotopes Sample<br />

gas<br />

Table 2.1 : Relevant characteristics <strong>of</strong> the light stable isotopes (Sessions, 2006).<br />

Interface type<br />

2 H/ 1 H H2 Pyrolysis<br />

13 C/ 12 C CO2 Combustion<br />

15 N/ 14 N N2<br />

Combustion/<br />

Reduction<br />

18 O/ 16 O CO Pyrolysis<br />

Reference<br />

standard<br />

(name)<br />

Water<br />

(VSMOW)<br />

Carbonate<br />

(VPDB)<br />

Isotope ratio<br />

<strong>of</strong> standard<br />

54<br />

Theoretical<br />

Sensitivity,<br />

nmol<br />

Typical<br />

precison (%0)<br />

Typical<br />

sensitivity (%0)<br />

First<br />

commercial<br />

GC-IRMS<br />

instrument<br />

0.0000 15576 21 2-5 10-50 1998<br />

0.011224 0.024 0.1-0.3 0.1-5 1988<br />

Air (AIR) 0.003663 0.11 0.3-0.7 1-10 1992<br />

Water<br />

(VSMOW)<br />

0.00200 52 0.19 0.3-0.6 4-14 1996<br />

34 S/ 32 S SO2 Not Available Trolite (VCDT) 0.04416 0.0048 Not Available Not Availble Not Available<br />

37 Cl/ 35 Cl CH3Cl Not Available<br />

Chloride<br />

(SMOC)<br />

0.3196 0.00066 Not Available Not Available Not Available


The distribution pattern <strong>of</strong> individual n-alkane carbon isotopes in different kinds<br />

<strong>of</strong> oil should respond to variations in <strong>organic</strong> source character (Murray et al., 1994).<br />

The shape <strong>of</strong> n-alkane carbon isotopes with a trend towards isotopically lighter values<br />

has been suggested for terrestrial <strong>organic</strong> matter; higher plant origin (Murray et al.,<br />

1994; Schouten et al., 2001; Yangming et al., 2005). Marine oil commonly show a<br />

flat or positively sloping pr<strong>of</strong>ile for n-alkane isotopes, few studies have been reported<br />

with respect to the sloping pr<strong>of</strong>iles <strong>of</strong> n-alkane carbon isotope from saline lake<br />

environments (Murray et al., 1994). However, as suggested by Colister et al. (1994),<br />

the mixing <strong>of</strong> materials from multiple sources could complicate n-alkane carbon<br />

isotope pr<strong>of</strong>iles, because n-alkanes and their precursors are produced by a broad<br />

range <strong>of</strong> organisms (e.g. algae, bacteria, and terrestrial vascular plants), with some<br />

overlap among different classes <strong>of</strong> organisms.<br />

2.3.2 Hydrogen Isotopes<br />

Hydrogen isotope system is flexible, robust, accurate and provides a means for<br />

measuring D/H ratios in many geochemically important <strong>organic</strong> compounds,<br />

including hydrocarbons, sterols and fatty acids (Sessions et al., 1999). Hydrogen has<br />

two (2) stable isotopes: proton and deuterium. The relative amount <strong>of</strong> these two<br />

isotopes varies widely in different <strong>organic</strong> materials as a result <strong>of</strong> natural processes,<br />

which cause isotopic fractionation. There is little isotopic variability within specific<br />

compound classes in individual organisms, but there is significant variability between<br />

compound classes, even among those that share common biosynthetic origins.<br />

Hydrogen isotope (D/H) ratios <strong>of</strong> bulk <strong>organic</strong> hydrogen are a useful diagnostic tool<br />

in fossil fuel research (Schimmelmann et al., 2004). The integration <strong>of</strong> hydrogen<br />

isotopic data with conventional molecular parameter could provide improved<br />

constraints for source depositional paleoenvironments, origin <strong>of</strong> <strong>organic</strong> matter, oil<br />

maturity, migration and reservoir fluid mixing (Peters, 2000; Tuo et al., 2006). The<br />

hydrogen isotope specific analysis was applied to oil and oil-source rock correlations<br />

(Sessions et al., 1999; Sache et al., 2004; Tuo et al., 2006).<br />

55


It was suggested that marine-derived sedimentary <strong>organic</strong> matter is likely to have<br />

δD values near –150 ‰ if no significant isotopic exchange has occurred as a result <strong>of</strong><br />

secondary processes (Santos Neto and Hayes, 1999). Oils derived from terrestrial<br />

<strong>organic</strong> matter preserve a much larger range <strong>of</strong> δD values than do oil derived from<br />

marine <strong>organic</strong> matter (Schimmelmann et al., 2004). This presumably reflects the<br />

greater variability <strong>of</strong> δD in the terrestrial hydrological cycle and provides a useful<br />

tool for oil-to-source correlations (Tuo et al., 2006). The D/H ratios were also used as<br />

a tool for correlating terrestrially sourced oils with potential rocks and for<br />

recognizing related oils in complex basins (Schimmelmann et al., 2004). Thus,<br />

compound specific isotope analysis can be used to enhance oil-source correlation in<br />

<strong>coal</strong> bearing basin and greater understanding in the source <strong>of</strong> crude oils (Tuo et al.,<br />

2006). The bulk δD values in oils can serve as proxy for isotopic composition <strong>of</strong><br />

ancient waters and thus serves as auxillary tool for paleoclimatic assessments (Santos<br />

Neto and Hayes, 1999). The recent advent <strong>of</strong> compound specific hydrogen isotope<br />

analysis (Burgoyne and Hayes, 1998; Hilkert et al., 1999) has allowed measurement<br />

<strong>of</strong> the D/H composition <strong>of</strong> individual compounds in complex mixtures. Hydrogen<br />

isotopes ratio <strong>of</strong> individual C27, C29 and C31 alkanes respond similarly to climatic and<br />

environmental factors based on their similar range <strong>of</strong> values that is consistent with<br />

their common origin from higher plant leaf waxes (Liu and Huang, 2005).<br />

Considerable differences in hydrogen isotopic compositions occur among n-alkanes<br />

from different terrestrial depsitional environments. A trend <strong>of</strong> depletion in deuterium<br />

is observed for individual n-alkanes from terrestrial depositional environments in the<br />

order <strong>of</strong> saline lacustrine, to fresh water paralic lacustrine, and to swamp. Therefore,<br />

the δD <strong>of</strong> individual n-alkanes can be used to assess the origin <strong>of</strong> the <strong>organic</strong> matter<br />

and depositional environment <strong>of</strong> the source rocks, and may possibly be helpful in<br />

differentiating <strong>coal</strong>-derived oils from interbedded mud-stone-derived oils in <strong>coal</strong><br />

measures (Xiong et al., 2005; Tuo et al., 2006). However additional independent<br />

geochemical evidence should be taken into account when depositional environments<br />

<strong>of</strong> source rocks are reconstructed based on δD values <strong>of</strong> individual n-alkanes (Xiong<br />

et al., 2005).<br />

56


2.3.3 Sulphur Isotopes<br />

Sulphur isotope ratios have been used to solve correlation problem but not as<br />

widely employed as carbon isotopes (Hunt, 1996) because <strong>of</strong> the alteration <strong>of</strong> the<br />

original δ 34 S <strong>of</strong> oils by secondary processes (Orr, 1974). Thode (1981) was able to<br />

group thirty-eight oils from the Willston Basin <strong>of</strong> North Dakota and Saskatchewan<br />

into source rocks <strong>of</strong> Ordovician, Mississippian (Bakken), and Pennsylvanian (Tyler)<br />

age by their sulphur isotope ratios. Bordenave and Burwood (1990) used a crossplot<br />

<strong>of</strong> δ 13 C versus δ 34 S to separate two groups <strong>of</strong> Iranian oils. It has also been shown<br />

through sulphur isotope ratios that the heavy immature Monterey oils from Miocene<br />

Monterey Formation was generated primarily by the Lower Monterey shale section;<br />

in which the <strong>organic</strong> sulphur <strong>of</strong> the kerogen is >1 %. Thus high sulphur content is<br />

characteristic <strong>of</strong> type IIA (II-S) kerogen and further support that the high sulphur<br />

kerogens are the prime generators <strong>of</strong> immature, heavy oils (Hunt, 1996).<br />

2.3.4 Oxygen Isotopes<br />

Information concerning fractionation during photosynthesis <strong>of</strong> <strong>organic</strong> oxygen is<br />

limited to studies <strong>of</strong> the δ 18 O <strong>of</strong> cellulose and other carbohydrates. Isotopic<br />

composition <strong>of</strong> plant water determines the δ 18 O <strong>of</strong> <strong>organic</strong> bound oxygen in cellulose.<br />

While there is no difference in δ 18 O <strong>of</strong> cellulose that can be related to photosynthetic<br />

pathway, such as the C3 or C4 fixation pathway, real differences in the δ 18 O <strong>of</strong><br />

cellulose between species exists. The possible mechanism <strong>of</strong> these differences is<br />

differential enrichment <strong>of</strong> leaf water with 18 O during evaporative transpiration<br />

(Dongmann et al., 1974). Variations in the δ 18 O were recorded in submerged aquatic<br />

plants (Sternberg et al., 1984). Overall photosynthetic rates and productivity <strong>of</strong> the<br />

plant may also cause variability in δ 18 O.The relationship between climate, leaf water<br />

and δ 18 O <strong>of</strong> cellulose may be substantially modified by subsequent exchange events.<br />

57


2.4 Analytical Methods Employed in Geochemical Studies<br />

Analytical techniques commonly in use for geochemical analysis has advanced<br />

greatly over the last two decades. Some <strong>of</strong> these methods include Pyrolysis (hydrous,<br />

non-hydrous and Rock-Eval pyrolysis), Gas Chromatography (GC), Gas<br />

Chromatography-Mass Spectrometry (GC-MS), Pyrolysis-Gas Chromatography<br />

Mass Spectrometry (Py-GC-MS), with the most recent addition <strong>of</strong> Gas<br />

Chromatography-Isotope Ratio Mass Spectrometry (IRMS). These techniques are<br />

briefly discussed below:<br />

2.4.1 Pyrolysis<br />

Most <strong>organic</strong> matter in source rock is insoluble in common <strong>organic</strong> solvents<br />

and cannot be analysed directly by GC and GC/MS.Asphalthene which is a vital<br />

constituent <strong>of</strong> source rock extracts and crude oil has limited solubility in most<br />

solvents after isolation by precipitation and thus not directly amenable to GC and<br />

GC/MS.Therefore characterization <strong>of</strong> asphalthenes and kerogens is possible through<br />

their initial degradation by pyrolysis. Pyrolysis is an important step in all <strong>coal</strong> thermal<br />

conversion processes. It is related with <strong>coal</strong> hydrogenation, combustion and<br />

gasification. It is responsible for up to 70 % <strong>of</strong> the weight loss <strong>of</strong> <strong>coal</strong>, and it is<br />

strongly dependent on its <strong>organic</strong> properties. The various forms <strong>of</strong> pyrolysis include:<br />

58


2.4.1.1 Flash or Rapid temperature-programmed pyrolysis: This is a fingerprinting method<br />

in which pyrolysis products are transferred rapidly to a GC column. The same<br />

approach can also be used in an <strong>of</strong>fline mode where the pyrolysis products are<br />

trapped and then fractionated, if necessary, prior to analysis by GC or GC/MS<br />

(Douglas and Grantham, 1974). Using a pyrobe analytical pyrolyser, directly coupled<br />

to the inlet <strong>of</strong> a gas chromatograph and equipped with a mass spectrometer as<br />

detector, pyrolysis becomes a powerful analytical technique for studying and<br />

identifying the products <strong>of</strong> <strong>coal</strong> heating processes. Often, pyrolysis studies are<br />

performed in small-scale reactors, simulating industrial scale reactors: wire mesh<br />

reactor, curie-point reactor, pyrobe reactor, thermo-gravimetric reactor and<br />

entrained-flow reactor (Bonfanti et al., 1997). A lot <strong>of</strong> products can be identified<br />

with the coupling pyrobe-gas chromatograph-mass spectrometer (Bonfanti et al.,<br />

1997).<br />

2.4.1.2 Hydrous pyrolysis: Introduction <strong>of</strong> hydrous pyrolysis makes a significant<br />

development in pyrolysis technique (Lewan et al., 1979; Winters et al., 1983). It is<br />

undertaken in a high-pressure vessel in the presence <strong>of</strong> water. Reactions are typically<br />

performed just below the critical temperature <strong>of</strong> water. By performing a series <strong>of</strong><br />

reactions at different temperature one can simulate the fate <strong>of</strong> the <strong>organic</strong> matter at<br />

different burial depths. Information that is obtained in this manner can be used for<br />

basin modelling purposes, or to determine the nature <strong>of</strong> the products generated and to<br />

correlate them with other oils in the basin. The most significant impact <strong>of</strong> hydrous<br />

pyrolysis is that it permits assessments <strong>of</strong> the fate <strong>of</strong> an immature rock as it is<br />

matured at greater depths <strong>of</strong> burial. The related method <strong>of</strong> hydrous pyrolysis applies<br />

thermal energy in the presence <strong>of</strong> water (Lewan et al., 1979; Winter et al., 1983;<br />

Lewan, 1985; Barth et al., 1989) and has proved useful for releasing hydrocarbon<br />

biomarkers weakly bound via sulphur or oxygen linkages to the kerogen moiety <strong>of</strong><br />

immature sulphur rich <strong>organic</strong> sediments (Greenwood et al., 2006).<br />

59


2.4.1.3 Rock-Eval pyrolysis: This technique was developed by Espitalié et al. (1977).<br />

It is a standardized pyrolysis method for source rock characterization and evaluation.<br />

It gains widespread use in petroleum industry, where it became a principal analytical<br />

tool. During the last 30 years the Rock-Eval apparatus has been routinely used in<br />

<strong>organic</strong> <strong>geochemistry</strong> for examining the oil and gas potential <strong>of</strong> source rocks. Many<br />

source evaluation parameters can be generated from the Rock-Eval analysis <strong>of</strong> source<br />

rock (Table 3). Less than 100 mg <strong>of</strong> a sample is heated in an inert or oxidative<br />

atmosphere under a programmed temperature pattern. The Rock-Eval instrument<br />

mode <strong>of</strong> operation is shown in Fig. 4. Evolution <strong>of</strong> volatiles is examined in Rock<br />

eval. A s<strong>of</strong>tware OPTKIN has been developed for acquisition <strong>of</strong> pyrolysis kinetic<br />

parameters using the Rock-Eval measurements. The more matured is an <strong>organic</strong><br />

matter; the higher is its Tmax. Rock-Eval pyrolysis is used to identify the type<br />

(kerogen) and maturity <strong>of</strong> <strong>organic</strong> matter and to detect petroleum potential in fossil<br />

fuels (Hunt, 1996; Peters et al., 2005; Moumouni et al., 2007). It has been noted that<br />

Hydrogen Index (HI) values obtained from Rock-Eval analysis can be misleading, as<br />

hydrocarbons may be adsorbed by the rock matrix (Langford and Blanc- Valleron<br />

1990). Shale source rocks may therefore yield Rock-Eval pyrolysis-generated HIs<br />

that are less than true average hydrogen index, while <strong>coal</strong>y source rocks may have<br />

their HIs over-estimated. Therefore proposed the use <strong>of</strong> TOC vs. S2 and measuring<br />

the adsorption <strong>of</strong> hydrocarbons by the rock matrix.<br />

60


S2<br />

S1<br />

Free hydrocarbon, FID<br />

Hydrocarbon, FID<br />

Distillation, 300 o C<br />

Total <strong>organic</strong> matter<br />

Pyrolysis, 300-650 o C<br />

Fig. 2.3 :Scheme <strong>of</strong> Rock-Eval analysis<br />

61<br />

S3<br />

Kerogen<br />

S4<br />

COX (IR)<br />

Residual<br />

fraction<br />

Oxidation,<br />

850 o C<br />

CO2 (IR)


Table 2.2:Rock-Eval Parameters (Johannes et al., 2007).<br />

Sample Formula Description<br />

S1 (mgHC/g sample) ― Free hydrocarbon (HC)<br />

S2 (mgHC/g sample) ―<br />

62<br />

Hydrocarbon generated through thermal<br />

cracking<br />

S3 (mg CO2/g sample) ― Amount <strong>of</strong> CO2 produced during pyrolysis<br />

S4 (mg CO2/g sample) ― Amount <strong>of</strong> CO2 produced during combustion<br />

Tmax ( o C) ― The temperature at top <strong>of</strong> S2 peak<br />

PI S1/S1+S2 Production Index<br />

PC (%)<br />

0.1[0.83(S1+S2) +0.273S3<br />

+0.429(S3CO+0.53S3′CO)]<br />

Pyrolysable <strong>organic</strong> carbon<br />

TOC (%) PC+RC Total <strong>organic</strong> carbon<br />

BI (mg HC/g TOC) 100S1/TOC Bitumen Index<br />

HI (mg HC/g TOC) 100S2/TOC Hydrogen Index<br />

OI (mg CO2/g TOC) 100S3/TOC Oxygen Index<br />

RC (%) RC CO + RC CO2<br />

Residual <strong>organic</strong> carbon


2.4.2 Gas chromatography/Mass spectrometry<br />

The advent <strong>of</strong> GC/MS makes rapid analysis <strong>of</strong> an <strong>organic</strong> mixture with more<br />

than 20 components possible (Simoneit, 2002). It requires very small sample size,<br />

rapid and very high sensitivity. Much <strong>of</strong> the early work done was accomplished using<br />

magnetic-sector mass spectrometers. When the quadrupole MS became<br />

commerecially available in the mid-1970s the popularity <strong>of</strong> GC/MS increased<br />

dramatically (Simoneit, 2002). Quadrupole systems although lacking high resolution<br />

capabilities <strong>of</strong> magnetic sector instruments, were able to scan rapidly, did not have<br />

any memory effects, were sensitive, easy to use; and did not require the elaborate<br />

turning procedures <strong>of</strong> the magnetic sector systems. Computerized GC/MS is the<br />

principal method used to evaluate biomarkers in fossil fuels. GC/MS can be used to<br />

detect and provisionally identify compounds using relative gas chromatographic<br />

retention times, elution patterns, and the mass spectral fragmentation patterns<br />

characteristic <strong>of</strong> their structures. GCMS analyses can be operated in various modes.<br />

Each mode provides a different type and quality <strong>of</strong> information.<br />

2.4.3 Gas Chromatography-Isotope Ratio Mass Spectrometer (GC-IRMS)<br />

GC-IRMS was first described by Mathews and Hayes (1978), but commercial<br />

systems were available in the late 1980s (Simoneit, 2002). High precision isotope-<br />

ratio detection <strong>of</strong> gas chromatographic analyte is now a mature technology, and is<br />

enjoying widespread growth in a variety <strong>of</strong> fields including bio<strong>geochemistry</strong> and<br />

paleoclimatology, archaeology, petroleum chemistry, environmental forensic, flavour<br />

authentication, drug testing, metabolite tracing and others (Sessions, 2006). GC-<br />

IRMS has been very useful in <strong>coal</strong> research through measurement <strong>of</strong> carbon,<br />

hydrogen, oxygen, nitrogen and sulphur isotopes <strong>of</strong> biomarkers. For instance,<br />

hydrogen isotopes are commonly fractionated to a much greater extent by different<br />

processes than carbon isotopes and better respond to the hydrological cycle e.g.<br />

meteoric water availability and isotope differences. Hydrogen isotope ratios are<br />

strongly influenced by environment and biochemical studies (Bi et al., 2005).<br />

63


Total <strong>organic</strong> hydrogen in <strong>coal</strong> encompasses hydrogen that is linked to carbon<br />

either directly or via bridging heteroatoms like O, S and N.Some <strong>of</strong> this <strong>organic</strong><br />

hydrogen can isotopically exchange with hydrogen from ambient water, with<br />

exchange half-lives ranging from seconds (in exposed hydroxyl, carboxyl and amino<br />

groups) to millions <strong>of</strong> years; aliphatic carbon linked hydrogen (Schimmelmann et al.,<br />

1999). The abundance <strong>of</strong> readily exchangeable <strong>organic</strong> hydrogen is expressed as<br />

“hydrogen exchangeability”. Exchangeable <strong>organic</strong> hydrogen is the most chemically<br />

reactive, polar hydrogen, and thus participates in oil and gas generation from <strong>coal</strong>. In<br />

contrasts, non-exchangeable <strong>organic</strong> hydrogen may have preserved an isotopic<br />

paleoenvironmental signal, but deuterium/hydrogen (D/H or 2 H/ 1 H) stable isotope<br />

ratios in <strong>coal</strong>s or <strong>coal</strong> kerogens are typically measured only for total hydrogen<br />

(Mastalerz and Schimmelmann, 2002). The isotopically non-exchangeable portion <strong>of</strong><br />

<strong>organic</strong> hydrogen is chemically more stable than exchangeable hydrogen, and δDn<br />

values <strong>of</strong> non-exchangeable hydrogen may therefore provide information about the<br />

depositional environments (Mastalerz and Schimmelmann, 2002). Organic hydrogen<br />

linked to oxygen and nitrogen tends to be more enriched in deuterium than carbon-<br />

linked <strong>organic</strong> hydrogen (Schimmelmann et al., 1999). Some <strong>of</strong> this 2 H in functional<br />

groups is sterically inaccessible to water during isotopic equilibration and thus<br />

contributes to the pool <strong>of</strong> non-exchangeable hydrogen (Mastalerz and Schimmelmann,<br />

2002). The integration <strong>of</strong> hydrogen isotopic data with conventional molecular<br />

parameter could provide improved constraints for source depositional<br />

paleoenvironments, origin <strong>of</strong> <strong>organic</strong> matter, oil maturity, migration and reservoir<br />

fluid mixing (Peters 2000; Tuo et al., 2006). Tuo et al. (2006) compared the values <strong>of</strong><br />

δD measured for diterpenoids compounds with n-alkanes present in <strong>coal</strong> and reported<br />

that diterpenoid hydrocarbons are 49-81 ‰ depleted in D relative to n-alkanes.<br />

Hydrogen isotopic variations also occur between different diterpenoid compounds,<br />

indicating a different source for these compounds (Tuo et al., 2006).<br />

64


Based on the comparison <strong>of</strong> δD values <strong>of</strong> different classes <strong>of</strong> diterpenoid<br />

compounds, which may be derived from the same biosynthetic precursors, a D<br />

enrichment process (especially dehydrogenation) will be expected when a compound<br />

is diagenetically altered from a natural precursor structure to a geological structure<br />

(Tuo et al., 2006). Saline lakes are likely D-enriched due to evaporation <strong>of</strong> water,<br />

while meteoric water as a main hydrogen source is relatively D-depleted in various<br />

terrestrial environments (Xiong et al., 2005).<br />

65


2.5 Geology <strong>of</strong> Benue Trough<br />

2.5.1 Geological Settings<br />

The Benue Trough (Fig. 2.4) is considered to have formed by the incipient<br />

rifting during the breakaway <strong>of</strong> South America from Africa and the opening <strong>of</strong> south<br />

Atlantic in Early Cretaceous (Albian) times (Whiteman, 1982). It trends SSW-NNE<br />

for about 800 km in length and 150 km in width (Abubakar et al., 2006; Jauro et al.,<br />

2007). The Benue Trough divide at its upper end to; the Gongola arm running north<br />

into the Chad basin and the Yola arm terminating eastwards against the Cameroon<br />

basement. There are various amounts <strong>of</strong> uplift and deformation that created a regional<br />

unconformity in the Benue Trough.This is because full rifting never developed and its<br />

sedimentary fill, accumulated in about 20Ma since Albian time, was folded by<br />

complex stress created at the African plate margin in Santonian thereby regarded as<br />

‘Failed Arm’ (Pearson and Obaje, 1999). The Trough contains about 6000 m thick <strong>of</strong><br />

Cretaceous-Palaeogene sediments that are <strong>organic</strong> rich in part. The Benue Trough can<br />

be physiographically and lithostratigraphically subdivided into lower, middle and<br />

upper BenueTrough (Fig. 2.5). The geology <strong>of</strong> Benue Trough has been extensively<br />

reviewed (Carter et al., 1963; Peters, 1982; Peters and Ekweozor, 1982; Obaje 1994,<br />

2000; Zaborski 2000; Obaje et al., 2004). Overlying the basal rift sandstones <strong>of</strong> Late<br />

Necomian to Aptian age are thick shales and thin limestone deposited during marine<br />

trangression (Peters and Ekweozor, 1982) in the Middle to Late Albian, Late<br />

Cenomanian to Early Turonian, Late Turonian to Early Santonian, Campanian to<br />

Maastrichtian, Paleocene, and Eocene.<br />

66


Fig. 2.4 :Geological map <strong>of</strong> Benue Trough, Nigeria (modified after Obaje, 1994).<br />

67


Fig. 2.5: Stratigraphic sequence <strong>of</strong> Benue Trough, Nigeria (Modified by Ehinola<br />

et al., 2002).<br />

68


The shale limestone sequence formed during Late Cenomanian to Early Turonian<br />

includes Odukpani Formation <strong>of</strong> Calabar area, Eze-Aku/Makurdi Formation that<br />

extends from North <strong>of</strong> Ishiagu in the Lower Benue Trough and Dukul, Gongila and<br />

Pindiga Formations <strong>of</strong> Upper Benue Trough in the North Eastern Nigeria.The<br />

limestone beds in these Formations are sufficiently thick in some places. After the<br />

deposition <strong>of</strong> the limestone bearing sequence, the sea became shallower resulting in<br />

the formation <strong>of</strong> swamps particularly in the Anambra basin. The swamps and<br />

associated vegetation were later buried under thick sediments to produce <strong>coal</strong>-bearing<br />

rocks.<br />

2.5.2 Lithology Description<br />

The <strong>coal</strong> bearing rocks consist essentially <strong>of</strong> Mamu Formation (lower <strong>coal</strong><br />

measure), Ajali sandstone (middle <strong>coal</strong> measure), and Nsukka Formation (upper <strong>coal</strong><br />

measures) in Anambra basin and Lower Benue Trough.In the Middle and Upper<br />

Benue Trough, <strong>coal</strong>iferous beds also occur within Lafia and Gombe Formations,<br />

which are considered as time equivalents <strong>of</strong> the Mamu and Nsukka Formations in the<br />

southeastern Nigeria. The Mamu Formation consists <strong>of</strong> shale, siltstones, sandstones<br />

and <strong>coal</strong>s <strong>of</strong> a fluvio-deltaic to fluvio-estuarine environments whose lateral<br />

equivalents are the conglomerates, cross-bedded and poorly sorted sandstones and<br />

claystones <strong>of</strong> the Lokoja and Bida Formations in Bida basin (Akande et al., 2005).<br />

Most <strong>of</strong> the previous authors described the depositional environment <strong>of</strong> the entire<br />

Benue Trough as shallow marine and (in parts) anoxic (Peters 1977,1982). However,<br />

more recent works on sedimentological structures and facies indicate different<br />

depositional environments, at least in the Benue Trough based on new interpretation<br />

<strong>of</strong> the limestone layers as turbidites and new faunal analysis (Gedhardt, 2001). On the<br />

other hand, middle Benue Trough was deposited under marginal marine (paralic) and<br />

continental conditions (Obaje, 1994; Gebhardt, 2001).<br />

69


In the middle Benue Trough, the Precambarian Basement is overlain<br />

uncomformably by the Asu River Group, and the Keana, Makurdi and<br />

Awgu Formation overlie this unconformably while Lafia Formation unconformably<br />

overlies the Awgu Formation (Ehinola et al., 2002). Different periods that were<br />

recognized in the Middle Benue Trough are Albian, Cenomanian, Turonian,<br />

Coniacian, Campanian-Maastrichtian and Paleocene.These periods have been<br />

previously described (Ehinola et al., 2002). The upper Benue Trough bifurcates at its<br />

northeastern end into the Gongola and the Yola basins (Abubakar et al., 2006; Jauro<br />

et al., 2007). In this part <strong>of</strong> Benue Trough, a compressional phase occurred at the end<br />

<strong>of</strong> the Maastrichtian.This event resulted in the folding and faulting <strong>of</strong> pre-Palaeogene<br />

sediments (Abubakar et al., 2006). The oldest sediment consists <strong>of</strong> the Late Jurassic<br />

to Albian continental Bima sandstone that rests uncomformably on Precambrian<br />

basement rocks (Abubakar et al., 2006). The Bima Sandstone is comformably<br />

overlain by the transitional marine Yolde Formation which is followed by marine<br />

sequences <strong>of</strong> Pindiga Formation in the Gongola Basin, and their lateral equivalent <strong>of</strong><br />

Dukul, Jessu/Sekiliye, Numanha and Lamja Formations in the Yola Basin.The<br />

youngest Cretaceous sediment in the upper Benue Trough is restricted to the Gongola<br />

Basin.It is represented by the coastal (deltaic) Gombe Sandstone which<br />

uncomformably covered the pre-mid-Santonian sequences in several places<br />

(Abubakar et al., 2006).The continental sands,silts and shales <strong>of</strong> the Palaeogene<br />

Kerri-Kerri Formation uncomformably overlie the Gombe Sandstone and mark the<br />

end <strong>of</strong> sedimentation in the upper Benue Trough.Detailed description <strong>of</strong> upper Benue<br />

Trough have been reported (Abubakar et al., 2006; Jauro et al., 2007).Regressive<br />

sandstones along the basin margin interfinger with basinal marine and paralic<br />

shale.These marine cycles and the early Cenomanian and Late Santonian<br />

uncomformities in the Benue Trough provides a practical basis for subdividing the<br />

sedimentary succession into the Albian Asu River group,a Late Cenomanian Early<br />

Santonian sequence(Cross River group),and a post Satonian deltaic,Coal measures<br />

and Paralic sequence (Peters and Ekweozor, 1982).<br />

70


CHAPTER THREE<br />

EXPERIMENTAL<br />

3.1 SAMPLING AND SAMPLE PREPARATION<br />

3.1.1 Sampling<br />

A total <strong>of</strong> nine samples comprising <strong>of</strong> six <strong>coal</strong>s, two carbonaceous shales and one<br />

<strong>coal</strong>y shale were collected from 2 boreholes (BH94 and BH120) from Lafia-Obi,<br />

Awgu Formation. The <strong>coal</strong> seams and interbedded shale in BH94 and BH120 were<br />

sampled between 218-431 m and 131- 289 m depths respectively. The geological map<br />

and detailed lithographic descriptions <strong>of</strong> the boreholes are presented in Fig. 3.1. In<br />

Mamu Formation, twelve samples consisting <strong>of</strong> nine <strong>coal</strong>s and three carbonaceous<br />

shales were collected from Okaba and Onyeama.Okaba mine exposed on the floor <strong>of</strong><br />

the Okaba open cast mine located at Odagbo village (N07 o 28’ 41.2”, E007 o 43’46.<br />

4”). The Onyeama seam is exposed along the Asata river channel situated at Enugu<br />

escarpment (N06 o 28’ 23.8”, E007 o 26’ 44.7”). The lithographic section <strong>of</strong> Okaba and<br />

Onyeama <strong>coal</strong> were presented in Fig. 3.2.<br />

71


Fig. 3.1: Geological map and Lithographic section <strong>of</strong> BH94 and BH120 <strong>of</strong><br />

Lafia-Obi <strong>coal</strong>, Awgu Formation (Ehinola et al., 2002).<br />

72


Fig. 3.2: Lithographic section <strong>of</strong> Onyeama and Okaba Mines in Mamu<br />

Formation.<br />

73


3.1.2 Extraction <strong>of</strong> Soluble Organic Matter<br />

The samples were crushed with agate mortar and powdered to less than 120<br />

mesh size prior to extraction. The powdered samples were Soxhlet extracted with<br />

chlor<strong>of</strong>orm for 72 h. Activated copper powder was added to remove elemental sulphur<br />

from the extracts. Excess solvent was distilled <strong>of</strong>f using a hot water bath to an aliquot<br />

volume <strong>of</strong> about 3 ml. The aliquot was then transferred to a weighed clean vial with a<br />

micropipette and the remaining solvent removed under nitrogen gas flow at<br />

temperature below 50 o C. The soluble <strong>organic</strong> matter was then weighed on a Mettler<br />

balance.<br />

3.1.3 Asphalthene Isolation<br />

The asphalthene was separated from the <strong>coal</strong> extracts by gradual addition <strong>of</strong> n-<br />

hexane with constant stirring (40 ml <strong>of</strong> n-hexane to 1 ml <strong>of</strong> extract). The resulting<br />

mixture was kept inside dark cupboard overnight. The precipitated asphalthene was<br />

removed by filteration on filter paper and purified by soxhlet extraction with n-hexane<br />

until the solvent in the soxhlet is colourless. The asphalthene was recovered by<br />

dissolution in dichloromethane. The dichloromethane was dried <strong>of</strong>f, leaving the solid<br />

asphalthene. The asphalthene was dried at 40 o C for 8 h in oven.<br />

3.1.4 Fractionation<br />

The maltene fraction left after asphaltene isolation was separated into saturated<br />

and aromatic hydrocarbons and polar compounds (NSO compounds) by column<br />

chromatography on neutral alumina over silica gel. Chromatographic column<br />

(length/width ratio 50:1) was wet packed with activated silica gel (80-120 mesh,<br />

ASTM) and aluminium oxide (70-325 mesh, ASTM) using n-hexane. Saturated,<br />

aromatic and polar fractions were eluted with hexane (150 ml), dichloromethane (150<br />

ml), and methanol (30 ml) respectively. The solvent was distilled <strong>of</strong>f using rotary<br />

evaporator to about 3 ml and thereafter transferred into a weighed clean vial. The<br />

remaining solvent was removed under nitrogen gas flow at temperature below 50 o C.<br />

The polar fraction was further derivatised prior to GC/MS analysis.<br />

74


3.1.5 Derivatisation <strong>of</strong> Polar fraction<br />

The polar fraction was derivatized using boron triflouric acid. 17-20 % BF3<br />

(Boron triflouric acid) was prepared by adding 200 ml <strong>of</strong> BF3 to 300 ml CH3OH<br />

(Methanol). 2 ml <strong>of</strong> prepared BF3 solution was added to about 1-2 mg polar fraction.<br />

The resulting solution was kept in the water bath at about 60 o C for 24 h. The<br />

esterified solution after 24 h was freed <strong>of</strong> the BF3 acid in a separating funnel by<br />

addition <strong>of</strong> distilled water and diethyl ether. The mixture was shaken vigorously and<br />

left for few minutes for the water and <strong>organic</strong> layer to separate. The aqueous layer<br />

containing the BF3 was drained <strong>of</strong>f. This procedure was repeated three times to wash<br />

<strong>of</strong>f the BF3 completely. The diethyl ether layer containing the derivatised polar<br />

fraction was then collected and the solvent removed by evaporation.<br />

3.2 ANALYTICAL METHODS<br />

3.2.1 Leco Analysis<br />

Total <strong>organic</strong> carbon (TOC) and Total Sulphur (TS) were measured by LECO<br />

analyses. Carbon and sulphur concentrations in whole samples were measured in<br />

duplicate by combustion in an induction furnace in a flow <strong>of</strong> oxygen, using a LECO<br />

carbon-analyzer IR 112. Total <strong>organic</strong> carbon (TOC) was measured using samples that<br />

had been treated with hydrochloric acid to remove carbonate.<br />

3.2.2 Elemental Analysis<br />

Elemental analysis (C, H, N, O) was performed using Carlo Erba 1108 CHNS-<br />

O analyser. Hydrogen was determined by sample oxidation (1100 o C) and infrared (IR)<br />

detection <strong>of</strong> H2O.The nitrogen content was obtained by thermal conductivity detection<br />

(TCD) <strong>of</strong> N2. Oxygen was analyzed separately by transformation <strong>of</strong> all oxygenated<br />

molecules into CO followed by oxidation over copper to CO2, which was measured<br />

directly by IR detector.<br />

75


3.2.3 Rock Eval Analysis<br />

Rock Eval analyses were performed using standard technique as described by<br />

Espitalié et al. (1977). This technique requires no demineralization or other<br />

preparation except powdering the whole samples.<br />

Less than 100 mg <strong>of</strong> a sample is heated in an inert or oxidative atmosphere under a<br />

programmed temperature pattern with characterization <strong>of</strong> four peaks. S1 which is the<br />

first peak represents hydrocarbon already present in the sample which are mainly<br />

stripped at temperatures about 300 o C .The second peak, S2 represents hydrocarbons<br />

generated through thermal cracking <strong>of</strong> kerogen at temperatures in the range <strong>of</strong> 300-<br />

650 o C, while S3 peak represents the CO2 which is generated from the kerogen at the<br />

same time the S2 hydrocarbons are being generated. The fourth peak, S4 indicates the<br />

amount <strong>of</strong> CO2 produced through oxidation during combustion at a temperature <strong>of</strong><br />

about 850 o C. A s<strong>of</strong>tware OPTKIN has been developed for acquisition <strong>of</strong> pyrolysis<br />

kinetic parameters. The parameters include S1, S2, S3, HI, OI, S2/S3, PI, PC% and<br />

Tmax for the assessment <strong>of</strong> kerogen qualities and maturation.<br />

3.2.4. Gas chromatography-mass spectrometry analysis (GC-MS)<br />

The GC-MS analyses <strong>of</strong> the fractions were performed on a Hewlett-Packard<br />

6890N gas chromatograph interfaced to a Hewlett-Packard 5973N Mass spectrometer.<br />

The gas chromatograph was equipped with a DB-5 MS fused silica capillary column<br />

(30 m x 0.25 mm) and helium was used as carrier gas with a flow rate <strong>of</strong> 1ml/min.<br />

The Mass spectrometer was operated with electron impact energy <strong>of</strong> 70 eV and ion<br />

source temperature <strong>of</strong> 230 o C. The GC oven temperature was isothermal for 1min at<br />

80 o C and then programmed from 80 to 280 o C at 3 o C/min and isothermal for 20 min<br />

at 280 o C. Individual saturated, aromatic and NSO- compounds were monitored by<br />

selected ion monitoring (SIM) at a cycle time <strong>of</strong> 1s.The GC-MS data were acquired<br />

and processed with a Hewlett-Packard Chemstation data system.<br />

76


3.2.5. Gas chromatography-isotope ratio mass spectrometry (GC-IRMS).<br />

The Carbon isotope analyses <strong>of</strong> individual compounds were performed on a<br />

Delta Plus XP Gas Chromatography-combustion-Isotope Ratio Mass Spectrometer.<br />

The gas chromatography was performed using a Thermo Finnigan GC<br />

COMBUSTION III system equipped with a DB-5 fused silica capillary column (30 m<br />

x 0.25 mm) and helium was the carrier gas with a flow rate <strong>of</strong> 1 ml/min. The GC oven<br />

temperature was isothermal for 1 min at 80 o C and then programmed from 80 to 280<br />

o C at 3 o C/min and isothermal for 20 min at 280 o C.<br />

Isotopic values were calculated by integrating the m/z 44,45 and 46 ion currents<br />

<strong>of</strong> the peaks produced by combustion (830 o C) <strong>of</strong> chromatographically separated<br />

compounds and those <strong>of</strong> CO2 standard spikes admitted at regular intervals. The<br />

reproducibility and accuracy were evaluated routinely using laboratory standards <strong>of</strong><br />

known δ 13 C values (C16-C32 n-alkanes). Laboratory standard was injected for every<br />

six-sample analysis. The isotope values are given with respect to the PDB standard.<br />

The analyses were repeated twice and the results presented as an average value.<br />

3.2.6. Vitrinite reflectance measurement<br />

These analyses were carried out on the polished <strong>coal</strong> samples using a Zeiss<br />

standard universal reflected microscope equipped with 100x oil immersion objective<br />

and a 40x air objective. Measuremets were done at 546 nm (wavelength) on clear<br />

spots <strong>of</strong> vitrinite particles (with distinct shape) <strong>of</strong> size approximately equal to or<br />

greater than 10 µm before each series <strong>of</strong> measurements. The microscope set up was<br />

calibrated with standard (glass) <strong>of</strong> known vitrinite reflectance within the range to be<br />

measured. Thereafter, reflectance values were read <strong>of</strong>f directly from the digital read<br />

out.<br />

77


4.1 ELEMENTAL ANALYTICAL DATA<br />

CHAPTER FOUR<br />

RESULTS AND DISCUSSIONS<br />

Representative samples from Awgu and Mamu Formation were analysed for C,<br />

H, O, N and S. The results in Table 4.1 were used to assess <strong>organic</strong> matter source,<br />

depositional environment and thermal maturity <strong>of</strong> the samples.<br />

4.1.1 Organic Matter Source and Depositional Environment<br />

4.1.1.1 Awgu Formation<br />

The C content varies from 76.04wt% to 82.10wt% with a mean <strong>of</strong> 79.40wt%<br />

whilst H ranges between (4.79 to 5.16) wt% with a mean value <strong>of</strong> 5.03 wt%. Oxygen<br />

values range from 8.10wt% to 10.40wt% with a mean value <strong>of</strong> 8.88wt% whilst<br />

Nitrogen ranges between (1.76 to 2.29) wt% with a mean value <strong>of</strong> 2.06. The S content<br />

varies from 0.38wt% to 1.29wt% with a mean value <strong>of</strong> 0.68wt%. The samples are<br />

characterized by low sulphur contents (0.38 to 1.08) except a sample from BH120<br />

with sulphur value <strong>of</strong> 1.29 wt%. The TOC/S ratio ranges from 8.1 to 52.8, indicating<br />

non-marine <strong>coal</strong> (Berner, 1984; Bechtel et al. 2007a, b). It can be inferred from these<br />

results that majority <strong>of</strong> the samples are <strong>of</strong> non-marine origin with few samples from<br />

BH120 with values ranging between 8.10wt% to 15.00wt% which suggest little<br />

marine incursion and/or <strong>organic</strong> matter deposited in lacustrine environment<br />

(Brown and Kenig, 2004).<br />

The TOC/N ratios range from 2.01-22.15 (Table 4.1). High TOC/N values (13.26-<br />

22.15) obtained in three <strong>of</strong> the samples reflect terrigenous <strong>organic</strong> matter input<br />

(Meyers, 1994; Bechtel et al., 2007a). The low values <strong>of</strong> (2.01-8.94) recorded in some<br />

<strong>of</strong> the samples, indicating the presence <strong>of</strong> marine <strong>organic</strong> matter and/or lacustrine<br />

algae (Meyers, 1994; Meyers et al., 2006; Bechtel et al., 2007a).<br />

The plot <strong>of</strong> H/C against O/C ratios (Fig. 4.1) revealed all the samples plot within<br />

the type III evolution path, indicating that the samples are derived mainly from<br />

terrigenous <strong>organic</strong> matter (Van Krevelen et al., 1961).<br />

78


4.1.1.2 Mamu Formation<br />

The C content varies from 69.87wt% to 75.72wt% with a mean <strong>of</strong> 72.83wt%<br />

whilst H ranges between (5.30 to 5.49) wt% with a mean value <strong>of</strong> 5.43 wt%. Oxygen<br />

values range from 14.82wt% to 20.51wt% with a mean value <strong>of</strong> 17.42wt% whilst<br />

Nitrogen ranges between (1.95 to 2.31) wt% with a mean value <strong>of</strong> 2.15wt%. The S<br />

content varies from 0.094wt% to 1.94wt% with a mean value <strong>of</strong> 0.46wt%. Majority<br />

<strong>of</strong> the samples have low sulphur content (0.09-0.52 wt%), except for the value <strong>of</strong><br />

1.94wt% obtained from sample OKB 8. The TOC/S ratios range from 12.9 to 234.7.<br />

The low S content as well as the values <strong>of</strong> the TOC/S ratios indicates non-marine <strong>coal</strong><br />

(Bechtel et al., 2007b; Brown and Kenig, 2004).<br />

The TOC/N ratios range from 20.49-31.51 (Table 4.1). These values reflect <strong>coal</strong><br />

derived from land plant <strong>organic</strong> matter (Meyers 1989, 1994; Dumitrescu and Brassell,<br />

2006; Bechtel et al., 2007a). The samples show low H/C (


Samples<br />

Awgu Formation<br />

LB387<br />

Table 4.1: Elemental Analysis and Vitrinite Reflectance Data <strong>of</strong> Nigerian Coal.<br />

Lithology<br />

(Rank)<br />

Carbonaceous shale<br />

Depth (m) TOC (%) % S % C % H % O % N H/C O/C C/N TOC/N TOC/S %VR<br />

218.5-222.5 4.56 0.40 76.55 5.16 10.40 2.27 0.81 0.1 33.72 2.01 11.50 0.94<br />

LB417 Coaly shale 407.4-412.5 6.72 0.45 ND ND ND ND ND ND ND ND 15.00 ND<br />

LB420 Coal (B) 417.0-422.0 13.74 0.38 76.04 4.85 9.38 1.80 0.77 0.09 42.24 7.63 35.90 ND<br />

LB514 Coal (B) 131.7-136.6 32.58 1.08 79.81 5.10 8.43 2.29 0.77 0.08 34.85 14.23 31.20 1.02<br />

LB518 Coal (B) 148.0 30.38 0.90 ND ND ND ND ND ND ND ND 33.90 ND<br />

LB523 Coal (B) 168.8-173.7 44.52 1.29 81.45 5.15 8.33 2.01 0.76 0.08 40.52 22.15 34.50 1.08<br />

LB534 Coal (B) 212.1-216.2 19.85 0.38 82.10 5.10 8.10 2.22 0.75 0.07 36.98 8.94 52.80 1.15<br />

LB542 Carbonaceous shale 247.0 5.81 0.72 ND ND ND ND ND ND ND ND 8.10 ND<br />

LB551 Coal (B) 286.0-289.0 23.34 0.51 80.42 4.79 8.66 1.76 0.71 0.08 45.69 13.26 45.40 1.15<br />

Mamu Formation<br />

OY5<br />

Carbonaceous shale<br />

4.6-5.8 4.53 0.09 ND ND ND ND ND ND ND ND 12.90 ND<br />

OY4 Coal (SB) 6.0-6.8 66.79 0.35 74.74 5.49 14.82 2.12 0.88 0.15 35.26 31.51 189.70 0.57<br />

OY3 Coal (SB) 6.8-7.3 61.09 0.39 ND ND ND ND ND ND ND ND 157.50 ND<br />

OY2 Coal (SB) 7.6-8.4 65.11 0.37 ND ND ND ND ND ND ND ND 177.90 ND<br />

OY1 Coal (SB) 8.4-8.8 46.71 0.20 75.72 5.30 14.92 1.95 0.84 0.15 38.83 23.95 234.70 0.60<br />

OY6 Carbonaceous shale 9.0-9.5 9.93 0.44 ND ND ND ND ND ND ND ND 22.60 ND<br />

OKB4<br />

Carbonaceous shale<br />

16.5-17.6 4.14 0.17 ND ND ND ND ND ND ND ND 24.60 ND<br />

OKB3 Coal (SB) 18.0-18.5 59 0.34 71.00 5.44 19.41 2.31 0.92 0.21 30.74 25.54 174.00 0.48<br />

OKB2 Coal (SB) 18.6-18.9 52.81 0.52 ND ND ND ND ND ND ND ND 101.40 ND<br />

OKB1 Coal (SB) 18.9-19.3 27.53 0.30 ND ND ND ND ND ND ND ND 93.00 ND<br />

OKB7 Coal (SB) 19.3-19.6 54.72 0.41 ND ND ND ND ND ND ND ND 132.50 ND<br />

OKB8 Coal (SB) 19.6-20.0 45.7 1.94 69.87 5.47 20.51 2.23 0.94 0.22 31.33 20.49 23.60 0.50<br />

ND – not determined<br />

OY – Onyeama<br />

OKB – Okaba<br />

(B) – Bituminous; (SB) – Sub-bituminous<br />

80


Fig. 4.1: Plot <strong>of</strong> Atomic H/C against O/C <strong>of</strong> Coal Samples from Benue<br />

Trough, Nigeria. (After Van Krevelen et al., 1961).<br />

81


4.2 Source Rock Evaluation <strong>of</strong> Nigerian Coal.<br />

The hydrocarbon potential <strong>of</strong> the <strong>coal</strong>s was assessed by Leco analysis and Rock<br />

Eval pyrolysis. Data from Leco and Rock Eval analyses are presented in Table 4.2.<br />

4.2.1 Organic Matter Concentration:<br />

The quantity <strong>of</strong> <strong>organic</strong> matter contained in the samples was evaluated from the Total<br />

Organic Carbon (TOC) content and Genetic Potential (GP) values. The TOC and GP<br />

values are presented in Table 4.2.<br />

4.2.1.1 Awgu Formation<br />

The TOC values for the carbonaceous shale and <strong>coal</strong>y shale range between<br />

4.56wt% to 6.72wt% with a mean value <strong>of</strong> 5.70wt%. While the <strong>coal</strong>, TOC values vary<br />

between 13.74wt% to 44.52wt% with a mean value <strong>of</strong> 27.40wt%. The TOC values in<br />

all the samples exceeded the minimal 0.5 wt% required for a potential source rock<br />

(Tissot and Welte, 1984; Killops and Killops, 1993; Hunt, 1996).<br />

The GP values for the carbonaceous shale and <strong>coal</strong>y shale range between<br />

3.2mg/g to 5.7mg/g with a mean value <strong>of</strong> 4.4mg/g. While the <strong>coal</strong>, GP values vary<br />

between 19.1mg/g to 90.8mg/g with a mean value <strong>of</strong> 42.4mg/g. The GP values are<br />

greater than 2 mg/g required for a potential source rock (Tissot and Welte, 1984;<br />

Killops and Killops 1993,2005; Hunt, 1996; Peters et al. 2005). The values indicate<br />

moderate to good source rock.<br />

82


Samples<br />

Awgu<br />

Formation<br />

LB387<br />

Lithology<br />

(Rank)<br />

Carbonaceous<br />

shale<br />

Table 4.2 : TOC and Rock-Eval Pyrolysis Data.<br />

OY – Onyeama<br />

OKB – Okaba<br />

(B) – Bituminous; (SB) – Sub-bituminous<br />

S 1= hydrocarbon already present in the sample which are mainly stripped at temperatures about 300 o C.<br />

S 2= hydrocarbons generated through thermal cracking <strong>of</strong> kerogen at temperatures in the range <strong>of</strong><br />

300-650 o C.<br />

Depth (m) TOC (%) Tmax ( o C) S1 (mg/g) S2 (mg/g) S3 (mg/g) PI (S1/S1+S2) HI (mg/gTOC) OI (mg/gTOC)<br />

218.5-222.5 4.6 446 0.91 4.82 1.1 0.16 105 24 5.7<br />

LB417 Coaly shale 407.4-412.5 6.7 447 0.89 3.55 0.64 0.2 52 9 4.4<br />

LB420 Coal (B) 417.0-422.0 13.7 453 1.92 17.76 2.08 0.1 129 15 19.7<br />

LB514 Coal (B) 131.7-136.6 32.6 451 12.88 77.92 2.36 0.14 239 7 90.8<br />

LB518 Coal (B) 148.0 30.4 469 3.58 25.16 2.48 0.12 82 8 28.7<br />

LB523 Coal (B) 168.8-173.7 44.5 459 10.28 57.84 3.42 0.15 129 7 68.1<br />

LB534 Coal (B) 212.1-216.2 19.9 455 4.32 23.48 1.44 0.16 118 7 27.8<br />

LB542<br />

Carbonaceous<br />

shale<br />

247.0 5.8 454 0.82 2.4 0.88 0.25 41 15 3.2<br />

LB551 Coal (B) 286.0-289.0 23.3 455 2.34 16.76 2.86 0.12 71 12 19.1<br />

Mamu<br />

Formation<br />

OY5<br />

Carbonaceous<br />

shale<br />

S 3= CO 2 that is generated from the kerogen at the same time the S 2 hydrocarbons are being generated.<br />

Tmax = Temperature <strong>of</strong> maximum generation <strong>of</strong> S 2 peak.<br />

83<br />

GP (mg/g)<br />

4.6-5.8 4.5 438 0.48 15.58 1.08 0.03 343 23 16.1<br />

OY4 Coal (SB) 6.0-6.8 66.8 432 4.18 166.48 7.2 0.02 249 10 170.7<br />

OY3 Coal (SB) 6.8-7.3 61.1 433 3.6 154 7 0.02 252 11 157.6<br />

OY2 Coal (SB) 7.6-8.4 65.1 435 4.66 168.96 6.92 0.03 259 10 173.6<br />

OY1 Coal (SB) 8.4-8.8 46.7 437 2.2 109.52 5.38 0.02 234 11 111.7<br />

OY6<br />

OKB4<br />

Carbonaceous<br />

shale<br />

Carbonaceous<br />

shale<br />

9.0-9.5 9.9 434 0.56 19.56 2.12 0.03 196 21 20.1<br />

16.5-17.6 4.1 425 0.24 4.92 7.56 0.05 118 182 5.2<br />

OKB3 Coal (SB) 18.0-18.5 59.0 421 6.8 160.72 14.32 0.04 272 24 167.5<br />

OKB2 Coal (SB) 18.6-18.9 52.8 416 19.32 148.4 21.44 0.12 281 40 167.7<br />

OKB1 Coal (SB) 18.9-19.3 27.5 425 4.58 127.2 26.96 0.03 462 97 131.8<br />

OKB7 Coal (SB) 19.3-19.6 54.7 421 3.86 116 22.88 0.03 211 41 119.9<br />

OKB8 Coal (SB) 19.6-20.0 45.7 423 4.44 121.12 21.92 0.04 265 47 125.6<br />

S1 + S2


4.2.1.2 Mamu Formation<br />

Samples from Mamu Formation have high TOC values ranging from<br />

27.53wt% to 59.00wt% with a mean value <strong>of</strong> 47.95wt% in Okaba and 46.71wt% to<br />

66.79wt% with a mean value <strong>of</strong> 59.93wt% in Onyeama <strong>coal</strong>s. The TOC <strong>of</strong><br />

interbedded carbonaceous shale is 4.14wt% in Okaba while Onyeama samples have<br />

values <strong>of</strong> 4.53wt% and 9.93 wt%.<br />

The GP values for the <strong>coal</strong> samples range from 119.9-167.7 mg/g with a mean<br />

value <strong>of</strong> 142.5mg/g in Okaba and 111.7mg/g to 173.6 mg/g with a mean value <strong>of</strong><br />

153.4mg/g in Onyeama <strong>coal</strong>s. The interbedded carbonaceous shale from Okaba has<br />

GP value <strong>of</strong> 5.2 mg/g while values <strong>of</strong> 16.1mg/g and 20.1 mg/g were recorded in<br />

Onyeama. These values are greater than 0.5 wt% and 2 mg/g respectively required for<br />

a potential source rock (Tissot and Welte, 1984; Killops and Killops 1993, 2005; Hunt,<br />

1996; Peters, et al. 2005), and thus indicate good to excellent source rock potential.<br />

4.2.2 Organic Matter Quality:<br />

The quality <strong>of</strong> the <strong>organic</strong> matter contained in the <strong>coal</strong> samples was evaluated from<br />

their Hydrogen Index (HI), plots <strong>of</strong> HI vs. OI (equivalent to van Krevelen diagram),<br />

plot <strong>of</strong> Tmax vs HI and plots <strong>of</strong> S2 vs. TOC.<br />

4.2.2.1 Awgu Formation<br />

The HI values in Awgu samples range from 41-239 mg/gTOC (av. 107.3). These<br />

low HI values indicate type III kerogen, capable <strong>of</strong> generating gas only (Peters 1986;<br />

Sachsenh<strong>of</strong>er et al. 1995).<br />

Plots <strong>of</strong> HI against OI, S2 vs. TOC and Tmax against HI for the samples are shown<br />

in Figs. 4.2, 4.3 and 4.4 respectively. The samples fall within type III evolution path<br />

on the plots <strong>of</strong> HI vs OI (Fig. 4.2) and S2 vs. TOC (Fig. 4.3), indicating gas prone<br />

(Killops and Killops 1993, 2005). The potential <strong>of</strong> the samples to generate mainly gas<br />

was further confirmed on the plot <strong>of</strong> Tmax vs HI where all the samples were plotted<br />

within the gas prone zone (Fig. 4.4).<br />

84


Fig. 4.2 : Plots <strong>of</strong> HI vs OI <strong>of</strong> Coal Samples from Benue Trough, Nigeria<br />

(After Van Krevelen et al., 1961).<br />

85


4.2.2.2 Mamu Formation<br />

The Hydrogen index (HI) values for Onyeama samples range from 196mg/gTOC to<br />

343mg/gTOC with a mean value <strong>of</strong> 255.5mg/gTOC, the highest value <strong>of</strong><br />

343mg/gTOC was recorded in the carbonaceous shale at the top <strong>of</strong> the seam. In Okaba<br />

samples the HI values range from 118mg/gTOC to 462mg/gTOC with a mean value<br />

<strong>of</strong> 268.2mg/gTOC. These values suggest that the <strong>coal</strong> is capable <strong>of</strong> generating both oil<br />

and gas (Killops and Killops 1993, 2005). The plots <strong>of</strong> HI against OI and S2 against<br />

TOC (Figs. 4.2 and 4.3), all the samples fall within type II/III kerogen fields. The<br />

Tmax vs HI plot (Fig. 4.4), further reveal that the samples plot within oil and gas zone.<br />

These features confirm their oil and gas generative potential.<br />

4.2.3 Thermal Maturity <strong>of</strong> Organic Matter<br />

The thermal maturity status <strong>of</strong> the samples was determined using pyrolysis<br />

parameters (Tmax and Production index (PI), plots <strong>of</strong> PI vs. Tmax and plots <strong>of</strong> HI vs.<br />

Tmax) and Vitrinite reflectance measurements from petrographic analyses.<br />

4.2.3.1 Awgu Formation<br />

The Tmax and PI values in Awgu samples range from 446-459 o C (av. 454) and<br />

0.10-0.29 (av. 0.15) respectively. These values indicate <strong>organic</strong> matter within the main<br />

phase <strong>of</strong> oil generation-late oil window. The Vitrinite reflectance (VR) values <strong>of</strong> Awgu<br />

samples range from 0.94-1.14 %R0 (av. 1.07). These values show that the samples are<br />

at the peak <strong>of</strong> oil generation (Killops and Killops, 2005; Peters et al., 2005). The high<br />

vitrinite reflectance recorded for Awgu Formation samples at the shallow depth is<br />

probably due to the activity <strong>of</strong> erosion, which has exposed the <strong>coal</strong> seam to the<br />

surface.<br />

86


Fig. 4.3 : Plots <strong>of</strong> S2 vs TOC <strong>of</strong> Coal Samples from Benue Trough, Nigeria<br />

(After Langford and Blanc-Valleron, 1990).<br />

87


Fig. 4.4 : Plot <strong>of</strong> Tmax vs HI. Of Coal Samples from Benue Trough, Nigeria<br />

(After modified by Akande et al., 2007).<br />

88


Vitrinite reflectance values were estimated from the plots <strong>of</strong> HI vs. Tmax for the<br />

Awgu samples (Fig. 4.5). Majority <strong>of</strong> the samples have VR <strong>of</strong> about 1.0 %R0. These<br />

values agreed with the vitrinite reflectance data obtained from the petrographic<br />

analysis. The samples also fall within the oil/condensate field (Fig. 4.7).<br />

4.2.3.2 Mamu Formation<br />

The Tmax values range from 432-438 o C (av. 435) and 416-425 o C (av. 422) for<br />

Onyeama and Okaba samples respectively. These values show that the samples are<br />

presently thermally immature and are probably within the late diagenesis stage<br />

(Killops and Killops 1993, 2005). The PI values for Okaba and Onyeama samples<br />

range from 0.03-0.12 (av. 0.05) and 0.02-0.03 (av. 0.025) respectively. The low PI<br />

values recorded for all the samples indicate that they are thermally immature.<br />

The Vitrinite reflectance values range from 0.48-0.50 %R0 (av. 0.49) and 0.57-<br />

0.60 %R0 (av. 0.58) Okaba and Onyeama samples respectively and indicate late<br />

diagenesis - early oil window. The Vitrinite reflectance values estimated from the plot<br />

<strong>of</strong> HI vs. Tmax. (Fig. 4.6) show VR


Fig. 4.5 : Plots <strong>of</strong> HI vs Tmax <strong>of</strong> Coal Samples from Awgu Formation.<br />

90


Fig. 4.6 : Plots <strong>of</strong> HI vs Tmax <strong>of</strong> Coal Samples from Mamu Formation.<br />

91


Fig. 4.7 : Plots <strong>of</strong> PI vs Tmax <strong>of</strong> Coal Samples from Benue Trough, Nigeria.<br />

92


4.3 Biomarker Geochemistry <strong>of</strong> the Nigerian Coal<br />

The source, depositional environment and thermal maturity status <strong>of</strong> the<br />

<strong>organic</strong> matter contained in the samples were determined based on the distributions<br />

and abundance <strong>of</strong> aliphatic biomarkers in the <strong>coal</strong> extracts.<br />

4.3.1 n-Alkane and Isoprenoid distributions<br />

The m/z 85 mass chromatograms showing the distribution <strong>of</strong> n-alkane and<br />

isoprenoids in the samples are shown in Figs.4.8, 4.9 and 4.10. Geochemical<br />

parameters calculated from the alkane distribution are given in Table 4.3.<br />

4.3.1.1 Awgu Formation<br />

The n-alkane distribution in Awgu samples range from C14-C35 maximizing at<br />

n-C16 or n-C18 (Fig.4.8). This pattern <strong>of</strong> distribution indicates <strong>organic</strong> matter derived<br />

from both marine and terrestrial (Peters et al., 2005). The samples plotted within the<br />

terrestrial <strong>organic</strong> matter zone on the plots <strong>of</strong> Pr/nC17 vs. Ph/nC18 in Fig. 4.11.<br />

Pr/Ph ratio calculated for the Awgu samples range from 3.04 to 11.07 (Table 4.3).<br />

All the samples have Pr/Ph ratio greater than 3.0, typical <strong>of</strong> land plant detritus<br />

deposited under aerobic (oxic) condition (Peters et al., 2005).<br />

The Carbon Preference Index (CPI) and Odd-Over-Even Predominance (OEP)<br />

values range from 0.98 to 1.12 and 0.98 to 1.06 respectively. These values reflect high<br />

maturity status <strong>of</strong> the samples (Peters et al., 2005). Also, the plots <strong>of</strong> CPI against OEP<br />

show that the samples are thermally mature (Fig. 4.12).<br />

93


Fig. 4.8 : m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Awgu samples (BH 94)<br />

showing the distribution <strong>of</strong> n-Alkanes.<br />

94


Fig. 4.8 (contd.): m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Awgu samples<br />

(BH 120) showing the distribution <strong>of</strong> n-Alkanes.<br />

95


Fig. 4.9 : m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu Formation samples<br />

(Okaba) showing the distribution <strong>of</strong> n-Alkanes.<br />

96


Fig. 4.9 (contd.): m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu Formation<br />

samples (Okaba) showing the distribution <strong>of</strong> n-Alkanes.<br />

97


Fig. 4.10 : m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu Formation samples<br />

(Onyeama) showing the distribution <strong>of</strong> n-Alkanes.<br />

98


Fig. 4.10 (contd.): m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu Formation<br />

samples (Onyeama) showing distribution <strong>of</strong> n-Alkanes.<br />

99


Table 4.3 : n-Alkanes and Isoprenoids Parameters.<br />

Samples Depth (m)<br />

Awgu Formation<br />

LB387 218.5-222.5<br />

Pr- Pristane, Ph- Phytane<br />

CPI = ½[(C25 + C27 +C29 + C31 + C33/C24 +C26+C28+C30+C32+C34) + (C25 + C27 +C29 + C31 +<br />

C33/C26+C28+C30+C32+C34)], CPI (1) = 2(C23+C25+C27+C29)/[C22+2(C24+C26+C28) +C30]<br />

OEP (1) = C21+6C23+C25/ 4(C22+C24), OEP (2) = C25+6C27+C29/ 4(C26+C28)<br />

OY – Onyeama<br />

OKB – Okaba<br />

Lithology<br />

(Rank)<br />

Carbonaceous<br />

shale<br />

(B) – Bituminous; (SB) – Sub-bituminous<br />

Pr/Ph Pr/nC17 Ph/nC18 CPI CPI (1) OEP (1) OEP (2) Crange Cmax<br />

6.14 1.81 0.23 1.12 1.08 1.03 1.06 C13-36 C23<br />

LB417 407.4-412.5 Coaly shale 3.16 0.50 0.13 1.02 1.01 0.98 1.00 C14-37 C18, C20<br />

LB420 417.0-422.0 Coal (B) 5.49 0.30 0.03 1.07 1.02 0.99 1.00 C16-32 C18<br />

LB514 131.7-136.6 Coal (B) 3.52 0.24 0.06 1.05 1.02 0.99 1.02 C13-35 C19<br />

LB518 148.0 Coal (B) 5.25 0.68 0.11 1.09 1.04 0.99 1.06 C13-35 C20<br />

LB523 168.8-173.7 Coal (B) 6.52 0.54 0.08 1.10 1.05 0.99 1.05 C13-35 C16<br />

LB534 212.1-216.2 Coal (B) 4.60 0.28 0.06 1.07 1.04 1.01 1.03 C13-35 C16, C18, C20<br />

LB542 247.0<br />

Carbonaceous<br />

shale<br />

11.07 1.69 0.09 1.04 0.99 0.98 0.99 C15-33 C27<br />

LB551 286-289 Coal (SB) 3.04 0.34 0.08 1.06 1.03 0.99 1.03 C13-35 C18, C20, C21, C22<br />

Mamu Formation<br />

OY5 4.6-5.8<br />

Carbonaceous<br />

shale<br />

9.33 10.89 0.83 2.11 1.99 1.22 1.86 C14-35 C29<br />

OY4 6.0-6.8 Coal (SB) 10.48 52.19 2.83 2.13 1.87 0.65 1.89 C13-35 C29<br />

OY3 6.8-7.3 Coal (SB) 11.55 40.09 2.23 1.97 1.91 0.73 1.93 C13-35 C29<br />

OY2 7.6-8.4 Coal (SB) 9.66 27.31 3.06 2.22 1.85 0.86 1.80 C13-35 C29<br />

OY1 8.4-8.8 Coal (SB) 9.87 40.33 1.63 2.17 1.96 1.16 1.88 C14-35 C27, C29<br />

OY6 9.0-9.5<br />

OKB4 16.5-17.6<br />

Carbonaceous<br />

shale<br />

Carbonaceous<br />

shale<br />

12.47 14.99 1.05 2.32 2.12 1.15 2.27 C14-35 C29<br />

2.11 3.40 1.08 3.76 3.00 1.30 3.08 C14-37 C29<br />

OKB3 18.0-18.5 Coal (SB) 1.73 2.94 0.72 5.18 5.82 1.29 5.16 C15-35 C29<br />

OKB2 18.6-18.9 Coal (SB) 3.76 5.42 1.27 7.38 8.72 1.29 7.06 C12-35 C29<br />

OKB1 18.9-19.3 Coal (SB) 4.08 4.51 0.74 6.51 7.20 1.28 6.28 C13-35 C29<br />

OKB7 19.3-19.6 Coal (SB) 3.30 2.57 0.70 5.93 6.16 1.38 6.13 C12-35 C29<br />

OKB8 19.6-20.0 Coal (SB) 2.11 2.53 0.86 6.02 6.48 1.16 6.02 C13-35 C29<br />

100


Fig. 4.11: Plots <strong>of</strong> Pr/nC17 against Ph/nC18 <strong>of</strong> Awgu and Mamu Formation samples.<br />

101


Fig. 4.12 : Plots <strong>of</strong> CPI against OEP <strong>of</strong> Awgu and Mamu samples.<br />

102


4.3.1.2 Mamu Formation<br />

The n-alkanes distribution in Mamu samples range from C14- C35, maximizing at<br />

C27 or C29 (Fig. 4.9, 4.10). This pattern indicates <strong>organic</strong> matter derived mainly from<br />

terrestrial <strong>organic</strong> matter. High proportions <strong>of</strong> long chain C27-C31 members relative to<br />

the total n-alkanes especially are typical <strong>of</strong> terrestrial higher plants (Eglinton and<br />

Hamilton, 1963; Caldicott and Eglinton, 1973; Tissot and Welte, 1984; Barthlot, et al.,<br />

1998; Miranda, et al., 1999). The Pr/Ph ratios range from 9.33 to 12.47 and 1.73 to<br />

4.08 for Onyeama and Okaba samples respectively (Table 4.3). All the samples,<br />

except some few from Okaba have Pr/Ph ratio greater than 3.0, typical <strong>of</strong> land plant<br />

detritus deposited under aerobic (oxic) condition (Killops and Killops, 2005; Peters et<br />

al., 2005; Tuo et al., 2007). The few samples having low Pr/Ph ratios can be<br />

interpreted as being deposited under sub-oxic to oxic settings (Killops and Killops,<br />

2005; Peters et al., 2005; Tuo et al., 2007).<br />

The Carbon Preference Index (CPI) and Odd-over-even predominance (OEP)<br />

values in Onyeama samples range from 1.85 to 2.32 and 0.65 to 2.27 respectively.<br />

The for Okaba samples range from 3.00 to 8.72 and 1.16 to 7.06 respectively. The<br />

high CPI (>>1) and OEP values observed are characteristic <strong>of</strong> low rank <strong>coal</strong> i.e. sub-<br />

bituminous (Bray and Evan, 1961; Scalan and Smith, 1970; Tissot and Welte, 1984;<br />

Bechtel et.al., 2004; Sabel et al., 2005; Stefanova et al., 2005). The plot <strong>of</strong> CPI<br />

against OEP show that these samples are immature to low mature (Fig. 4.12).<br />

4.3.2 Fatty acids and alkanones<br />

The distributions <strong>of</strong> fatty acids in the polar fraction have been successfully<br />

used to differentiate the biological source <strong>of</strong> geological materials (Duan et al., 1997).<br />

The m/z 58 and m/z 74 mass chromatograms showing the distributions <strong>of</strong> the saturated<br />

n-fatty acids and alkan-2-ones in the <strong>coal</strong> extracts are shown in Figs.<br />

4.13,4.14,4.15,4.16,4.17 and 4.18 respectively. Parameters calculated from the fatty<br />

acids and alkanones distribution in the <strong>coal</strong> extracts are listed in Table 4.4.<br />

103


4.3.2.1 Awgu Formation<br />

Awgu samples have n-fatty acids ranging from C14 to C30, maximizing at nC16 or<br />

nC18 (Fig. 4.13). The short chain/long chain saturated fatty acid (ATRFA) ratios for the<br />

samples which range from 0.97-1.00, indicate both terrestrial and marine <strong>organic</strong><br />

matter derived material (Wilkes et al., 1999). Abundance <strong>of</strong> short chain saturated<br />

n-fatty acids (C20) in the samples can be attributed to curticular waxes <strong>of</strong> higher plants (Cranwell,<br />

1974).<br />

The carbon preference index (CPIFA) <strong>of</strong> the long chain n-fatty acids (C24-C30)<br />

range between 1.98 and 3.29 (Table 4.4), indicating a strong even over odd-<br />

predominance. These values inferred high maturity status for the samples (Wilkes et<br />

al., 1999).<br />

The distribution <strong>of</strong> straight chain n-alkan-2-ones range from nC14-nC33,<br />

maximizing at nC17 (Fig. 4.16). Similar distribution has previously been observed in<br />

stalagmites (Xie et al., 2003; Bai et al., 2006). However, some <strong>of</strong> the samples<br />

maximize at nC23 or nC25 (Fig. 4.16), an indication <strong>of</strong> contribution from higher plants,<br />

microalgae and phytoplankton <strong>organic</strong> matter inputs (Hernandez et al., 2001;<br />

Gonzalez-Vila et al., 2003).<br />

104


Fig. 4.13: m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in<br />

Awgu samples (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

105


Fig 4.13 (contd.): m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty<br />

acids in Awgu samples (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

106


Fig. 4.14: m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in<br />

Mamu samples (Okaba) (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

107


Fig. 4.15: m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in Mamu<br />

samples (Onyeama) (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

108


Fig. 4.15 (contd.): m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in<br />

Mamu samples (Onyeama) (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids). (Numbers<br />

refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

109


Fig. 4.16: m/z 58 mass chromatograms showing the distributions <strong>of</strong> alkan-2-ones in Awgu<br />

samples (Numbers refer to carbon chain lengths <strong>of</strong> alkan-2-ones).<br />

110


Fig. 4.17: m/z 58 mass chromatograms showing the distributions <strong>of</strong> alkan-2-ones in Mamu<br />

samples (Okaba) (Numbers refer to carbon chain lengths <strong>of</strong> alkan-2-ones).<br />

111


Fig. 4.18: m/z 58 mass chromatograms showing the distributions <strong>of</strong> alkan-2-ones in Mamu<br />

samples (Onyeama) (Numbers refer to carbon chain lengths <strong>of</strong> alkan-2-ones).<br />

112


Table 4.4: Parameters calculated from n-Fatty acids and alkanones composition <strong>of</strong> Nigerian Coal.<br />

Samples Depth (m) Lithology (Rank) ATR FA CPI LFA Pr-2-one/C17 CPI (alkanone)<br />

Awgu<br />

Formation<br />

LB387 218.5-222.5 Carbonaceous shale 0.99 ND 0.58 0.84<br />

LB417 407.4-412.5 Coaly shale 0.99 ND 0.80 0.85<br />

LB420 417.0-422.0 Coal (B) 1.00 ND 1.09 1.14<br />

LB514 131.7-136.6 Coal (B) 1.00 ND 0.51 0.89<br />

LB518 148.0 Coal (B) 0.99 2.46 0.60 0.76<br />

LB523 168.8-173.7 Coal (B) 0.99 1.98 0.39 1.01<br />

LB534 212.1-216.2 Coal (B) 0.99 3.29 1.59 0.76<br />

LB542 247.0 Carbonaceous shale 0.97 3.20 0.86 0.93<br />

LB551 286.0-289.0 Coal (B) 0.98 ND 0.82 1.20<br />

Mamu<br />

Formation<br />

OY5 4.6-5.8 Carbonaceous shale 0.95 1.47 4.39 1.27<br />

OY4 6.0-6.8 Coal (SB) 0.91 1.37 10.54 1.45<br />

OY3 6.8-7.3 Coal (SB) 0.92 1.61 8.69 1.51<br />

OY2 7.6-8.4 Coal (SB) 0.92 2.07 12.18 1.69<br />

OY1 8.4-8.8 Coal (SB) 0.96 1.25 7.45 1.30<br />

OY6 9.0-9.5 Carbonaceous shale 0.94 1.61 4.10 1.60<br />

OKB4 16.5-17.6 Carbonaceous shale ND ND ND ND<br />

OKB3 18.0-18.5 Coal (SB) 0.85 2.78 8.68 2.66<br />

OKB2 18.6-18.9 Coal (SB) 0.93 1.37 7.69 2.25<br />

OKB1 18.9-19.3 Coal (SB) 0.94 1.95 12.69 2.59<br />

OKB7 19.3-19.6 Coal (SB) 0.96 1.44 13.71 2.49<br />

OKB8 19.6-20.0 Coal (SB) 0.91 2.51 21.43 2.10<br />

ATR FA = Short chain/long chain saturated fatty acid<br />

CPI LFA= Carbon Preference Index (Longchainfatty acids)<br />

CPI (alkanones)= Carbon Preference Index (alkan-2-ones).<br />

ATR FA = C 14 + C 16 + C 18 / C 14 + C 16 + C 18 + C 26 + C 28 + C 30<br />

CPI LFA=½{(C24+C26+C28+C30/C21+C23+C25+C27)+(C24+C26+C28+C30/C23+C25+C27+C29)}<br />

CPI(alkanones)=1/2{(C25+C27+C29+C31+C33)/(C22+C24+C26+C28+C30+C32)<br />

+(C25+C27+C29+C31+C33)/ C24+C26+C28+C30+C32+34}}<br />

OY – Onyeama<br />

OKB – Okaba<br />

(B) – Bituminous<br />

(SB) – Sub-bituminous<br />

ND – Not determined<br />

113


4.3.2.2 Mamu Formation<br />

Mamu samples have saturated n-fatty acids ranging from C8 to C32,<br />

maximizing at nC16 or nC18 (Fig. 4.14 and 4.15). These distributions reflect <strong>organic</strong><br />

matter from both marine and terrestrial materials (Volkman et al., 1998). However the<br />

dominance <strong>of</strong> short chain (nC22) in the samples can<br />

be attributed to the contribution <strong>of</strong> higher plants to the <strong>organic</strong> matter (Cranwell,<br />

1974). The ATRFA ratios range from 0.85 to 0.96. These values indicate <strong>organic</strong> matter<br />

derived from mixed origin (Wilkes et al., 1999). The carbon preference index (CPILFA)<br />

for the long chain saturated n-fatty acids range between 1.25 and 2.78, indicating a<br />

slight even over odd predominance (Table 4.4). These values indicate low maturity<br />

(Wilkes et al., 1999).<br />

The n-alkan-2-ones range from nC12 to nC33, maximizing at nC17 or nC29<br />

(Fig. 4.17 and 4.18). These distributions reflect higher plants and algae inputs to the<br />

<strong>organic</strong> matter (Bai et al., 2006). The CPI values range from 1.27 to 1.69 and 2.10 to<br />

2.66 in Onyeama and Okaba samples respectively (Table 4.4). These values indicate<br />

low maturity status for all the samples (Tuo et al., 2007).<br />

4.3.3 Tricyclic and C24 tetracyclic terpanes<br />

The m/z 191 showing the distributions <strong>of</strong> tricyclic and tetracyclic terpanes in the<br />

samples are shown in Fig. 4.19,4.20 and 4.21. Peak identities are listed in Table 4.6.<br />

4.3.3.1 Awgu Formation<br />

A series <strong>of</strong> tricyclic terpanes ranging from C19 to C29 are observed in Awgu<br />

samples (Figs. 4.19). Higher percentages <strong>of</strong> C19-C21 compared to C23 tricyclic terpanes<br />

indicate <strong>organic</strong> matter derived from terrestrial origin (Ozcelik and Altunsoy, 2005).<br />

The C24 tetracyclic terpane is present in appreciable amounts in all the samples.<br />

The C24tetra/C26tri(R+S) ratios range between 0.96 and 3.45, probably reflecting<br />

terrigenous <strong>organic</strong> matter input (Philp and Gilbert, 1986).<br />

114


Various ratios <strong>of</strong> tricyclic terpanes have been used to distinguish marine carbonate,<br />

lacustrine, paralic, <strong>coal</strong>/resin and evaporitic source depositional environments<br />

(De Grande et al., 1993; Tuo et al., 1999; Yangming et al., 2005; Peters et al., 2005).<br />

C22/C21 tricyclic terpane ratio in the samples range from 0.16 to 0.42, suggesting<br />

<strong>organic</strong> matter deposited in lacustrine-fluvial/deltaic depositional environment (Peters<br />

et al., 2005). C24tetra/C30hopane ratio has also been used to assess depositional<br />

environment <strong>of</strong> source rock (Peters et al., 2005). C24tetra/C30hopane ratios in the<br />

sample range between 0.11 and 0.43. These values also indicate <strong>organic</strong> matter<br />

deposited in lacustrine-fluvial/deltaic depositional environment (Peters et al., 2005).<br />

115


Fig. 4.19 :m/z 191 showing the distribution <strong>of</strong> tricyclic and tetracyclic terpane in<br />

Awgu samples.<br />

116


Fig. 4.20: m/z 191 showing the distribution <strong>of</strong> tricyclic and tetracyclic terpane in<br />

Mamu Formation samples (Okaba).<br />

117


Fig. 4.21 : m/z 191 showing the distribution <strong>of</strong> tricyclic and tetracyclic terpane in<br />

Mamu Formation samples (Onyeama).<br />

118


Sample<br />

Table 4.5 : Tri- and tetracyclic terpanes source and depositional environment parameters.<br />

Awgu Formation<br />

Depth<br />

(m)<br />

Lithology<br />

(Rank)<br />

C24tetra/<br />

C30hopane<br />

119<br />

C24tetra/<br />

C26(R+S)tri<br />

C22/C21 triterpane<br />

%C19-C21<br />

triterpane<br />

LB387 218.5-222.5 Carbonaceous shale 0.10 2.50 0.24 83.70 16.30<br />

LB417 407.4-412.5 Coaly shale 0.18 2.47 0.24 77.30 22.70<br />

LB420 417.0-422.0 Coal (B) 0.43 2.84 0.23 82.50 17.50<br />

LB514 131.7-136.6 Coal (B) 0.18 2.55 0.17 82.40 17.60<br />

LB518 148.0 Coal (B) 0.24 2.14 0.26 86.20 13.80<br />

LB523 168.8-173.7 Coal (B) 0.24 3.45 0.17 84.70 15.30<br />

LB534 212.1-216.2 Coal (B) 0.21 2.29 0.42 93.60 6.40<br />

LB542 247 Carbonaceous shale 0.32 0.96 0.16 82.10 17.90<br />

LB551 286.0-289.0 Coal (B) 0.18 1.33 0.27 72.70 27.30<br />

Mamu Formation<br />

OY5 4.6-5.8 Carbonaceous shale 0.08 2.40 0.65 83.00 17.00<br />

OY4 6.0-6.8 Coal (SB) 0.09 2.86 0.46 90.00 10.00<br />

OY3 6.8-7.3 Coal (SB) 0.05 2.11 0.61 72.30 27.70<br />

OY2 7.6-8.4 Coal (SB) 0.08 2.21 0.64 87.50 12.50<br />

OY1 8.4-8.8 Coal (SB) 0.09 2.70 0.55 78.50 21.50<br />

OY6 9.0-9.5 Carbonaceous shale 0.06 0.98 0.61 85.80 14.20<br />

OKB4 16.5-17.6 Carbonaceous shale 0.18 1.66 0.79 73.90 26.10<br />

OKB3 18.0-18.5 Coal (SB) 0.09 2.04 0.66 53.40 46.60<br />

OKB2 18.6-18.9 Coal (SB) 0.08 1.41 ND ND ND<br />

OKB1 18.9-19.3 Coal (SB) 0.06 1.94 0.64 70.40 29.60<br />

OKB7 19.3-19.6 Coal (SB) 0.10 2.53 0.84 81.00 19.00<br />

OKB8 19.6-20.0 Coal (SB) 0.15 2.28 0.72 79.50 20.50<br />

C24tetra/C30hopane = C24tetracyclic terpane/C30hopane<br />

C24tetra/C26(R+S)tri= C24tetracyclic terpane/C26(R+S)tricyclic terpane<br />

OY – Onyeama<br />

OKB – Okaba<br />

(B) – Bituminous<br />

(SB) – Sub-bituminous<br />

ND – Not determined<br />

%C23<br />

triterpane


4.3.3.2 Mamu Formation<br />

Mamu samples are dominated by C19-C21 tricyclic terpanes (Table 4.5). These<br />

distributions indicate <strong>organic</strong> matter derived from terrestrial matter (Ozcelik and<br />

Altunsoy, 2005). The C22/C21 tricyclic terpane ratios range from 0.46 to 0.65 and 0.64<br />

to 0.84 in Onyeama and Okaba samples respectively (Table 4.5). The observed<br />

C22/C21 tricyclic ratios indicate fluvial/deltaic and lacustrine-fluvial/deltaic<br />

depositional environment for Onyeama and Okaba samples respectively (Peters et al.,<br />

2005).<br />

The C24tetra/C26tri(R+S) ratios range between 0.98 and 2.86, probably reflecting<br />

terrigenous <strong>organic</strong> matter input (Philp and Gilbert, 1986). The C24tetra/C30hopane<br />

ratios also range from 0.05 to 0.09 and 0.06 to 0.18 in Onyeama and Okaba samples<br />

respectively. These values also reflect fluvial/deltaic depositional environment and<br />

lacustrine-fluvial/deltaic for Onyeama and Okaba samples respectively (Peters et al.,<br />

2005).<br />

4.3.4 Hopanes and homohopanes<br />

The m/z 191 mass chromatograms showing the distribution <strong>of</strong> pentacyclic<br />

triterpanes in the samples are shown in Figs. 4.22,4.23 and 4.24 and the peak<br />

identities are given in Table 4.6.<br />

4.3.4.1 Awgu Formation<br />

C29 and C30αβ-hopane occur in appreciable amount in all the Awgu <strong>coal</strong>y shale<br />

samples (Fig. 4.22), indicating significant contribution <strong>of</strong> prokaryotic organisms<br />

(i.e. bacteria, cyanobacteria and blue algae) to the source <strong>organic</strong> matter.<br />

The sterane/hopane ratio is <strong>of</strong>ten used as a measure <strong>of</strong> relative inputs <strong>of</strong> eukaryotic<br />

versus prokaryotic debris (Peters and Moldowan, 1993). The sterane/hopane ratio<br />

values range from 0.04-0.51(Table 4.7).<br />

120


The ratio values (


Fig. 4.22: m/z 191 Mass chromatogram showing the distribution <strong>of</strong> hopanes in<br />

Awgu samples.<br />

122


Fig. 4.23 : m/z 191 Mass chromatogram showing the distribution <strong>of</strong> hopanes and<br />

benzohopanes in Mamu sample (Okaba).<br />

123


Fig. 4.24 : m/z 191 mass chromatogram showing the distribution <strong>of</strong> hopanes and<br />

benzohopanes in Mamu samples (Onyeama).<br />

124


Table 4.6 : Peak identities on m/z 191 mass chromatograms.<br />

Peak Compound<br />

C19t C19 tricyclic terpane<br />

C20t C20 tricyclic terpane<br />

C21t C21 tricyclic terpane<br />

C22t C22 tricyclic terpane<br />

C23t C23 tricyclic terpane<br />

C24t C24 tricyclic terpane<br />

C25tS C25 22(S)-tricyclic terpane<br />

C25tR C25 22(R)-tricyclic terpane<br />

C24tetra Tetracyclic hopane (secohopane)<br />

C26tS C26 22(S)-tricyclic terpane<br />

C26tR C26 22(R)-tricyclic terpane<br />

C28tS C28 22(S)-tricyclic terpane<br />

C28tR C28 22(R)-tricyclic terpane<br />

1 C27 18α(H)-22,29,30-trisnorneohopane (Ts)<br />

2 C27 17α(H)-22,29,30-trisnorhopane(Tm)<br />

3 C27 17β(H)-22,29,30-trisnorhopane<br />

4 C29 17α(H), 21β(H)-30-norhopane<br />

4* Oleanene isomers; olean-18-ene, olean-13 (18)-ene, olean-12-ene<br />

5 C29 17β(H), 21α(H)-normoretane<br />

6 C30 17α(H), 21β(H)-hopane<br />

7 C30 17α(H), 21α(H)-30-norhopane<br />

8 C30 17β(H), 21α(H)-moretane<br />

9 C31 17α(H), 21β(H)-30-homohopane (22S)<br />

10 C31 17α(H), 21β(H)-30-homohopane (22R)<br />

10* C30 17β(H), 21β(H)-hopane<br />

11 C32 17α(H), 21β(H)-30,31-bishomohopane (22S)<br />

12 C32 17α(H), 21β(H)-30,31-bishomohopane (22R)<br />

11* C31 17β(H), 21β(H)-hopane<br />

13 C33 17α(H), 21β(H)-30,31,32-trishomohopane (22S)<br />

14 C33 17α(H), 21β(H)-30,31,32-trishomohopane (22R)<br />

125


Table 4.6 (contd.) : Peak identities on m/z 191 mass chromatograms.<br />

Peak Compound<br />

15<br />

C34 17α(H), 21β(H)-30,31,32,33-tetrakishomohopane (22S)<br />

16 C34 17α(H), 21β(H)-30,31,32,33-tetrakishomohopane (22R)<br />

17 C35 17α(H), 21β(H)-30,31,32,33,34-pentakishomohopane (22S)<br />

18 C35 17α(H), 21β(H)-30,31,32,33,34-pentakishomohopane (22R)<br />

Sp30 18α(H)-28-noroleanane<br />

O 18α(H)-oleanane + 18β(H)-oleanane<br />

126


Samples<br />

Awgu Formation<br />

Depth<br />

(m)<br />

Table 4.7: Source and depositional environment parameters computed from the hopane and<br />

sterane distributions in the <strong>coal</strong>s.<br />

Lithology<br />

(Rank)<br />

%C27<br />

sterane<br />

%C28<br />

sterane<br />

%C29<br />

sterane<br />

%C27<br />

diast.<br />

LB387 218.5-222.5 Carbonaceous shale 25.43 20.04 54.53 27.46 54.50 18.14 0.47 55.74 0.11 0.06 0.43 0.06<br />

LB417 407.4-412.5 Coaly shale 18.13 28.27 53.60 21.58 32.14 46.28 0.34 68.03 0.11 0.06 0.25 0.04<br />

LB420 417.0-422.0 Coal (B) 24.04 29.02 46.94 30.53 31.25 38.22 0.51 67.35 0.51 0.07 0.15 0.05<br />

LB514 131.7-136.6 Coal (B) 17.47 29.66 52.87 15.05 33.05 51.90 0.33 55.13 0.27 0.03 0.34 0.04<br />

LB518 148.0 Coal (B) 23.19 23.28 53.53 10.79 32.29 56.93 0.43 73.80 0.32 0.03 0.37 0.04<br />

LB523 168.8-173.7 Coal (B) 16.14 26.10 57.77 19.42 42.67 37.91 0.28 46.81 0.43 0.03 0.35 0.05<br />

LB534 212.1-216.2 Coal (B) 18.71 36.88 44.41 27.97 45.89 26.14 0.42 70.97 0.04 0.06 0.33 0.06<br />

LB542 247.0 Carbonaceous shale 31.89 31.20 36.91 25.86 42.26 31.89 0.86 62.81 0.11 0.26 0.58 0.10<br />

LB551 286.0-289.0 Coal (B) 17.58 32.90 49.52 24.41 29.04 46.55 0.36 67.92 0.13 0.11 0.89 0.10<br />

Mamu Formation<br />

OY5 4.6-5.8 Carbonaceous shale 24.02 20.18 55.8 28.40 23.37 48.27 0.43 89.00 0.07 0.05 0.23 0.04<br />

OY4 6.0-6.8 Coal (SB) 11.83 33.64 54.53 23.05 28.61 48.34 0.22 84.51 0.06 0.05 0.25 0.03<br />

OY3 6.8-7.3 Coal (SB) 10.73 29.71 59.56 25.44 24.33 50.23 0.18 66.56 0.29 0.02 0.32 0.02<br />

OY2 7.6-8.4 Coal (SB) 16.51 26.46 57.03 20.14 30.31 49.56 0.29 66.30 0.04 0.05 0.29 0.03<br />

OY1 8.4-8.8 Coal (SB) 12.97 20.08 66.95 17.00 24.06 58.94 0.20 87.23 0.06 0.01 0.28 0.02<br />

OY6 9.0-9.5 Carbonaceous shale 20.00 23.38 56.62 17.31 25.94 56.75 0.35 71.37 0.35 0.03 0.59 0.04<br />

OKB4 16.5-17.6 Carbonaceous shale 20.98 43.2 35.82 13.91 29.66 56.43 0.59 50.32 0.16 0.09 0.62 0.13<br />

OKB3 18.0-18.5 Coal (SB) 19.06 41.28 39.66 11.60 35.63 52.77 0.48 50.69 0.20 0.32 0.81 0.11<br />

OKB2 18.6-18.9 Coal (SB) 10.59 27.73 61.68 21.89 42.34 35.78 0.17 63.72 0.10 0.06 0.76 0.09<br />

OKB1 18.9-19.3 Coal (SB) 11.57 29.61 58.83 26.45 32.19 41.36 0.20 51.76 0.08 0.01 0.59 0.03<br />

OKB7 19.3-19.6 Coal (SB) 17.06 39.20 43.79 32.76 31.00 36.24 0.40 73.84 0.08 0.02 0.92 0.11<br />

OKB8 19.6-20.0 Coal (SB) 10.78 34.31 54.91 19.53 42.62 37.85 0.20 46.77 0.12 0.47 0.85 0.12<br />

Sterane/Hopane=C27+C28+C29steranes/[(C29+C30)αβhopane + (C31+C32+C33)αβ(R+S)homohopane]<br />

%C28<br />

diast.<br />

127<br />

%C29<br />

diast.<br />

C27/C29<br />

sterane<br />

C35/C30 = C35αβ(R+S) homohopane/ C30αβ hopane + C30βα moretane<br />

Homohopane ratio ,C35/C34 αβS = C35αβS/C34αβS homohopane<br />

Homohopane index = C35/ C31+C32+C33+C34+C35) αβ(R+S) homohopane<br />

OY – Onyeama<br />

OKB – Okaba<br />

(B) – Bituminous<br />

(SB) – Sub-bituminous<br />

%Diast./<br />

sterane<br />

Sterane/<br />

hopane<br />

C35/C30<br />

hopane<br />

C35/C34<br />

αβS hopane<br />

Homohopane/<br />

index


Samples Depth (m)<br />

Awgu<br />

Formation<br />

LB387 218.5-222.5<br />

Table 4.8 : Maturity parameters computed from the hopane and sterane distributions in the <strong>coal</strong>s.<br />

Lithology<br />

(Rank)<br />

Carbonaceous<br />

shale<br />

Mor/Hop = Moretane/Hopane (C30)<br />

Hop/Hop + Mor = Hopane/Hopane + Moretane (C30)<br />

C32HH = C32homohopane<br />

OY – Onyeama<br />

OKB – Okaba<br />

(B) – Bituminous<br />

(SB) – Sub-bituminous<br />

Mor/Hop Hop/Hop+ Mor Ts/Ts+Tm<br />

128<br />

22S/22S+22RC32 HH<br />

20S/20S+20R C29 steranes ββ/ββ+αα C29 sterane<br />

0.12 0.89 0.61 0.60 0.58 0.47<br />

LB417 407.4-412.5 Coaly shale 0.08 0.92 0.46 0.62 0.43 0.55<br />

LB420 417.0-422.0 Coal (B) 0.11 0.9 0.64 0.61 0.44 0.51<br />

LB514 131.7-136.6 Coal (B) 0.10 0.91 0.46 0.61 0.44 0.48<br />

LB518 148.0 Coal (B) 0.09 0.92 0.53 0.60 0.48 0.49<br />

LB523 168.8-173.7 Coal (B) 0.09 0.92 0.34 0.58 0.45 0.54<br />

LB534 212.1-216.2 Coal (B) 0.06 0.94 0.59 0.58 0.44 0.42<br />

LB542 247.0<br />

Carbonaceous<br />

shale<br />

0.14 0.88 0.49 0.53 0.45 0.48<br />

LB551 286.0-289.0 Coal (B) 0.08 0.93 0.66 0.61 0.44 0.51<br />

Mamu<br />

Formation<br />

OY5 4.6-5.8<br />

Carbonaceous<br />

shale<br />

0.46 0.69 0.04 0.54 0.04 0.25<br />

OY4 6.0-6.8 Coal (SB) 0.56 0.64 0.03 0.59 0.15 0.19<br />

OY3 6.8-7.3 Coal (SB) 0.63 0.61 0.04 0.55 0.17 0.20<br />

OY2 7.6-8.4 Coal (SB) 0.54 0.65 0.03 0.54 0.11 0.24<br />

OY1 8.4-8.8 Coal (SB) 0.61 0.62 0.02 0.55 0.19 0.16<br />

OY6 9.0-9.5<br />

OKB4 16.5-17.6<br />

Carbonaceous<br />

shale<br />

Carbonaceous<br />

shale<br />

0.64 0.61 0.05 0.58 0.19 0.25<br />

0.84 0.54 0.13 0.48 0.19 0.30<br />

OKB3 18.0-18.5 Coal (SB) 0.85 0.54 0.16 0.43 0.10 0.33<br />

OKB2 18.6-18.9 Coal (SB) 0.62 0.62 0.24 0.31 0.18 0.21<br />

OKB1 18.9-19.3 Coal (SB) 0.59 0.62 0.11 0.37 0.19 0.50<br />

OKB7 19.3-19.6 Coal (SB) 0.57 0.64 0.14 0.33 0.39 0.56<br />

OKB8 19.6-20.0 Coal (SB) 0.93 0.52 0.24 0.16 0.22 0.45


4.3.4.2 Mamu Formation<br />

The C27 to C35 hopanes are detected in all the Mamu samples but C28 was not<br />

detected (Fig. 4.23 and 4.24). The most prominent hopane in Onyeama samples are<br />

C29αβ-norhopane and C30αβ-hopane while C29αβ norhopane is predominant in Okaba<br />

samples (Fig 4.23 and 4.24). In the Okaba samples, abundance <strong>of</strong> C29αβ-hopane in all<br />

the samples reflects major contribution from terrestrial <strong>organic</strong> matter; however,<br />

contribution from prokaryotic organisms is not excluded while abundant C30αβ-<br />

hopane with notable presence <strong>of</strong> C29αβ in Onyeama samples reflects significant<br />

contribution from prokaryotic organisms as well as vitrinitic (terrestrial) <strong>organic</strong><br />

matter. The unusual high abundance <strong>of</strong> 22R compared to 22S in the C31-17α (H),<br />

21β(H) homohopane is evident in all Okaba samples. This is likely due to co-elution<br />

<strong>of</strong> gammacerane (Peter and Moldowan, 1993; Kagya, 1996; Farrimond et al., 1998;<br />

Peters et al., 2005).<br />

Benzohopanes with different distributions were found in Onyeama and Okaba <strong>coal</strong><br />

samples (Fig. 4.23 and 4.24). There is no previous record <strong>of</strong> presence <strong>of</strong><br />

benzohopanes in Nigerian <strong>coal</strong> and <strong>coal</strong>y <strong>organic</strong> matter. The C32-C35 benzohopanes<br />

were detected in Onyeama samples while C32-C33 benzohopanes were detected in<br />

Okaba samples. Benzohopanes are thought to be secondary transformation products <strong>of</strong><br />

C35 bacteriohopanepolyol derivatives (Grice et al., 1998; Peters et al., 2005; Killops<br />

and Killops, 2005; Bechtel et al., 2007a).<br />

129


Fig. 4.25: Mass frangmentogram and spectra <strong>of</strong> Olean-18-ene in Okaba Mine<br />

samples.<br />

130


Fig. 4.26: Mass frangmentogram and spectra <strong>of</strong> Olean-13(18)-ene in Okaba Mine<br />

samples.<br />

131


Fig. 4.27: Mass frangmentogram and spectra <strong>of</strong> Olean-12-ene in Okaba Mine<br />

samples.<br />

132


Three isomers <strong>of</strong> oleanenes; olean-18-ene, olean-13 (18)-ene and olean-12-ene were<br />

identified in Okaba samples (Fig. 4.25,4.26 and 4.27). Similar to the benzohopanes,<br />

this is the first time oleanene isomers are being identified in Nigerian <strong>coal</strong> and <strong>coal</strong>y<br />

<strong>organic</strong> matter. These three oleanene isomers are products <strong>of</strong> late diagenesis from<br />

taraxerol and β-amyrin, which are biomarkers for angiosperm (Ten Haven and<br />

Rullkötter, 1988; Ekweozor and Telnǽs, 1990; Rullkötter et al., 1994; Curiale, 1995).<br />

They have also been found useful as indicators <strong>of</strong> thermal immaturity (Eneogwe et al.,<br />

2002).<br />

In Onyeama sample, C35/C30 hopane ratio range from 0.01 to 0.05 while Okaba<br />

samples have values ranging from 0.01 to 0.47. These values indicate fluvial/deltaic<br />

and lacustrine-fluvial/deltaic depositional environments for Onyeama and Okaba<br />

samples respectively.<br />

The homohopane index and homohopane ratio for the samples range from 0.02-<br />

0.13 and 0.23-0.92 respectively (Table 4.7). These values indicate oxic depositional<br />

environment for Onyeama samples and suboxic-oxic depositional environments for<br />

Okaba samples (Peters and Moldowan, 1991; Hanson et al., 2001; Killops and Killops,<br />

2005; Peters et al., 2005; Yangming et al., 2005). There is presence <strong>of</strong> gammacerane<br />

in Okaba samples (Fig. 4.23 and 4.24), which indicate water column stratification<br />

during <strong>organic</strong> matter source deposition (Sinninghe Damsté et al., 1995;<br />

Yangming et al., 2005).<br />

The Moretane/Hopane, Hopane/Hopane + Moretane, Ts/Ts + Tm, 22S/22S + 22R<br />

C32 homohopane ratios in Onyeama samples range from 0.46 to 0.64; 0.61 to 0.69;<br />

0.02 to 0.05; and 0.48 to 0.58 respectively. These values suggest low maturity status<br />

(Rullkötter et al., 1985; Kagya, 1996; Peters et al., 2005). The Okaba samples have<br />

Moretane/Hopane, Hopane/Hopane + Moretane, Ts/Ts + Tm, 22S/22S + 22R C32<br />

homohopane ratios ranging from 0.59-0.93; 0.54-0.64; 0.11-0.24; and 0.16-0.48<br />

respectively. These values also indicate that Onyeama samples are thermally<br />

immature (Rullköter et al., 1985;Kagya, 1996; Peters et al., 2005).<br />

133


4.3.5 Steranes<br />

The m/z 217 mass chromatograms showing the distribution <strong>of</strong> steranes and<br />

diasteranes in all the samples are shown in Figs. 4.28,4.29 and 4.30. Peak identities<br />

are listed in Table 4.9.<br />

4.3.5.1 Awgu Formation<br />

The occurrence <strong>of</strong> C27 to C29 steranes and diasteranes were detected in Awgu<br />

<strong>coal</strong>y samples (Fig. 4.28). The sterane and diasterane distributions for all the samples<br />

occur in the order <strong>of</strong> C29>C28>C27 (Table 4.7). The predominance <strong>of</strong> C29 sterane over<br />

C27 sterane reflects a greater input <strong>of</strong> terrestrial relative to marine <strong>organic</strong> matter<br />

(Huang and Meinschein, 1979; Volkman 1988; Kagya, 1996; Sari and Bahtiyar, 1999;<br />

Otto et al., 2005; Peters et al., 2005). The ternary plots <strong>of</strong> sterane distribution in Awgu<br />

samples (Fig. 4.31) indicate <strong>organic</strong> matter derived from terrestrial materials<br />

deposited in lacustrine – fluvial/deltaic settings (Huang and Meinschein, 1979;<br />

Killops and Killops, 1993, 2005; Peters et al., 2005). The diasterane ternary plots<br />

(Fig. 4.32) also show that Awgu samples are from terrestial <strong>organic</strong> matter. This<br />

observation is supported by C27/C29 ratios (Table 4.7), which range from 0.28 to 0.56<br />

(Peters et al., 2005). The dominance <strong>of</strong> dinosterol over C30 steranes in these samples<br />

reflects typical fresh water lacustrine source rocks (Köhler and Clausing, 2000; Peters<br />

et al., 2005).<br />

The 20S/20S+20R and αββ/αββ+ααα C29 ratios range from 0.43 to 0.58 and 0.42<br />

to 0.55 respectively. These values show that the samples are within the oil generative<br />

window (Peters et al., 2005). Plots <strong>of</strong> 22S/22S+22R C32 hopanes against<br />

C29αββ/αββ+ααα steranes show that Awgu samples are thermally mature (Fig. 4.33).<br />

134


Fig. 4.28 : m/z 217 mass chromatograms showing the distribution <strong>of</strong> steranes<br />

and diasteranes in Awgu samples.<br />

135


Fig. 4.29: m/z 217 mass chromatograms showing the distribution <strong>of</strong> steranes and<br />

diasteranes in Mamu samples (Okaba).<br />

136


Fig. 4.30 : m/z 217 mass chromatograms showing the distribution <strong>of</strong> steranes and<br />

diasteranes in Mamu samples (Onyeama).<br />

137


Table 4.9 : Peak identities on m/z 217 mass chromatograms <strong>of</strong> Nigerian <strong>coal</strong>.<br />

Peak Compound<br />

1<br />

C27 13β(H), 17α(H)-Diasterane (20S)<br />

2 C27 13β(H), 17α(H)-Diasterane (20R)<br />

3 C27 13α(H), 17β(H)-Diasterane (20S)<br />

4 C27 13α(H), 17β(H)-Diasterane (20R)<br />

5 C28 13β(H), 17α(H)-Diasterane (20S)<br />

6 C28 13β(H), 17α(H)-Diasterane (20R)<br />

7 C28 13α(H), 17β(H)-Diasterane (20S)<br />

8 C27 5α(H), 14α(H), 17α(H)-Sterane (20S) + C28 13α(H), 17β(H)-Diasterane (20S)<br />

9 C27 5α(H), 14β(H), 17β(H)-Sterane (20R) + C28 13β(H), 17α(H)-Diasterane (20S)<br />

10 C27 5α(H), 14β(H), 17β(H)-Sterane (20S) + C28 13β(H), 17α(H)-Diasterane (20R)<br />

11 C27 5α(H), 14α(H), 17α(H)-Sterane (20R)<br />

12 C29 13β(H), 17α(H)-Diasterane (20R)<br />

13 C29 13α(H), 17β(H)-Diasterane (20S)<br />

14 C28 5α(H), 14α(H), 17α(H)-Sterane (20S)<br />

15 C28 5α(H), 14β(H), 17β(H)-Sterane (20R)<br />

16 C28 5α(H), 14β(H), 17β(H)-Sterane (20S)<br />

17 C28 5α(H), 14α(H), 17α(H)-Sterane (20R)<br />

18 C29 5α(H), 14α(H), 17α(H)-Sterane (20S)<br />

19 C29 5α(H), 14β(H), 17β(H)-Sterane (20R)<br />

20 C29 5α(H), 14β(H), 17β(H)-Sterane (20S)<br />

21 C29 5α(H), 14α(H), 17α(H)-Sterane (20R)<br />

C30 C30 5α(H), 14α(H), 17α(H)-Sterane<br />

138


Fig. 4.31 : Ternary plots <strong>of</strong> C27, C28 and C29 steranes distributions in Nigerian<br />

<strong>coal</strong> (After Huang and Meinschein, 1979).<br />

139


Fig. 4.32 : Ternary plots <strong>of</strong> C27, C28 and C29 diasteranes distributions in Nigerian<br />

<strong>coal</strong> (After Huang and Meinschein, 1979).<br />

140


Fig. 4.33 : Plots <strong>of</strong> 22S/22S+22R C32hopanes against C29αββ/αββ+ααα steranes<br />

(After Inaba et al., 2001).<br />

141


4.3.5.2 Mamu Formation<br />

C29 Diasteranes and steranes are the most abundant steranes in all the samples<br />

except few samples from Okaba where C28 predominates. The sterane and diasterane<br />

distributions in Okaba samples are increasing in the order <strong>of</strong> C29>C28>C27.<br />

The predominance <strong>of</strong> C29 over C27 sterane reflects a greater input <strong>of</strong> terrestrial relative<br />

to marine <strong>organic</strong> matter (Huang and Meinschein, 1979; Volkman, 1988; Kagya, 1996;<br />

Sari and Bahtiyar, 1999; Otto et al., 2005; Peters et al., 2005). The appreciable<br />

quantity <strong>of</strong> C27 and C28 in these samples also reflect contributions from phytoplankton;<br />

algae, diatoms, din<strong>of</strong>lagellates (Volkman, 1986; Volkman et al., 1998; Sari and<br />

Bahtiyar, 1999; Peters et al., 2005). The ternary plot <strong>of</strong> C27, C28 and C29 sterane <strong>of</strong><br />

Mamu samples (Fig. 4.31) reflects major terrestrial input in Onyeama samples while<br />

Okaba samples consist <strong>of</strong> both terrestrial and marine <strong>organic</strong> matter (Huang and<br />

Meinschein, 1979; Killops and Killops, 1993, 2005; Peters et al., 2005).<br />

The diasterane ternary plot (Fig. 4.32) shows that most <strong>of</strong> the Onyeama samples<br />

are derived from terrestrial <strong>organic</strong> matter with few samples having mixed inputs<br />

(i.e. terrestrial and marine). Samples from Okaba are majorly derived from mixed<br />

origin. There is little variation in sterane and diasterane distribution in Onyeama<br />

samples while significant variations are noticed in Okaba samples. This observation<br />

possibly reflects same depositional environments for Onyeama samples<br />

(fluvial/deltaic) and lacustrine-fluvial/deltaic depositional environments for Okaba<br />

samples. This observation can be supported by the ratio <strong>of</strong> C27/C29 (Table 4.7) for the<br />

samples. The values range from 0.2 to 0.7 and 0.18 to 0.43 in Okaba and Onyeama<br />

samples respectively. The dominance <strong>of</strong> dinosterol over C30 steranes in Okaba<br />

samples reflects typical fresh water lacustrine source rocks (Köhler and Clausing,<br />

2000; Peters et al., 2005).<br />

142


The 20S/20S+20R and αββ/αββ+ααα C29 ratios range from 0.04 to 0.19 and 0.16<br />

to 0.25 respectively in Onyeama samples while the values range from 0.1 to 0.39 and<br />

0.21 to 0.56 respectively in Okaba samples. The generally low values recorded<br />

indicate that the samples are thermally immature. The plot <strong>of</strong> 22S/22S+22R C32<br />

hopanes against C29αββ/αββ+ααα steranes also confirm the thermal immaturity status<br />

<strong>of</strong> Mamu samples (Fig. 4.33).<br />

143


4.4 POLYCYCLIC AROMATIC HYDROCARBONS IN NIGERIAN COAL<br />

Mass chromatograms showing the distributions <strong>of</strong> naphthalenes, phenanthrenes<br />

and dibenzothiophenes and their alkyl derivatives are shown in Figs.<br />

4.34,4.35,4.36,4.37,4.38,4.39 and 4.41 respectively. Peak identities are given in Tables<br />

4.10,4.11 and 4.13. These compounds were also used to assess the origin, depositional<br />

environment and thermal maturity <strong>of</strong> the <strong>coal</strong> samples.<br />

4.4.1 Naphthalene and Alkylnaphthalenes<br />

4.4.1.1 Awgu Formation<br />

The m/z 156 and 170 showing the distribution <strong>of</strong> dimethyl and trimethyl<br />

naphthalenes are shown in Fig. 4.34. In the aromatic fractions <strong>of</strong> Awgu <strong>coal</strong>ly samples,<br />

10 dimethylnaphthalene (DMN) and 10 trimethylnaphathalene (TMN) were detected.<br />

Also the presence <strong>of</strong> appreciable amounts <strong>of</strong> 1,6-, 1,7-, and 2,6-DMNs in all the<br />

samples indicate terrestrial <strong>organic</strong> matter input (Day and Edman, 1963; Achari et al.,<br />

1973).The presence <strong>of</strong> 1,2,5- and 1,2,7-TMN in all the samples indicate both<br />

angiosperm and gymnosperms material contribution to the <strong>organic</strong> matter that formed<br />

the <strong>coal</strong>s (Killops and Killops, 2005). There is predominance <strong>of</strong> 1,2,5-TMN over<br />

1,2,7-TMNs in most <strong>of</strong> the samples except in LB542, LB523 and LB420. The other<br />

prominent trimethylnaphthalene in the samples is 1,2,6-TMN. The occurrence <strong>of</strong> 1,2,6<br />

TMN in sediments has been traced to microbial source (Alexander et al., 1992).<br />

Therefore its presence in the samples also indicates microbial input to the biomass.<br />

144


Fig. 4.34 : m/z 156, 170 mass chromatograms showing the distribution <strong>of</strong> naphthalene and<br />

alkylnaphthalenes in Awgu Samples.<br />

145


Fig. 4.34 (contd.): m/z 156, 170 mass chromatograms showing the distribution <strong>of</strong> naphthalene<br />

and alkylnaphthalenes in Awgu Samples.<br />

146


Fig. 4.35 : m/z 156, 170 mass chromatograms showing the distribution <strong>of</strong> naphthalene and<br />

alkylnaphthalenes in Mamu samples (Okaba).<br />

147


Fig. 4.36 : m/z 156, 170 mass chromatograms showing the distribution <strong>of</strong> naphthalene and<br />

alkylnaphthalenes in Mamu Samples (Onyeama).<br />

148


Table 4.10 : Peak identities on m/z 156 and 170 mass chromatograms.<br />

Peak Compound<br />

1,2-DMN<br />

1,2-dimethylnaphthalene<br />

1,3-DMN 1,3- dimethylnaphthalene<br />

1,4-DMN 1,4- dimethylnaphthalene<br />

1,5-DMN 1,5- dimethylnaphthalene<br />

1,6-DMN 1,6- dimethylnaphthalene<br />

1,7-DMN 1,7- dimethylnaphthalene<br />

1,8-DMN 1,8- dimethylnaphthalene<br />

2,3-DMN 2,3- dimethylnaphthalene<br />

2,6-DMN 2,6- dimethylnaphthalene<br />

2,7-DMN 2,7- dimethylnaphthalene<br />

1,2,5-TMN 1,2,5- trimethylnaphthalene<br />

1,2,6-TMN 1,2,6- trimethylnaphthalene<br />

1,2,7-TMN 1,2,7- trimethylnaphthalene<br />

1,3,5-TMN 1,3,5- trimethylnaphthalene<br />

1,3,6-TMN 1,3,6- trimethylnaphthalene<br />

1,3,7-TMN 1,3,7- trimethylnaphthalene<br />

1,4,6-TMN 1,4,6- trimethylnaphthalene<br />

1,6,7-TMN 1,6,7- trimethylnaphthalene<br />

2,3,6-TMN 2,3,6- trimethylnaphthalene<br />

149


4.4.1.2 Mamu Formation<br />

Figures 4.35 and 4.36 show the m/z 156 and 170 distribution <strong>of</strong> dimethyl and<br />

trimethyl naphthalenes in Mamu samples. The samples have similar alkyl naphthalene<br />

distributions as observed in Awgu samples. Also 10 dimethylnaphthalene (DMN) and<br />

10 trimethylnaphthalene (TMN) were detected in all the samples. The presence <strong>of</strong><br />

appreciable amounts <strong>of</strong> 1,6-, 1,7-, and 2,6-DMNs in all the samples indicate terrestrial<br />

<strong>organic</strong> matter input (Day and Edman, 1963; Achari et al., 1973). Both 1,2,5- and<br />

1,2,7-TMN were detected in the samples which indicate contributions from both<br />

gymnosperm and angiosperm <strong>organic</strong> matter (Killops and Killops, 2005). In addition,<br />

the occurrence <strong>of</strong> 1,2,6-TMN also indicates microbial contribution to the biomass<br />

(Alexander et al., 1992).<br />

4.4.2 Phenanthrene and alkylphenanthrenes<br />

The m/z 178, 192 and 206 chromatograms showing the distribution <strong>of</strong><br />

phenanthrene and alkylphenanthrenes in the samples are shown in Fig. 4.37. The peak<br />

identities are listed in Table 4.11. Phenanthrene, four methylphenanthrene (MP) and<br />

ten dimethylphenanthrene (DMP) were detected in the aromatic fraction <strong>of</strong> the<br />

samples.<br />

4.4.2.1 Awgu Formation<br />

Plots <strong>of</strong> 9MP/1MP+9MP ratio against Paq for Awgu samples are shown in<br />

Fig. 4.40. This plot has been proposed to discriminate between terrestrial and marine<br />

<strong>organic</strong> matter (modified after Fickens et al., 2000). All the samples fall within the<br />

field <strong>of</strong> marine <strong>organic</strong> matter source (Fig. 4.40), an indication that the <strong>coal</strong>s were<br />

derived from <strong>organic</strong> matter formed mainly from marine materials. Also the<br />

occurrence <strong>of</strong> 1,7-DMP (pimanthrene) in the samples is attributed to terrestrial<br />

<strong>organic</strong> matter (Simoneit et al., 1986).<br />

150


Various maturity parameters calculated from the distributions <strong>of</strong> phenanthrene and<br />

methylphenanthrenes in the <strong>coal</strong> samples are given in Table 4.14. The calculated<br />

vitrinite reflectance (%Rc), MPI-1, MPI-1, MPI-2, MPI-1*, MPDF and PAI values<br />

range from, 0.78 to 1.17, 0.63 to 1.88, 0.78 to 1.23, 0.59 to 1.72, 0.46 to 0.77 and 1.01<br />

to 1.68 respectively.<br />

There is no remarkable difference between MPI-1, MPI-2 and MPI-1* ratios in the<br />

samples. The calculated vitrinite reflectance (Rc) correlates well with vitrinite<br />

reflectance obtained from the plot <strong>of</strong> HI vs. Tmax (Figs. 4.5 and 4.6) for type III<br />

kerogen. The Rc values show that the samples are in late window/gas phase.<br />

151


Fig. 4.37 : m/z 178, 192, 206 mass chromatograms showing the distribution <strong>of</strong> phenanthrene and<br />

alkylphenanthrenes in Awgu Samples.<br />

152


Fig. 4.38 : m/z 178, 192, 206 mass chromatograms showing the distribution <strong>of</strong> phenanthrene and<br />

alkylphenanthrenes in Mamu Samples (Okaba).<br />

153


Fig. 4.39 : m/z 178, 192, 206 mass chromatograms showing the distribution <strong>of</strong> phenanthrene and<br />

alkylphenanthrenes in Mamu Samples (Onyeama).<br />

154


Table 4.11 : Peak identities on m/z 178,192,206 mass chromatograms.<br />

Peaks Compounds<br />

P<br />

155<br />

Phenanthrene<br />

1-MP 1-methylphenanthrene<br />

2-MP 2-methylphenanthrene<br />

3-MP 3-methylphenanthrene<br />

9-MP 9-methylphenanthrene<br />

1,2-DMP 1,2-dimethylphenanthrene<br />

1,7-DMP 1,7-dimethylphenanthrene<br />

1,8-DMP 1,8-dimethylphenanthrene<br />

1,9-DMP 1,9-dimethylphenanthrene<br />

2,3-DMP 2,3-dimethylphenanthrene<br />

2,6-DMP 2,6-dimethylphenanthrene<br />

2,7-DMP 2,7-dimethylphenanthrene<br />

2,9-DMP 2,9-dimethylphenanthrene<br />

3,6-DMP 3,6-dimethylphenanthrene<br />

3,9-DMP 3,9-dimethylphenanthrene


Samples Depth (m)<br />

Awgu<br />

Formation<br />

Table 4.12 : Source and Depositional Environment and maturity parameters<br />

derived from phenanthrene and dimethylbenzothiophene and their alkyl<br />

derivatives.<br />

Paq= nC23 + nC25/ nC23 + nC25 + nC29 + nC31 - Alkanes<br />

MPI-1 = 1.5 (2-MP + 3-MP)/ 0.69P + 1-MP + 9-MP; MPI-2 = 3(2-MP)/P + 1-MP + 9-MP<br />

MPI-1* = 1.89(2-MP + 3-MP) / P + 1.26(1-MP + 9-MP); MPDF = 3-MP + 2-MP/ sum <strong>of</strong> methylphenanthrenes<br />

PAI = 1-MP + 2-MP + 3-MP + 9-MP / Phenanthrene<br />

%Rc, MPI-1 = 0.6(MPI-1) + 0.4; R0 < 1.35%, -0.6(MPI-1) + 2.3; R0>1.35%<br />

MDR=4-MDBT/1-MDBT; MDR’=4-MDBT/1-MDBT+4-MDBT; EDR’=4,6-DMDBT/4-EDBT+4,6-DMDBT<br />

OY – Onyeama<br />

OKB – Okaba<br />

Lithology<br />

(Rank)<br />

(B) – Bituminous<br />

(SB) – Sub-bituminous<br />

DBT / P MDBT / MP 9MP / 9MP+1MP Paq<br />

156<br />

%Rc,<br />

MPI-1<br />

MPI-1 MPI-2 MPI-1* MPDF PAI MDR MDR’ EDR’<br />

LB387 218.5-222.5 Carbonaceous shale 0.07 0.08 0.53 0.64 0.88 0.8 0.78 0.75 0.49 1.68 4.23 0.81 0.80<br />

LB417 407.4-412.5 Coaly shale 0.14 0.37 0.44 0.64 0.78 0.63 ND 0.59 0.46 1.28 8.74 0.90 0.85<br />

LB420 417-422 Coal (B) 0.1 0.18 0.47 0.78 0.96 0.93 0.99 0.86 0.6 1.23 6.53 0.87 0.84<br />

LB514 131.7-136.6 Coal (B) 0.16 0.28 0.66 0.69 1.08 1.13 ND 1.04 0.65 1.36 4.38 0.81 0.84<br />

LB518 148 Coal (B) 0.01 0.1 0.48 0.70 0.96 0.93 0.91 0.86 0.56 1.47 5.53 0.85 0.86<br />

LB523 168.8-173.7 Coal (B) 0.15 0.32 0.5 0.66 1.2 1.33 ND 1.21 0.71 1.37 6.13 0.86 0.84<br />

LB534 212.1-216.2 Coal (B) 0.03 0.13 0.63 0.69 1.17 1.88 ND 1.72 0.77 1.83 6.49 0.87 0.91<br />

LB542 247 Carbonaceous shale 0.14 0.36 0.53 0.61 0.95 0.92 0.96 0.83 0.64 1.01 7.08 0.88 0.9<br />

LB551 286-289 Coal (B) 0.13 0.24 0.57 0.65 1.14 1.23 1.23 1.14 0.65 1.6 9 0.9 0.88<br />

Mamu<br />

Formation<br />

OY5 4.6-5.8 Carbonaceous shale 0.04 0.05 0.58 0.26 0.8 0.66 0.68 0.62 0.46 1.42 3.06 0.75 0.78<br />

OY4 6-6.8 Coal (SB) 0.18 0.33 0.94 0.16 0.61 0.34 0.26 0.32 0.3 1.09 0.42 0.30 0.92<br />

OY3 6.8-7.3 Coal (SB) 0.15 0.18 0.93 0.19 0.64 0.4 0.31 0.31 0.26 1.42 0.67 0.40 0.90<br />

OY2 7.6-8.4 Coal (SB) 0.13 0.16 0.82 0.29 0.69 0.48 0.4 0.45 0.41 0.98 0.75 0.43 0.87<br />

OY1 8.4-8.8 Coal (SB) 0.31 0.14 0.75 0.29 0.77 0.61 0.63 0.57 0.41 1.66 1.01 0.50 0.91<br />

OY6 9-9.5 Carbonaceous shale 0.02 0.11 0.57 0.25 0.65 0.42 0.4 0.38 0.48 0.59 2.61 0.72 0.92<br />

OKB4 16.5-17.6 Carbonaceous shale 0.06 0.12 0.46 0.15 0.73 0.55 0.51 0.51 0.48 0.89 2.38 0.70 0.53<br />

OKB3 18.0-18.5 Coal (SB) 0.1 0.11 0.72 0.06 0.68 0.46 0.42 0.39 0.44 0.7 2.71 0.73 0.68<br />

OKB2 18.6-18.9 Coal (SB) 0.09 0.17 0.62 0.05 0.59 0.31 0.29 0.28 0.46 0.41 2.85 0.74 0.78<br />

OKB1 18.9-19.3 Coal (SB) 0.08 0.09 0.67 0.06 0.64 0.41 0.41 0.37 0.39 0.82 2.68 0.73 0.52<br />

OKB7 19.3-19.6 Coal (SB) 0.09 0.15 0.72 0.09 0.65 0.42 0.39 0.38 0.53 0.49 3.06 0.75 0.55<br />

OKB8 19.6-20 Coal (SB) 0.1 0.21 0.59 0.07 0.64 0.41 0.36 0.36 0.59 0.4 5.42 0.84 0.79


Fig. 4.40 :Organic matter source discrimination in the samples (modified after Ficken et al., 2000).<br />

157


4.4.2.2 Mamu Formation<br />

The m/z 178,192 and 206 mass chromatograms showing the distribution <strong>of</strong><br />

phenanthrene and alkylphenanthrenes in the samples are shown in Figs. 4.38 and 4.39.<br />

Plot <strong>of</strong> 9MP/1MP+9MP ratio against Paq is shown in Fig. 4.40. The samples classified<br />

under <strong>organic</strong> matter derived majorly from terrestrial materials. Maturity parameters<br />

derived from phenanthrene and methylphenanthrenes are presented in Table 4.12.<br />

The calculated vitrinite reflectance (% Rc), MPI-1, MPI-2, MPI-1*, MPDF and PAI<br />

values range from, 0.59 to 0.8, 0.31 to 0.66, 0.26 to 0.68, 0.28 to 0.62, 0.26 to 0.59<br />

and 0.4 to 1.66 respectively. All the MPI ratios i.e. MPI-1, MPI-2, MPI-1* are in close<br />

agreement with each other. The calculated vitrinite reflectance, Rc shows that<br />

Onyeama samples are more mature than Okaba samples and are at the beginning <strong>of</strong><br />

oil generative window. The enhanced values <strong>of</strong> Rc recorded for Mamu samples<br />

compared to vitrinite reflectance obtained from the plot <strong>of</strong> HI vs. Tmax<br />

(Figs. 4.5 and 4.6) is consistent with type II and type II/III kerogen classification for<br />

these samples. Rc has been reported to be more effective in estimating thermal<br />

maturity <strong>of</strong> type III <strong>organic</strong> matter (e.g. Buldzinski et al., 1995; Dzou et al., 1995;<br />

Hughes et al., 1995; Radke et al., 1995).<br />

4.4.3 Dibenzothiophene and alkyldibenzothiophenes<br />

The m/z 184, 198, 212 and 226 mass chromatograms showing the distributions<br />

<strong>of</strong> dibenzothiophene and alkydibenzothiophenes are shown in Fig. 4.41. The peak<br />

identities are listed in Table 4.13. Dibenzothiophene, four methyldibenzothiophene<br />

(MDBT), 4-ethyldibenzothiophene, eleven dimethyldibenzothiophene (DMDBT) and<br />

unidentified trimethyldibenzothiophene (TMDBT) were detected in the aromatic<br />

fraction <strong>of</strong> the samples. Various geochemical parameters calculated from the<br />

distributions are given in Table 4.12.<br />

158


4.4.3.1 Awgu Formation<br />

Dibenzothiophene / phenanthrene and methyldibenzothiophene /methylphenanthrene<br />

ratios in the samples range from 0.07 to 0.16 and 0.08 to 0.37 respectively<br />

(Table 4.12). Low amounts <strong>of</strong> dibenzothiophene to phenanthrene and<br />

methyldibenzothiophene to methylphenanthrene indicate deposition <strong>of</strong> the <strong>organic</strong><br />

matter predominantly in lacustrine and/ or fluvial/deltaic environment<br />

(Hughes et al., 1995). These values indicate appreciable quantity <strong>of</strong> terrestrial input<br />

(Requejo, 1994). The plots <strong>of</strong> %S vs. DBT/Phen and %S vs. MDBT/MP have been<br />

used to assess the depositional environment <strong>of</strong> <strong>organic</strong> matter (Hughes et al., 1995).<br />

Majority <strong>of</strong> Awgu samples plot within marine shale, other lacustrine and<br />

fluvial/deltaic field while few samples plot within mixed marine (Hughes et al., 1995).<br />

The plots <strong>of</strong> dibenzothiophene/phenanthrene vs pristane/phytane and<br />

methyldibenzothiophene/methylphenanthrene vs pristane/phytane are versatile tools<br />

in discriminating different depositional environment <strong>of</strong> petroleum source rock<br />

(Hughes et al., 1995). Majority <strong>of</strong> the samples fall within the field <strong>of</strong> fluvial/deltaic<br />

depositional environment on both plots (Fig. 4.44 and 4.45).<br />

159


Fig. 4.41: m/z 184,198,212,226 mass chromatograms showing the distributions <strong>of</strong><br />

dibenzothiophene and alkyldibenzothiophenes in Awgu Samples.<br />

160


Fig. 4.41 (contd.): m/z 184,198,212,226 mass chromatograms showing the distributions <strong>of</strong><br />

dibenzothiophene and alkyldibenzothiophenes in Awgu samples.<br />

161


Table 4.13 : Peak identities on m/z 184, 198, 212, 226 mass chromatograms.<br />

Peaks Compounds<br />

DBT Dibenzothiophene<br />

4-MDBT 4-methyldibenzothiophene<br />

2-MDBT 2-methyldibenzothiophene<br />

3-MDBT 3-methyldibenzothiophene<br />

1-MDBT 1-methyldibenzothiophene<br />

4-EDBT 4-ethyldibenzothiophene<br />

4,6-DMDBT 4,6-dimethyldibenzothiophene<br />

2,4-DMDBT 2,4-dimethyldibenzothiophene<br />

2,6-DMDBT 2,6-dimethyldibenzothiophene<br />

3,6-DMDBT 3,6-dimethyldibenzothiophene<br />

3,7-DMDBT 3,7-dimethyldibenzothiophene<br />

1,4-DMDBT 1,4-dimethyldibenzothiophene<br />

1,6-DMDBT 1,6-dimethyldibenzothiophene<br />

1,8-DMDBT 1,8-dimethyldibenzothiophene<br />

1,3-DMDBT 1,3-dimethyldibenzothiophene<br />

1,2-DMDBT 1,2-dimethyldibenzothiophene<br />

1,9-DMDBT 1,9-dimethyldibenzothiophene<br />

TMDBTs Trimethyldibenzothiophenes<br />

162


Fig. 4.42 : Cross plot <strong>of</strong> total Sulphur vs. dibenzothiophene/phenanthrene (DBT/P) for<br />

Nigerian <strong>coal</strong> (Hughes et al., 1995).<br />

163


Fig. 4.43 : Cross plot <strong>of</strong> total Sulphur vs. methyldibenzothiophene/methylphenanthrene<br />

(MDBT/MP) for Nigerian <strong>coal</strong> (Hughes et al., 1995).<br />

164


Fig. 4.44 : Cross plot <strong>of</strong> dibenzothiophene/phenanthrene (DBT/P) vs. pristane/phytane <strong>of</strong><br />

Nigerian <strong>coal</strong> (Hughes et al., 1995).<br />

165


Fig. 4.45 : Cross plot <strong>of</strong> methyldibenzothiophene/methylphenanthrene (MDBT/MP) vs.<br />

pristane/phytane <strong>of</strong> Nigerian <strong>coal</strong> (Hughes et al., 1995).<br />

166


4.4.3.2 Mamu Formation<br />

The dibenzothiophene/phenanthrene and methyldibenzothiophene<br />

/methylphenanthrene ratios range from 0.02 to 0.31 and 0.05 to 0.33 (Table 4.12)<br />

respectively. These values indicate appreciable quantity <strong>of</strong> terrestrial <strong>organic</strong> matter<br />

(Requejo, 1994). Majority <strong>of</strong> the samples plot within zone I on the plots <strong>of</strong> %S vs.<br />

DBT/Phen and %S vs. MDBT/MP (Figs. 4.42 and 4.43) which represents marine<br />

shale, other lacustrine and fluvial/deltaic while only one Okaba sample plot at the<br />

boundary with zone III which represent marine mixed (Hughes et al., 1995). The low<br />

amounts <strong>of</strong> dibenzothiophene to phenanthrene and methyldibenzothiophene to<br />

methylphenanthrene (Table 4.12) indicate deposition <strong>of</strong> the <strong>organic</strong> matter<br />

predominantly in lacustrine and/ or fluvial/deltaic environment (Hughes et al., 1995).<br />

Figure 31 show the plots <strong>of</strong> dibenzothiophene/phenanthrene vs pristane/phytane and<br />

methyldibenzothiophene/methylphenanthrene vs pristane/phytane for Mamu samples<br />

and most <strong>of</strong> the samples plot within the field <strong>of</strong> fluvial/deltaic depositional<br />

environments.<br />

167


4.5 Carbon Isotopic Composition <strong>of</strong> Nigerian Coal<br />

The δ 13 C isotopic values for individual n-alkanes and whole saturate fractions<br />

<strong>of</strong> the samples are presented in Tables 4.14 and 4.15.<br />

4.5.1 Origin and Depositional Environment <strong>of</strong> Organic Matter<br />

4.5.1.1 Awgu Formation<br />

The carbon isotopic compositions <strong>of</strong> individual alkanes (nC14-nC33) for Awgu<br />

samples range between –30.6 to –25.8 δ 13 ‰PDB. The most depleted values were<br />

observed for the long chain alkanes (nC21-nC33). These features are characteristics <strong>of</strong><br />

plant wax derived n-alkanes <strong>of</strong> C3-plants (Schouten et al., 2000; Hu et al., 2002).<br />

Significant contribution from marine <strong>organic</strong> matter (i.e.C3 algae or cyanobacteria) is<br />

reflected in heavier δ 13 C isotope values <strong>of</strong> short (nC14-nC22) n-alkanes. The n-alkanes<br />

become isotopically lighter with increasing chain length (Fig. 4.46). This observation<br />

reflect terrigenous input (Murray et al., 1994; Schouten et al., 2001; Yangming et al.,<br />

2005) deposited in fluvial/ deltaic depositional environment while the carbon isotopic<br />

composition <strong>of</strong> n-alkanes for the samples that range from –29.0 to –26.3 δ 13 ‰PDB is<br />

typical <strong>of</strong> lacustrine phytoplankton (Murray et al., 1994). Therefore Awgu samples<br />

can be said to consist <strong>of</strong> both terrestrial and marine <strong>organic</strong> matter deposited in<br />

lacustrine-fluvial/deltaic depositional environment.<br />

The δ 13 C isotope ratios <strong>of</strong> Pr and Ph range from –30.2 to –25.5 δ 13 ‰PDB (av.<br />

–27.9) and –30.9 to –26.6 δ 13 ‰PDB (av. –28.22) respectively. There is no significant<br />

difference between the 13 C-values <strong>of</strong> Pr and Ph, which suggest that they were formed<br />

from the same group <strong>of</strong> organisms or organisms depending on the same source <strong>of</strong><br />

carbon (Schwas and Spangenberg, 2007).<br />

168


Table 4.14: Carbon Isotopic Composition <strong>of</strong> n-Alkanes in Awgu Samples (δ 13 ‰PDB).<br />

Sample no Depth (m) C13 C14 C15 C16 C17 C18 C19 C20 C21 C22 C23 C24 C25 C26 C27 C28 C29 C30 C31 C32 C33 C34 C35 Weighted Av. Pr Ph<br />

LB387 218.5-222.5 ND -26.5 -26.6 -26.6 -28.6 -27.2 -27.4 -27.6 -28.0 -28.0 -28.0 -27.9 -28.3 -28.8 -29.3 -28.5 -28.0 -28.5 -30.2 -29.6 -30.6 ND ND -28.2 -26.7 -26.6<br />

LB417 407.4-412.5 ND ND -24.7 -26.8 -27.3 -30.1 -27.4 -27.8 -30.2 -29.5 -31.5 -28.4 -28.2 -27.3 -27.3 -27.5 -28.3 -29.4 -29.2 -29.5 -30.2 -29.4 -28.6 -27.3 -25.5 -29.4<br />

LB420 417-422 ND ND -24.0 -27.7 -28.1 -28.5 -30.8 -29.1 -29.6 -28.2 -28.9 -29.5 -30.5 -28.8 -30.8 -28.8 -28.7 -29.0 -30.8 -25.8 -33.3 ND ND -29.0 -29.6 -26.1<br />

LB514 131.7-136.6 ND -26.2 -25.6 -25.5 -26.5 -26.2 -25.5 -25.8 -26.0 -26.3 -26.4 -27.0 -26.9 -28.3 -27.5 -27.6 -26.2 -27.9 -28.7 -29.6 -30.1 ND ND -27.0 -30.2 -29.6<br />

LB518 148.0 ND -26.0 -26.0 -26.6 -26.5 -26.7 -25.6 -27.0 -27.3 -27.5 -27.8 -28.3 -29.3 -28.0 -28.1 -27.9 -27.8 -28.4 -29.1 -28.8 -29.5 ND ND -27.6 -26.9 -28.6<br />

LB523 168.8-173.0 -25.5 -26.0 -26.2 -27.2 -26.9 -26.7 -27.0 -27.3 -27.7 -27.5 -27.5 -27.4 -27.7 -27.4 -28.6 -28.7 -28.6 -29.8 -29.8 -29.9 -30.9 ND ND -27.8 -28.8 -27.0<br />

LB534 212.1-216.2 ND -26.3 -25.9 -25.3 -26.7 -26.4 -26.1 -27.2 -27.6 -27.4 -28.9 -28.4 -28.1 -27.9 -27.9 -27.9 -27.8 -28.9 -29.3 -28.9 -30.6 ND ND -26.3 -29.9 -30.9<br />

LB542 247.0 ND -24.3 -26.3 -26.2 -26.5 -27.9 -28.3 -26.7 -27.9 -27.1 -27.2 -27.5 -28.2 -28.6 -28.4 -28.4 -28.2 -28.4 -28.9 -29.1 -30.3 ND ND -27.7 -27.0 -26.5<br />

LB551 286-289 ND ND -26.4 -26.3 -27.9 -29.4 -26.9 -26.7 -27.8 -27.4 -27.5 -28.2 -28.2 -28.1 -28.3 -28.4 -28.1 -30.5 -28.6 -28.1 -29.8 ND ND -28.0 -26.8 -29.3<br />

Average <strong>of</strong> -25.5 -25.9 -25.8 -26.5 -27.2 -27.7 -27.2 -27.2 -28.0 -27.7 -28.2 -28.1 -28.4 -28.1 -28.5 -28.2 -28.0 -29.0 -29.4 -28.8 -30.6 -29.4 -28.6 -27.9 -28.2<br />

Individual<br />

n-Alkanes<br />

169


Fig. 4.46 : Carbon isotopic distribution <strong>of</strong> individual n-alkanes in Awgu Samples<br />

(Murray et al., 1994).<br />

170


4.5.1.2 Mamu Formation<br />

The δ 13 C isotopic values for n-alkanes range from –31.7 to –29.2 δ 13 ‰PDB and<br />

–30.1 to –28.2 δ 13 ‰PDB in Onyeama and Okaba samples respectively. The isotopic<br />

values are characteristics <strong>of</strong> plant wax derived n-alkanes <strong>of</strong> C3-plants (Canuel et al.,<br />

1997; Sun et al., 2000; Bi et al., 2005; Tanner et al., 2007; Tuo et al., 2007). The<br />

carbon isotopic compositions <strong>of</strong> individual alkanes (nC15-nC36) in Onyeama and<br />

Okaba samples range from –32.19 to –27.15 δ 13 ‰PDB and –32.48 to –27.50<br />

δ 13 ‰PDB respectively. The most depleted values were observed for the long chain<br />

alkanes, which are characteristics <strong>of</strong> plant wax derived n-alkanes <strong>of</strong> C3-plants<br />

(Schouten et al., 2000; Hu et al., 2002).<br />

Significant contribution from marine <strong>organic</strong> matter (i.e.C3 algae or cyanobacteria)<br />

is reflected in heavier δ 13 C isotope values observed in the short chain (nC15-nC18)<br />

alkanes in all the samples (Figs. 4.47 and 4.48). Both pristane and phytane have<br />

essentially the same δ 13 C values as the n-alkanes, suggesting that they were formed<br />

from the same group <strong>of</strong> organisms (Schwas and Spangenberg, 2007).<br />

However, there is notably flat portion pattern <strong>of</strong> the n-alkane pr<strong>of</strong>ile between nC20-<br />

nC25 and nC27-nC31 (Fig. 4.47 and 4.48) in Onyeama and Okaba samples respectively,<br />

which is an indication <strong>of</strong> marine incursion (Murray et al., 1994). These features show<br />

that Mamu samples consist <strong>of</strong> both terrestrial and marine <strong>organic</strong> matter deposited in<br />

fluvial/deltaic (Onyeama) or lacustrine-fluvial/deltaic (Okaba) settings.<br />

171


Table 4.15: Carbon Isotopic Composition <strong>of</strong> n-Alkanes in Mamu Samples (δ 13 ‰PDB).<br />

Sample no Depth (m) C14 C15 C16 C17 C18 C19 C20 C21 C22 C23 C24 C25 C26 C27 C28 C29 C30 C31 C32 C33 C34 C35 C36 Weighted Av. Pr Ph<br />

Okaba<br />

OKB4 16.5-17.6 ND ND -27.4 ND -29.3 -31.3 -30.1 -33.9 -30.8 -31.9 -28.5 -30.1 -29.9 -28.2 -28.2 -30.0 -30.1 -31.1 ND ND ND ND ND -30.1 -29.8 -32.1<br />

OKB3 18.0-18.5 ND ND ND ND -29.7 -26.4 -29.0 ND ND ND ND ND ND -31.5 -32.0 -28.5 -28.5 -32.0 -31.1 -31.5 -32.2 -28.7 ND -30.1 -31.5 -35.4<br />

OKB2 18.6-18.9 -26.2 ND -28.5 ND -29.7 ND ND ND ND ND ND ND ND -28.4 -26.3 -28.2 -26.1 -27.5 -29.7 -31.5 -27.9 ND ND -28.2 -30.4 -29.7<br />

OKB1 18.9-19.3 -26.1 -27.2 -29.2 -27.4 -28.9 -28.3 -32.5 -31.2 -29.3 -30.0 -25.5 -31.2 -31.9 -28.8 -30.9 -27.5 -28.6 -32.6 -30.6 -31.1 -32.8 -28.1 -28.7 -29.5 -28.8 -30.6<br />

OKB7 19.3-19.6 ND ND -27.1 -32.1 -28.2 -28.6 -27.1 -32.3 -32.1 -29.9 -32.4 -31.2 -32.1 -29.1 -28.0 -28.1 -27.7 -29.1 -28.6 -28.2 ND -29.2 ND -29.5 -28.5 -32.0<br />

OKB8 19.6-20.0 ND ND -25.3 -26.1 -26.3 -31.0 -32.4 ND ND ND ND ND ND -29.4 ND -30.2 -29.4 -30.2 -30.9 ND -30.4 ND ND -29.3 -27.6 -30.6<br />

Average <strong>of</strong> -26.1 -27.2 -27.5 -28.6 -28.8 -29.1 -30.2 -32.5 -30.7 -30.6 -29.8 -30.9 -31.3 -29.2 -29.1 -28.7 -28.4 -30.4 -30.2 -30.6 -30.8 -28.7 -28.7 -29.4 -31.7<br />

Individual<br />

n-Alkanes<br />

Onyeama<br />

OY5 4.6-5.8 ND -28.8 -28.9 -28.0 -30.0 -32.2 -32.2 -33.1 -33.3 -33.3 -33.0 -31.6 -31.8 -30.8 -31.7 -31.2 -34.2 -30.7 ND ND -33.6 ND ND -31.6 -29.8 -31.6<br />

OY4 6-6.8 ND -27.6 -27.4 ND -30.1 -31.1 -30.1 -31.5 -30.6 -30.5 -30.1 -31.3 -30.0 -29.2 -30.4 -29.6 -32.1 -30.5 -32.0 -31.9 -29.4 -32.7 -33.1 -30.5 -28.9 -29.7<br />

OY3 6.8-7.3 ND ND -33.9 ND -33.3 -34.1 -31.2 -29.9 -30.9 -31.0 -31.4 -31.7 -29.8 -29.2 -30.2 -29.1 -31.5 -30.6 -29.7 -29.5 -30.4 -33.0 ND -31.7 -29.8 -32.6<br />

OY2 7.6-8.4 ND -27.9 -26.6 -26.8 -26.3 -28.8 -29.3 -30.6 -29.6 -29.9 -30.9 -29.5 -29.2 -28.8 -29.4 -28.8 -29.9 -30.1 -31.0 -28.6 -30.4 -31.4 -29.3 -29.2 -28.3 -29.5<br />

OY1 8.4-8.8 ND -27.6 -27.5 ND -28.3 -29.1 -28.9 -29.5 -29.5 -30.4 -30.4 -31.0 -29.8 -28.7 -29.4 -28.8 -31.0 -30.0 -30.8 -33.3 -32.2 -31.6 -33.6 -30.1 -29.3 -28.6<br />

OY6 9.0-9.5 ND -27.6 -27.7 -26.7 -30.8 -30.4 -30.9 -30.7 -33.1 -32.6 -33.6 -32.2 -31.9 -29.6 -30.7 -29.8 -32.3 -30.4 -32.2 -32.8 ND ND ND -30.8 -28.8 -30.7<br />

Average <strong>of</strong> ND -27.9 - 28.7 -27.2 -29.8 -30.9 -30.4 -30.9 -31.1 -31.3 -31.6 -31.2 -30.4 -29.4 -30.3 -29.5 -31.8 -30.4 -31.1 -31.2 -31.2 -32.2 -32.0 -29.1 -30.5<br />

Individual<br />

n-Alkanes<br />

172


Fig. 4.47: Carbon isotopic distribution <strong>of</strong> individual n-alkanes in Okaba Samples, Mamu<br />

(Murray et al., 1994).<br />

173


Fig. 4.48: Carbon isotopic distribution <strong>of</strong> individual n-alkanes in Onyeama Samples,<br />

Mamu (Murray et al., 1994).<br />

174


CHAPTER FIVE<br />

SUMMARY AND CONCLUSION<br />

Coal and <strong>coal</strong>y <strong>organic</strong> matter samples were collected from <strong>coal</strong> bearing<br />

measures <strong>of</strong> Lower and Middle Benue Trough, Nigeria.These samples were subjected<br />

to Leco, Rock-Eval pyrolysis, Elemental, Vitrinite reflectance measurement, Gas<br />

Chromatography-Mass Spectrometry and Gas Chromatography-Isotope Ratio- Mass<br />

Spectrometry analyses. This study was undertaken to re-appraise the hydrocarbon<br />

potential <strong>of</strong> Nigerian <strong>coal</strong> through source rock evaluation studies, biomarker<br />

parameters and carbon isotopic compositions.<br />

Majority <strong>of</strong> the samples were characterized by low sulphur contents (


The Pr/Ph ratios <strong>of</strong> samples from Awgu and Onyeama samples reflect <strong>organic</strong><br />

matter deposition under oxic conditions in freshwater-lacustrine and freshwater<br />

depositional environment respectively while Pr/Ph ratios <strong>of</strong> Okaba samples indicate<br />

suboxic-oxic conditions during sedimentation in freshwater-lacustrine depositional<br />

environments.<br />

The presence <strong>of</strong> hopane, homohopane (C31-C35) in all the samples showed that<br />

bacteriohopanetetrol and other polyfunctional C35 hopanoids common in prokaryotic<br />

microorganisms have significant contributions to the <strong>organic</strong> matter that formed the<br />

<strong>coal</strong>s. The occurrence <strong>of</strong> oleanene isomers in Okaba samples favoured terrestrial<br />

<strong>organic</strong> matter deposited in lacustrine-fluvial/deltaic environment. In addition, the<br />

detection <strong>of</strong> gammacerane in Okaba samples represents water stratification during<br />

<strong>organic</strong> matter source deposition. The abundance <strong>of</strong> C29 Steranes and diasteranes in<br />

the samples indicate land input to the <strong>organic</strong> matter that formed the <strong>coal</strong>.<br />

The distribution <strong>of</strong> polyaromatic hydrocarbons and carbon isotopic compositions<br />

<strong>of</strong> the individual alkanes in the samples showed that Awgu samples were formed from<br />

<strong>organic</strong> matter derived from both terrestrial and marine material deposited in<br />

lacustrine-fluvial/deltaic depositional environment. The Mamu samples were derived<br />

from mixed <strong>organic</strong> materials (terrestrial and marine) deposited in fluvio-deltaic<br />

(Onyeama) and lacustrine-fluvial/deltaic settings (Okaba) respectively.<br />

The Vitrinite reflectance values (0.48-1.15 %Ro) and all the maturity parameters<br />

derived from the elemental, Rock Eval analysis and biomarkers distributions show<br />

that Awgu samples were in the late oil window while Mamu samples have low<br />

thermal maturity status. The Awgu Formation <strong>coal</strong>s might have generated gases into<br />

yet-to-be identified reservoir. However, <strong>coal</strong>s from Mamu Formation have potential to<br />

generate both oil and gas but are presently immature to have formed any significant<br />

hydrocarbons.<br />

176


REFERENCES<br />

Abubakar, M.B., Obaje, N.G., Luterbacher, H.P., Dike, E.F.C. and Ashraf, A.R.<br />

2006. A report on the occurrence <strong>of</strong> Albian-Cenomanian elater-bearin<br />

pollen in Nasara-1 well, Upper Benue Trough, Nigeria:<br />

Biostratigraphic and palaeoclimatological implications. Journal <strong>of</strong><br />

African Earth Sciences 45:347- 354.<br />

Achari, R. G., Shaw, G. and Holleyhead, R. 1973. Identification <strong>of</strong> ionene and other<br />

carotenoid degradation products from the pyrolysis <strong>of</strong> sporopollenins derived<br />

from some pollen exines, a spore <strong>coal</strong> and the green river shale. Chemical<br />

Geology 12: 229–234.<br />

Adam, P., Schaeffer, P. and Albrecht, P. 2006.C40 monoaromatic lycopane derivatives<br />

as indicators <strong>of</strong> contribution <strong>of</strong> the algal Botryococcus braunii race L to the<br />

<strong>organic</strong> matter <strong>of</strong> Messel oil shale (Eocene, Germany). Organic Geochemistry<br />

37:584-596.<br />

Akande, S.O., Ojo, O.J., Erdtmann, B.D.and Hetenyi, M. 2005.Paleoenvironments,<br />

<strong>organic</strong> petrology and Rock-Eval studies on source rock facies <strong>of</strong> the Lower<br />

Maastrichtian Patti Formation, southern Bida Basin, Nigeria.Journal <strong>of</strong><br />

African Earth Sciences 41:394-406.<br />

Akande, S.O., Ogunmoyero, I.B., Petersen, H.I. and Nyt<strong>of</strong>t, H.P. 2007. Source Rock<br />

Evaluation <strong>of</strong> Coals from the Lower Maastrichtian Mamu Formation, SE<br />

Nigeria. Journal <strong>of</strong> Petroleum Geology 30.4: 303-324.<br />

Albrecht, P. and Ourisson, G. 1971.Biogenic substances in sediments and fossils.<br />

Angew. Chem. Int. Ed 10:209-225.<br />

Aitzetmuller, K. and Vosmann, K. 1998. Cyclopropenoic fatty acids in gymnosperms:<br />

The seed oil <strong>of</strong> Welwitschia. Journal <strong>of</strong> the American Oil Chemists<br />

Society 75: 1761–1765.<br />

Alexander, R., Strachan, M.G., Kagi, R.I. and van Bronswijk, W. 1986.Heating rates<br />

effects on aromatic maturity indicators. Organic Geochemistry 10:97-1003.<br />

Alexander, R., Bastow, T. P., Kagi, R. I. and Singh, R. K. 1992. Identification <strong>of</strong><br />

1,2,2,5-Tetramethyltetralin and 1,2,2,3,4-Pentamethyltetralin as Racemates<br />

in Petroleum. Journal <strong>of</strong> the Chemical Society, Chemical Communications<br />

1906:1712–1714.<br />

177


Alexander, R., Bastow, T. P., Fisher, S. J. and Kagi, R. I. 1995. Geosynthesis <strong>of</strong><br />

Organic compounds: II. Methylation <strong>of</strong> phenanthrene and<br />

alkylphenanthrenes. Geochimica et Cosmochimica Acta 59:4259–4266.<br />

Aquino Neto, F.R., Tendel, J.M., Restle, A., Connan, J. and Albrecht, P.<br />

1983.Occurrence and formation <strong>of</strong> tricyclic and tetracyclic terpanes in<br />

sediments and petroleums. In: Bjoroy, M.et.al. (Eds.), Advances in Organic<br />

Geochemistry. New York: John Wiley and Sons.659-676.<br />

Arens, N.C., Jahren, A.H. and Amundson, R. 2000.Can C3 plants faithfully record<br />

the carbon isotopic composition <strong>of</strong> atmospheric carbon dioxide? Paleobiology<br />

261:137-164.<br />

Audino, M., Grice, K., Alexander, R. and Kagi, R.I. 2002.Macrocyclic alkanes in<br />

crude oils from the algaenan <strong>of</strong> Botryococcus braunii. Organic Geochemistry<br />

33:979-984.<br />

Azevedo, D.A., Aquino Neto, F.R., Simoneit, B.R.T. and Pinto, A.C. 1992.Novel<br />

series <strong>of</strong> tricyclic aromatic terpanes characterized in Tasmanian tasmanite.<br />

Organic Geochemistry 18:9-16.<br />

Azevedo, D.A., Aquino Neto, F.R. and Simoneit, B.R.T. 1998.Extended ketones <strong>of</strong><br />

the tricyclic terpanes in Tasmanian tasmanite bitumen. Organic Geochemistry<br />

28:289-295.<br />

Bai, Y., Fang, X., Wang, Y., Kenig, F., Miao, Y. and Wang, Y. 2006.Distribution <strong>of</strong><br />

aliphatic ketones in Chinese soils: Potential environmental implications.<br />

Organic Geochemistry 37:860-869.<br />

Balogun, F.A., Mokorbia, C.E., Fasasi, M.K. and Ogundare, F.O. 2003.Natural<br />

radioactivity associated with bituminous <strong>coal</strong> mining in Nigeria.Nuclear<br />

Methods And Methods In Physical Resources A 505:444-448.<br />

Barakat, A.O. and Rullkötter, J. 1997. A comparative study <strong>of</strong> molecular paleosalinity<br />

indicators: chromans, tocopherols and C20 isoprenoid thiophenes in Miocene<br />

lake sediments (Nordlinger Ries, Southern Germany). Aquatic Geochemistry<br />

3:169-190.<br />

Barth, T., Borgund, A.E. and Hopland, A.L. 1989.Generation <strong>of</strong> <strong>organic</strong> compounds<br />

by hydrous pyrolysis <strong>of</strong> Kimmeridge oil shale-bulk results and activation<br />

calculations. Organic Geochemistry 14:69-76.<br />

178


Barthlot, W., Neinhuis, C., Cutler, D., Ditsch, F., Meusel, I., Theisen, I. and Wilhelmi,<br />

H. 1998.Classification and terminology <strong>of</strong> plant epicuticular waxes. Botanical<br />

Journal <strong>of</strong> the Linnean Society 126:237-260.<br />

Bastow, T.P., Alexander, R., Sosrowidjojo, I.B. and Kagi, R.I.<br />

1998.Pentamethylnaphthalenes and related compounds in sedimentary <strong>organic</strong><br />

matter. Organic Geochemistry 28.9-10:585-595.<br />

Bastow, T.P., Alexander, R., Fisher, S.J., Singh, R.K., van Aarssen, B.G.K. and Kagi,<br />

R.I. 2000.Geosynthesis <strong>of</strong> <strong>organic</strong> compounds. Part V-methylation <strong>of</strong><br />

alkylnaphthalenes. Organic Geochemistry 31:523-534.<br />

Bastow, T.P., Snigh, R.K., van Aarssen, B.G.K. and Alexander, R.I. 2001.<br />

2- Methylretene in sedimentary material: a new higher plant biomarker.<br />

Organic Geochemistry 32:1211-1217.<br />

Bechtel, A., Gruber, W., Sachsenh<strong>of</strong>er, R.F., Gratzer, R. and Püttmann, W.<br />

2001.Organic geochemical and stable carbon isotopic investigations <strong>of</strong> <strong>coal</strong>s<br />

formed in low-lying and raised mires within the Eastern Alps (Austria).<br />

Organic <strong>geochemistry</strong> 32:1289-1310.<br />

Bechtel, A., Sachsenh<strong>of</strong>er, R.F., Kolcon, I., Gratzer, R., Otto, W. and Püttmann, W.<br />

2002.Organic <strong>geochemistry</strong> <strong>of</strong> the Lower Miocene Oberd<strong>of</strong>f lignite (Styrian<br />

Basin, Austria): its relation to petrography, palynology and the<br />

palaeoenvironment. International Journal <strong>of</strong> Coal Geology 51:31-57.<br />

Bechtel, A., Sachsenh<strong>of</strong>er, R., F., Markic, M., Gratzer, R., Lucke, A. and Püttmann,<br />

W. 2003.Paleoenvironmental implications from biomarker and stable isotope<br />

investigations on the Pliocene Velenje lignite seam (Slovenia). Organic<br />

Geochemistry 34:1277-1298.<br />

Bechtel, A., Markic, M., Sachsenh<strong>of</strong>er, R.F., Jelen, B., Gratzer, R., Lucke, A.,<br />

Püttmann, W. 2004. Paleoenvironment <strong>of</strong> the upper Oligocene Trbovlje <strong>coal</strong><br />

seam (Slovenia). International Journal <strong>of</strong> Coal Geology 57: 23–48.<br />

Bechtel, A., Gawlick, H. -J., Gratzer, R., Tomaselli, M. and Püttmann, W. 2007.<br />

Molecular indicators <strong>of</strong> palaeosalinity and depositional environment <strong>of</strong> small<br />

scale basins within carbonate platforms: The Late Triassic Hauptdolomite<br />

Wiestalstausee section near Hallein (Northern Calcareous Alps, Austria).<br />

Organic Geochemistry 38.1: 92-111.<br />

179


Bechtel, A., Woszczyk, M., Reischenbacher, D., Sachsenh<strong>of</strong>er, R.F., Gratzer, R.,<br />

Püttmann, W. and Spychalski, W. 2007.Biomarkers and geochemical<br />

indicators <strong>of</strong> Holocene environmental changes in coastal LakeSarbsko<br />

(Poland). Organic Geochemistry 38.7: 1112-1131.<br />

Berner, R.A. 1984. Sedimentary pyrite formation: an update. Geochimica et<br />

Cosmochimica Acta 48:605–615.<br />

Benner, R., Fogel, M.L.Sprague, E.K. and Hodson, R.E. 1987.Depletion <strong>of</strong> 13C in<br />

lignin and its implications for stable carbon isotope studies. Nature 329:708-<br />

710.<br />

Betrand, P., Behar, F. and Durand, B. 1986.Composition <strong>of</strong> potential oil from humic<br />

<strong>coal</strong>s in relation to their petrographic nature. Organic Geochemistry 10:601-<br />

608.<br />

Bi, X., Sheng, G., Liu, X., Li, C. and Fu, J. 2005.Molecular and carbon and hydrogen<br />

isotopic composition <strong>of</strong> n-alkanes in plant leaf waxes. Organic Geochemistry<br />

36:1405-1417.<br />

Bogacheva, M.P. and Galimov, E.M. 1979.Intramolecular distribution <strong>of</strong> carbon<br />

isotopes in chlorophyll and hemine. Geokhimia 7:1166-1172.<br />

Bonfanti, L., Comellas, L., Lliberia, J.L., Vallhonrat-Matalonga, R., Pich-Santacana,<br />

M. and López-Piñol, D. 1997. Pyrolysis gas chromatography <strong>of</strong> some <strong>coal</strong>s by<br />

nitrogen and phosphorous, flame ionization and mass spectrometer detectors.<br />

Journal <strong>of</strong> Analytical and Applied Pyrolysis 44:101-119.<br />

Bordenave, M.L. and Burwood, R. 1990.Source rock distribution and maturation in<br />

the Zagros Orogenic Belt: Provenance <strong>of</strong> the Asmari and Bangestan Reservoir<br />

oil accumulation. Organic Geochemistry 16.1-3:369-389.<br />

Boreham, C.J. and Powell, T.G. 1991. Variation in pyrolysate composition <strong>of</strong><br />

sediments from the Jurassic Walloon <strong>coal</strong> measures, eastern Australia as a<br />

function <strong>of</strong> thermal maturation. Organic Geochemistry 17:723-733.<br />

Boreham, C.J., Summons, R.E., Roksanolic, Z. and Dowling, L.M. 1994.Chemical,<br />

molecular and isotopic differentiation <strong>of</strong> <strong>organic</strong> facies in the Tertiary<br />

lacustrine Duaringa oil shale deposit, Queensland, Australia.Organic<br />

Geochemistry 21:655-712.<br />

Boreham, C.J., Blevin, J.E., Radlinski, A.P. and Trigg, K.R. 2003.Coal as source <strong>of</strong><br />

180


oil and gas, a case study from the Bass basin, Australia. APPEA Journal<br />

43.1:117-148.<br />

Borrego, A.G., Blanco, C.G. and Püttmann, W. 1997.Geochemical significance <strong>of</strong><br />

aromatic hydrocarbon distribution in the bitumens <strong>of</strong> the Puertollano oil shales,<br />

Spain. Organic Geochemistry 26.3/4:219-228.<br />

Brassell, S.C., Eglinton, G., Marlowe, I.T., Pflaumann, U. and Sarnthein, M.<br />

1986.Molecular stratigraphy: a new tool for climatic assessment. Nature<br />

320:129-133.<br />

Brassell, S.C. 1992.Biomarkers in recent and ancient sediments: the importance <strong>of</strong> the<br />

diagenetic continuum. In: Whelan, J.K., Farrington, J.W., editors. Organic<br />

matter-productivity, accumulation, and preservation in recent and ancient<br />

sediments. New York: Columbia University Press. 339-367.<br />

Bray, E.E. and Evans, E.D. 1961.Distribution <strong>of</strong> n-paraffins as a clue to recognition <strong>of</strong><br />

source beds. Geochimica et Cosmochimica Acta 22:2-15.<br />

Brooks, J.D. and Smith, J.W. 1967.The diagenesis <strong>of</strong> plant lipids during the formation<br />

<strong>of</strong> <strong>coal</strong>, petroleum and natural gas-I. Changes in the n-paraffin hydrocarbon.<br />

Geochimica et Cosmochimica Acta 31:2389-2397.<br />

Brooks, J.D. and Smith, J.W. 1969.The diagenesis <strong>of</strong> plant lipids during the formation<br />

<strong>of</strong> <strong>coal</strong>, petroleum and natural gas-II. Coalification and the formation <strong>of</strong> oil<br />

and gas in the Gippsland Basin. Geochimica et Cosmochimica Acta 33:1183-<br />

1194.<br />

Brown, F.S., Baedeckner, M.J., Nissenbaum, A. and Kaplan, I.R. 1972.Early<br />

diagenesis in a reducing fjord, Saanich Inlet, British Coloumbia-III.Changes <strong>of</strong><br />

in<strong>organic</strong> constituents <strong>of</strong> sediments. Geochimica et Cosmochimica Acta<br />

36:1185-1203.<br />

Brown T.C. and Kenig F. 2004.Water column structure during deposition <strong>of</strong> middle<br />

Devonian–Lower Mississipian black and green/gray shales <strong>of</strong> the Illinois<br />

basins: a biomarker approach. Paleogeography Paleoclimate Paleoecol<br />

215:59–85.<br />

Budzinski, H., Garrigues, P.H., Connan, J., Devillers, J., Domine, D., Radke, M. and<br />

Oudin, J.L. 1995.Alkylated phenanthrene distributions as maturity and origin<br />

indicators in crude oils and rock extracts. Geochimica et Cosmochimica Acta<br />

59.10:2043-2056.<br />

181


Burgoyne, T.W. and Hayes, J.M. 1998.Quantitative production <strong>of</strong> H2 by pyrolysis <strong>of</strong><br />

gas chromatographic effluents. Analytical Chemistry 70:5136-5141.<br />

Caldicott, A.B. and Eglinton, G. 1973.Surface waxes. In: Miller, L.P. (Ed.),<br />

Phytochemistry 3,In<strong>organic</strong> Elements and Special Groups <strong>of</strong> Chemicals, New<br />

York: Van Nostrand Reinhold.162.<br />

Canuel, E.A., Freeman, K. and Wakeham, S., 1997. Isotopic compositions <strong>of</strong> lipid<br />

biomarker compounds in estuarine plants and surface sediments. Limnological<br />

Oceanography 42.7:1570–1583.<br />

Carter, J.D., Barber, W. and Jones, G.P. 1963.The Geology <strong>of</strong> Parts <strong>of</strong> Adamawa,<br />

Bauchi and Bornu provinces in Northeastern Nigeria.Bulletin <strong>of</strong> the<br />

Geological Survey <strong>of</strong> Nigeria 30:109.<br />

Chabbi, A., Rumpel, C. and Kögel-Knabner, I. 2007.Stable carbon isotope signature<br />

and chemical composition <strong>of</strong> <strong>organic</strong> matter in lignite-containing mine soils<br />

and sediments are closely linked. Organic Geochemistry 38.6:835-844.<br />

Chakhmakhchev, A., Suzuki, M. and Takayama, K. 1997.Distribution <strong>of</strong> alkylated<br />

dibenzothiophenes in petroleum as tool for maturity assessments. Organic<br />

Geochemistry 26:483-490.<br />

Clark, R.C. and Blummer, M. 1967.Distribution <strong>of</strong> n-paraffins in marine organisms<br />

and sediment. Limnology and Oceanography 12:79-87.<br />

Collister, J.W., Lichtfouse, E., Hiesshima, G. and Hayes, J.M. 1994.Partial resolution<br />

<strong>of</strong> sources <strong>of</strong> n-alkanes in the saline portion <strong>of</strong> the Parachute Creek Member,<br />

Green River Formation (Piceance Creek Basin, Colorado). Organic<br />

Geochemistry 21:645-659.<br />

Connan, J., Bouroullec, J., Dessort, D. and Albrecht, P. 1986.The microbial input in<br />

carbonate-anhydrite facies <strong>of</strong> a sabkha palaeoenvironment from Guatemala: a<br />

molecular approach. Organic Geochemistry 10:29-50.<br />

Cranwell, P.A. 1974.Monocarboxylic acids in lake sediments:Indicators, derived from<br />

terrestrial and aquatic biota <strong>of</strong> paleoenvironmental trophic levels. Chemical<br />

Geology 14:1-4.<br />

Cranwell, P.A. 1977.Organic <strong>geochemistry</strong> <strong>of</strong> Cam Loch (Sutherland) sediments.<br />

182


Chemical Geology 20:205-221.<br />

Cranwell, P.A. 1984.Lipid <strong>geochemistry</strong> <strong>of</strong> sediments from Uptan Broad, a small<br />

productive lake. Organic Geochemistry 7:25-37.<br />

Cranwell, P.A., Eglinton, G. and Robinson, N. 1987.Lipids <strong>of</strong> aquatic organisms as<br />

potential contributors to lacustrine sediments. Organic Geochemisry 11:513-<br />

527.<br />

Curiale, J.A. 1995. Saturated and olefinic terrigenous triterpenoid hydrocarbons in a<br />

biodegraded Tertiary oil <strong>of</strong> northeast Alaska. Organic Geochemistry 23:177-<br />

182.<br />

Curiale, J.A. 2002. A review <strong>of</strong> the occurrences and causes <strong>of</strong> migration<br />

contamination in crude oil. Organic Geochemistry 33:1389–1400.<br />

Day, W. C. and Erdman, J. G. 1963. Ionene: a thermal degradation product <strong>of</strong><br />

carotene. Science 141:808.<br />

De Graaf, W., Sinninghe Damsté, J.S. and de Leeuw, J.W. 1992.Laboratory<br />

simulation <strong>of</strong> natural sulphurisation I.Formation <strong>of</strong> monomeric and oligomeric<br />

isoprenoid polysulphides with low temperature reactions <strong>of</strong> in<strong>organic</strong><br />

polysulphides with phytol and phytadienes. Geochimica et Cosmochimica<br />

Acta 56:4321-4328.<br />

De Grande, S.M.B., Aquino Neto, F.R. and Mello, M.R. 1993.Extended tricyclic<br />

terpanes in sediments and petroleum. Organic Geochemistry 20:1039-1047.<br />

De Leeuw J. W. and Sinninghe Damsté, J.S. 1990 Organic sulphur compounds and<br />

other biomarkers as indicators <strong>of</strong> palaeosalinity. In Geochemistry <strong>of</strong> Sulfur in<br />

Fossil Fuels (ed. W. L. Orr and C. M. White); ACS Symposium Series 429:<br />

417-443.<br />

Derenne, S., Largeau, C., Casadevall, E. and Connan, J. 1988.Comparison <strong>of</strong><br />

torbanites <strong>of</strong> various origins and evolutionary stages. Bacterial contribution to<br />

their formation. Cause <strong>of</strong> the lack <strong>of</strong> boryococcane in bitumens. Organic<br />

Geochemistry 12:43-59.<br />

Didyk, B.M., Simoneit, B.R.T., Brassel, S.C. and Eglinton, G. 1978.Organic<br />

geochemical indicators <strong>of</strong> paleoenvironmental conditions <strong>of</strong> sedimentation.<br />

Nature 272:216-222.<br />

Dongmann, G., Nurnberg, H.W., Forstel, H. and Wagener, K. 1974.On the enrichment<br />

183


<strong>of</strong> H2 18 O in the leaves <strong>of</strong> transpiring plants. Radiation Environ.Biophys. 11:41-<br />

52.<br />

Douglas, A.G. and Grantham, P. J. 1974.Fingerprint gas chromatography in the<br />

analysis <strong>of</strong> some native bitumens, asphalts and related substances.In:Tissot,B.,<br />

Bienner,F(Eds.), Advances in Organic Geochemistry 1973.Editions<br />

Technip,Paris:261-276.<br />

Duan, Y., Wen, Q.B., Zheng, G., Luo, B.J. and Ma, L. 1997. Isotopic composition and<br />

probable origin <strong>of</strong> individual fatty acids in modern sediments from Ruoergai<br />

Marsh and Nansha Sea, China. Organic Geochemistry 27.7/8:583-589.<br />

Duan, Y., Song, J.M., Cui, M.Z. and Luo, B.J. 1998.Organic geochemical studies <strong>of</strong><br />

sinking particulate material in China Sea area In:<strong>organic</strong> matter influxes and<br />

distributional features <strong>of</strong> hydrocarbon compounds and fatty acids. Science in<br />

China 41:208-214.<br />

Dumitrescu, M. and Brassell, S.C. 2006. Compositional and isotopic characteristics <strong>of</strong><br />

<strong>organic</strong> matter for the early Aptian Oceanic Anoxic Event at Shatsky Rise,<br />

ODP Leg 198. Palaeogeography, Palaeoclimatology, Palaeoecology<br />

235:168–191.<br />

Durand, B. 2003.A History <strong>of</strong> Organic Geochemistry.Oil and Gas and Technology-<br />

Rev., IFP 58.2:203-231.<br />

Dzou, L. I. P., Noble, R. A. and Senflte, J. T. 1995. Maturation effects on absolute<br />

biomarker concentrations in a suite <strong>of</strong> <strong>coal</strong>s and associated vitrinite<br />

concentrates. Organic Geochemistry 23:681–697.<br />

Eglinton, G. and Hamilton, R.J. 1963.The distribution <strong>of</strong> n-alkanes In: Chemical<br />

Plant Taxonomy, ed. London: T.Swain.Academic Press. 187-217.<br />

Ehinola, O.A., Ekweozor, C.M., Oros, D.R. and Simoneit, B.R.T. 2002.Geology,<br />

<strong>geochemistry</strong> and biomaker evaluation <strong>of</strong> Lafia-Obi <strong>coal</strong> Benue Trough,<br />

Nigeria. Fuel 81:219-233.<br />

Ekweozor, C.M. and Telnǽs, N. 1990.Oleanane parameters: verification by<br />

quantitative study <strong>of</strong> the biomarker occurrence in sediments <strong>of</strong> the Niger<br />

Delta.Organic Geochemistry 16:401-413.<br />

184


Ellis, V.O., De Baroos, A.M.A., De Baroos, A.B., Simoneit, B.R.T. and Cardoso, J.N.<br />

1997.Sesquiterpenoids in sediments <strong>of</strong> hypersaline lagoon: A possible algal<br />

origin. Organic Geochemistry 26.11-12:721-730.<br />

Eneogwe, C., Ekundayo, O. and Patterson, B., 2002. Source derived oleanenes<br />

identified in Niger Delta oils. Journal <strong>of</strong> Petroleum Geology 25: 83–95.<br />

Espitalié, J., Lar Porte, J.L.Madec, M., Marquis, F., Le Plat, P., Paulet, J. and Boutfen,<br />

A. 1977.Method rapide de caracterisation des roches meres de leur potential<br />

petrolier et de leur degree d’evolution. Revue l’Inst. Francais du Petrole<br />

32.1:23-42.<br />

Farrimond, P., Taylor, A. and Telnǽs, N. 1998.Biomarker maturity parameters. The<br />

role <strong>of</strong> generation and thermal degradation. Organic Geochemistry 29:1181-<br />

1197.<br />

Ficken, K.J., Li, B., Swain, D.L. and Eglinton, G. 2000.An n-alkane proxy for the<br />

sedimentary input <strong>of</strong> submerged/floating fresh water aquatic macrophytes.<br />

Organic Geochemistry 31.7-8:745-749.<br />

Fowler, M.G., Abolins, P. and Douglas, A.G. 1986.Monocyclic alkanes in Ordovician<br />

<strong>organic</strong> matter. Organic Geochemistry 10:815-823.<br />

Galimov, E.M. 2006.Isotope Organic Geochemistry.Organic Geochemistry 37:1200-<br />

1262.<br />

Gallegos, E.J. 1971. Identification <strong>of</strong> new steranes, terpanes and branched paraffins in<br />

Green River Shale by combined capillary gas chromatography. Analytical<br />

Chemistry, 43:1151-1160.<br />

Gebhardt, H. 2001.Calcareous nann<strong>of</strong>ossils from the Nkalagu Formation type locality<br />

(Middle Turonian to Coniacian, southern Nigeria): biostratigraphy and palaeoecologic<br />

implications. Journal <strong>of</strong> African Earth Sciences 32.3:391-402.<br />

Gonzalez-villa, F.J., Polvillo, O., Boski, T., Moura, D. and de Andres, J.R.<br />

2003.Biomarker patterns in a time-resolved Holocene/terminal Pleistocene<br />

sedimentary sequence fro the Guadiana river estuarine area (SW<br />

Portugal/Spain border). Organic Geochemistry 34:1601-1613.<br />

Grantham, P.J. and Douglas, A.G. 1980.The nature and origin <strong>of</strong> sesquiterpenoids in<br />

some Tertiary fossil resins. Geochimica et. Cosmochimica Acta 44:1801-1810.<br />

Greenwood, P.F., Leenher, J.A., McIntyre, C., Berwick, L. and Franzmann, P.D.<br />

185


2006. Bacterial biomarkers thermally released from dissolved <strong>organic</strong> matter.<br />

Organic Geochemistry 37:597-609.<br />

Grice, K., Schouten, S., Nissembaum, A., Charrach, J. and Sinninghe Damsté, J.S.<br />

1998.A remarkable paradox: Sulfurised freshwater algal (Botryococcus<br />

braunii) lipids in an ancient hypersaline euxinic ecosystem. Organic<br />

Geochemistry 28:195-216.<br />

Grice, K., Audino, M., Alexander, R., Boreham, C.J. and Kagi, R.I. 2001.Distribution<br />

and stable carbon isotopic compositions <strong>of</strong> biomakers in torbanites from<br />

different palaeogeographical locations. Organic Geochemistry 32:1195-1210.<br />

Haberer, R.M., Mangelsdorf, K., Wilkes, M. and Horsefield, B. 2006.Occurrence and<br />

paleoenvironmental significance <strong>of</strong> aromatic hydrocarbon biomarkers in<br />

Oligocene sediments from the Mallik 5L-38 Gas Hydrate Production Research<br />

Well (Canada). Organic Geochemistry 37:519-538.<br />

Hayes, T.M., Freeman, K.H., Popp, B. and Hoham, C.H. 1990.Compound specific<br />

isotope analyses: a novel tool for reconstruction <strong>of</strong> biogeochemical proceses.<br />

Organic Geochemistry 16:1115-1128.<br />

Hayes, J.M. 1993.Factors controlling 13 C contents <strong>of</strong> sedimentary <strong>organic</strong> compounds:<br />

Principles and evidence. Marine Geology 113:111-125.<br />

Hendrix, M.S., Brassel, S.C., Carrol, A.R. and Graham, S.A. 1995. Sedimentology,<br />

<strong>organic</strong> <strong>geochemistry</strong> and petroleum potential <strong>of</strong> Jurassic <strong>coal</strong> measures:<br />

Tarim, Junggar and Turpain basins, Northwest China. American Association <strong>of</strong><br />

Petroleum Geologists Bulletin 79:929-959.<br />

Hernandez, M.E., Mead, R., Peralba, M.C. and Jaffe, R. 2001.Origin and transport <strong>of</strong><br />

n-alkan-2-ones in a subtropical estuary: potential biomarkers for seagrassderived<br />

<strong>organic</strong> matter. Organic Geochemistry 32:21-32.<br />

Hilkert, A.W., Douthitt, C.B., Schluter, H.J. and Brand, W.A. 1999.Isotope ratio<br />

monitoring gas chromatography/mass spectrometry <strong>of</strong> D/H by high<br />

temperature conversion isotope ratio mass spectrometry. Rapid<br />

Communication in Mass Spectrometry 13:1226-1230.<br />

Hu, J., Peng, P., Jia, G., Fang, D., Zhang, G., Fu, J. and Wang, P. 2002. Biological<br />

markers and their carbon isotopes as an approach to the paleoenvironmental<br />

reconstruction <strong>of</strong> Nansha area, South China Sea, during the last 30 ka.<br />

Organic Geochemistry 33: 1197–1204.<br />

Huang, W.Y. and Meinschein, W. G., 1979. Sterols as ecological indicators.<br />

186


Geochimica et Cosmochimica 43:739–745.<br />

Huang, H. and Pearson, M.J. 1999.Source rock paleoenvironments and controls on the<br />

distribution <strong>of</strong> dibenzothiophenes in lacustrine crude oil, Bohai basin, Eastern<br />

China.Organic Geochemistry 30:1455-1470.<br />

Huang, Y., Lockheart, M.J., Collister, J.W. and Eglinton, G. 1995.Molecular and<br />

isotopic bio<strong>geochemistry</strong> <strong>of</strong> the Miocene clarkia Formation: hydrocarbons and<br />

alcohols. Organic Geochemistry 23:785-801.<br />

Huang, Y., Lockheart, M.J., Logan, G.A. and Eglinton, G. 1996.Isotope and<br />

Molecular evidence for the diverse origins <strong>of</strong> carboxylic acids in leaf fossils<br />

and sediments from the Miocene Lake clarkia deposit, Idaho, USA.Organic<br />

Geochemistry 24:289-299.<br />

Huang, Y., Street-Perrott, F.A., Perrott, R.A., Metzger, P. and Eglinton, G.<br />

1999.Glacial/interglacial environmental changes inferred from molecular and<br />

compound-specific δ 13 analyses <strong>of</strong> sediments from Sacred Lake, Mt.<br />

Kenya.Geochimica et Cosmochimica Acta 63:1383-1404.<br />

Huang, Y., Dupont, L., Sarnthein, M., Hayes, J.M. and Eglinton, G. 2000.Mapping C4<br />

plants input from the North West Africa into North East Atlantic sediments.<br />

Geochimica et. Cosmochimica Acta 65:3505-3513.<br />

Hughes, W.B., Holba, A.G. and Dzhou, L.I.P. 1995.The ratios <strong>of</strong> dibenzothiophene to<br />

phenanthrene and pristane to phytane as indicators <strong>of</strong> depositional<br />

environment and lithology <strong>of</strong> petroleum source rocks. Geochimica et<br />

Cosmochimica Acta 59:3581-3598.<br />

Hunt, J.M. 1991.Generation <strong>of</strong> gas and oil from <strong>coal</strong> and other terrestrial <strong>organic</strong><br />

matter. Organic Geochemistry 17:673-680.<br />

Hunt, J.M. 1996.Petroleum Geochemistry and Geology.Second Edition. New York:<br />

W.H. Freeman and Company.<br />

Hunt, J.M., Philp, R.P. and Kvenvolden, K.A. 2002.Early developments in petroleum<br />

<strong>geochemistry</strong>. Organic Geochemistry 33:1025-1052.<br />

Hutton, A., Daulay, B., Herudiyanto Nas, C., Pujobroto, A. and Sutarwan, H.<br />

1994.Liptinite in Indonesian Tertiary <strong>coal</strong>s. Energy and Fuels 8:1469-1477.<br />

Inaba, T., Suzuki, N., Hirai, A., Sekiguchi, K. and Watanabe, T. 2001. Source rock<br />

lithology prediction based on oil diacholestane abundance in the siliceousclastic<br />

Akita sedimentary basin, Japan.Organic Geochemistry 32.7: 877-890.<br />

187


Inbus, S.W. and McKirdy, D.M. 1993.Organic <strong>geochemistry</strong> <strong>of</strong> Precambrian<br />

sedimentary rocks. In: Engel, M.H., Macko, S.A., editors. Organic<br />

<strong>geochemistry</strong>. New York: Pergamon Press. 657-684.<br />

Isasken, G.H., Curry, D.J., Yeakel, J.D. and Jensen, A.I. 1998.Controls on the oil and<br />

gas potential <strong>of</strong> humic <strong>coal</strong>s. Organic <strong>geochemistry</strong> 29:23-44.<br />

Jauro, A., Obaje, N.G., Agho, M.O., Abubakar, M.B. and Tukur, A. 2007.Organic<br />

<strong>geochemistry</strong> <strong>of</strong> Cretaceous Lamza and Chikila <strong>coal</strong>s, upper Benue Trough,<br />

Nigeria.Fuel 86:520-532.<br />

Johannes, I., Kruusement, K. and Veski, R. 2007. Evaluation <strong>of</strong> oil potential and<br />

pyrolysis kinetics <strong>of</strong> renewable fuel and shale samples by Rock-Eval analyzer.<br />

Journal <strong>of</strong> Analytical and Applied Pyrolysis 79:183-190.<br />

Johns, R.B. 1986.Biological markers in sedimentary record. Amsterdam: Elsevier<br />

Science Publisher.<br />

Katz, B. J. 1983. Limitations <strong>of</strong> Rock-Eval pyrolysis for typing <strong>organic</strong> matter.<br />

Organic Geochemistry 4:195–199.<br />

Kagya, M.L.N. 1996. Geochemical characterization <strong>of</strong> Triassic petroleum source rock<br />

in the Mandawa basin, Tanzania. Journal <strong>of</strong> African Earth Sciences 23.1: 73-<br />

88.<br />

Kennicult, M.C. and Brooks, J.M. 1990.Unusual normal alkane distributions in<br />

<strong>of</strong>fshore New Zealand sediments. Organic Geochemistry 15:193-197.<br />

Killops, S.D. and Killops, V.J. 1993.An introduction to <strong>organic</strong> <strong>geochemistry</strong>. UK:<br />

Longman Group Ltd.<br />

Killops, S.D. and Frewin, N.L. 1994.Triterpenoids diagenesis and cuticular<br />

preservation. Organic Geochemistry 21:1193-1209.<br />

Killops, S.D., Funnel, R.H., Suggate, R.P., Sykes, R., Peters, K.E., Walters, C.,<br />

Woolhouse, A.D., Weston, R.J. and Boudou, J.P. 1998.Predicting generation<br />

and expulsion <strong>of</strong> paraffinic oil from the vitrinite-rich <strong>coal</strong>s. Organic<br />

Geochemistry 29:1-9.<br />

Killops, S.D. and Killops, V.J. 2005.Introduction to <strong>organic</strong> <strong>geochemistry</strong>. Second<br />

edition.U.K: Blackwell Publishing Limited.<br />

188


Kleemann, G., Poralla, K., Englert, G., Kjosen, H., Liaaen-Jensen, S., Neunlist, S. and<br />

Rohmer, M. 1990.Tetrahymanol from the phototrophic bacterium<br />

Rhodopseudomonas palustris: first report <strong>of</strong> a gammacerane triterpane from a<br />

prokaryote. Journal <strong>of</strong> General Microbiology 136:2551-2553.<br />

Kogel-Knabner, I. 2002.The macromolecular <strong>organic</strong> composition <strong>of</strong> plant and<br />

microbial residues as inputs to soil <strong>organic</strong> matter. Soil Biology &<br />

Biochemistry 34:139-162.<br />

Köhler, J. and Clausing, A. 2000. Taxonomy and palaeoecology <strong>of</strong> din<strong>of</strong>lagellate<br />

cysts from Upper Oligocene freshwater sediments <strong>of</strong> Lake Enspel, Westerwald<br />

area, Germany. Review <strong>of</strong> Palaeobotany and Palynology 112:39-49.<br />

Kolonic, S., Sinninghe Damsté, J.S., Böttcher, M.E., Kuypers, M.M., Beckmann, B.,<br />

Scheeder, G. and Wagner, T. 2002.Geochemical characterization <strong>of</strong><br />

Cenomanian/Turonian black shales from the Tarfaya basin (SW Morocco).<br />

Journal <strong>of</strong> Petroleum Geology 25.3:325-350.<br />

Kotarba, M.J., Clayton, J.L., Rice, D.D. and Wagner, M. 2002.Assesment <strong>of</strong><br />

hydrocarbon source rock potential <strong>of</strong> Polish bituminous <strong>coal</strong>s and<br />

carbonaceous shales.Chemical Geology 184.1-2:11-35.<br />

Kvenvolden, K.A. 1962.Normal paraffin hydrocarbons in sediments from San<br />

Francisco Bay, Califonia.American Association <strong>of</strong> Petroleum Geologists<br />

Bulletin 46:1643-1652.<br />

Kvenvolden, K.A. and Squires, R.M. 1967.Carbon isotopic composition <strong>of</strong> crude oils<br />

from Ellenburger group (Lower Ordovician), Permian Basin west Texas and<br />

eastern New Mexico.American Association <strong>of</strong> Petroleum Geologists Bulletin<br />

5:1293-1303.<br />

Kwan-Hwa, S., Jun-Chin, S., Ying-Ju, C. and Wuu-Liang, H. 2006.Generation <strong>of</strong><br />

hydrocarbon gases and CO2 from a humic <strong>coal</strong>: Experimental study on the<br />

effect <strong>of</strong> water, minerals and transition metals. Organic Geochemistry 37:437-<br />

453.<br />

Langford, F.F. and Blanc-Valleron, M.M. 1990. Interpreting Rock Eval pyrolysis data<br />

using graphs <strong>of</strong> pyrolyzable hydrocarbons versus total <strong>organic</strong> carbon.<br />

American Association <strong>of</strong> Petroleum Geologists Bulletin 74:799-804.<br />

Lehtonen, K. and Ketola, M. 1990.Occurrence <strong>of</strong> long chain acyclic methyl ketones<br />

in sphagnum and carex peats <strong>of</strong> various degrees <strong>of</strong> humification. Organic<br />

Geochemistry 15:275-280.<br />

189


Leif, R.N. and Simoneit, B.R.T. 1995.Ketones in hydrothermal petroleums and<br />

sediments extracts from Guaymas Basin, Gulf <strong>of</strong> Califonia.Organic<br />

Geochemistry 23:889-904.<br />

Lewan, M.D., Winters, J.C. and Mc Donald, J.H. 1979.Generation <strong>of</strong> oil-like<br />

pyrolysates from <strong>organic</strong> rich shales. Science 203:897-899.<br />

Lewan, M.D. 1985.Evaluation <strong>of</strong> petroleum generation by hydrous pyrolysis<br />

experimentation. Philosophical Transactions Royal Society, London A<br />

315:124-134.<br />

Lis, G.P., Schimmelmann, A. and Mastalerz, M. 2006.D/H ratios and hydrogen<br />

exchangeability <strong>of</strong> type-II kerogens with increasing thermal maturity. Organic<br />

Geochemistry 37:342-353.<br />

Littke, R., Leythaeuser, D., Radke, M. and Schaefer, R. G. 1990. Petroleum<br />

generation and migration in <strong>coal</strong> seams <strong>of</strong> the Carboniferous Ruhr Basin,<br />

northwest Germany. Organic Geochemistry 16:247–258.<br />

Liu, W. and Huang, Y. 2005.Compound specific D/H ratios and molecular<br />

distributions <strong>of</strong> higher plant leaf waxes as novel paleoenvironmental indicators<br />

in the Chinese Loess Plateau.Organic Geochemistry 36:851-860.<br />

Lockheart, M.J., van Bergen, P.F. and Evershed, R.P. 2000.Chemotaxonomic<br />

classification <strong>of</strong> fossil leaves from the Miocene Clarkia lake deposit, Idaho,<br />

USA based on n-alkyl lipid distributions and principal component analyses.<br />

Organic Geochemistry 31:1223-1246.<br />

Logan, G.A. and Eglinton, G. 1994.Bio<strong>geochemistry</strong> <strong>of</strong> the Miocene lacustrine<br />

deposit at Clarkia, nothern Idaho, USA.Organic Geochemistry 21:857-870.<br />

Lu, H., Peng, P. and Sun, Y. 2003.Molecular and stable carbon isotopic composition<br />

<strong>of</strong> monomethylalkanes from one oil sand sample: source implications.<br />

Organic Geochemistry 34:745-754.<br />

Lücke, A., Helle, G., Schleser, G.H., Figueral, I., Mosbrugger, V., Jones, T.P. and<br />

Rowe, N.P. 1999.Environmental history <strong>of</strong> the German Lower Rhine<br />

Embayment during the Middle Miocene as reflected by carbon isotopes in<br />

brown <strong>coal</strong>s. Paleogeography, Paleoclimatology, Paleoecology 154:339-352.<br />

Mackenzie, A.S., Brassell, S.C., Eglinton, G. and Maxwell, J.R. 1982.Chemical<br />

fossils-the geological fate <strong>of</strong> steroids. Science 217:491-504.<br />

190


Mackenzie, A.S. 1984. Application <strong>of</strong> biomarkers in petroleum <strong>geochemistry</strong>. In:<br />

Brooks, J., Welte, D.H. (Eds.), Advancesin Petroleum Geochemistry, vol. 1.<br />

London: Academic Press.115–214.<br />

Macko, S.A. 1994.Compound-specific approaches using stable isotopes. In: Laytha,<br />

K., Michener, R.H. (Eds.), Stable Isotopes and Environmental Sciences.<br />

Oxford: Blacwell.241-247.<br />

Mallory, F.B., Gordon, J.T. and Connon, R.L. 1963.The isolation <strong>of</strong> a pentacyclic<br />

triterpenoid alcohol from a protozoan. Journal <strong>of</strong> the American Chemical<br />

Society 85:1362-1363.<br />

Mario, G.G., Ronald, C.S. and Milton, L.L. 1997.Generation and expulsion <strong>of</strong><br />

petroleum and gas from Almond formation <strong>coal</strong>, Greater Green river basin,<br />

Wyoming. American Association <strong>of</strong> Petroleum Geologists Bulletin 81:62-81.<br />

Mastalerz, M. and Schimmelmann, A. 2002.Isotopically exchangeable <strong>organic</strong><br />

hydrogen in <strong>coal</strong> relates to thermal maturity and maceral composition.<br />

Organic Geochemistry 33:921-931.<br />

Mathews, D.E. and Hayes, J.M. 1978.Isotope-ratio-monitoring gas chromatographymass<br />

spectrometry. Analytical Chemistry 50:1465-1473.<br />

Matsumoto, G.I. and Watanuki, K. 1990.Geochemical features <strong>of</strong> hydrocarbons and<br />

fatty acids in sediments <strong>of</strong> the inland hydrothermal environments <strong>of</strong><br />

Japan.Organic Geochemistry 29:1921-1952.<br />

Meinschein, W.G. 1961.Significance <strong>of</strong> hydrocarbons in sediments and petroleum.<br />

Geochimica et. Cosmochimica Acta 22:58-64.<br />

Mello, M.R., Gallianone, P.C., Brassel, S.C., Maxwell, J.R. 1988.Geochemical and<br />

biological marker asesment <strong>of</strong> depositional environments using Brazillian<br />

<strong>of</strong>fshore oils. Marine and Petroleum Geology 3:205-223.<br />

Meyers, P.A. 1989.Sources and deposition <strong>of</strong> <strong>organic</strong> matter in Cretaceous passive<br />

margin deep-sea sediments: A synthesis <strong>of</strong> <strong>organic</strong> geochemical studies from<br />

Deep Sea Drilling Project Site 603 outer Hatteras Rise.Marine and Petroleum<br />

Geology 6:182-189.<br />

Meyers, P.A. 1994. Preservation <strong>of</strong> source identification <strong>of</strong> sedimentary <strong>organic</strong><br />

matter during and after deposition. Chemical Geology 144:289–302.<br />

191


Meyers, P.A. 2003. Application <strong>of</strong> Organic Geochemistry to paleolimnological<br />

reconstructions: a summary <strong>of</strong> examples from Laurentian Great<br />

Lakes.Organic Geochemistry 34.2:261-289.<br />

Meyers, P.A., Bernasconi, S.M. and Forster, A. 2006. Origins and accumulation <strong>of</strong><br />

<strong>organic</strong> matter in expanded Albian to Santonian black shale sequences on the<br />

Demerara Rise, South American margin. Organic Geochemistry 37: 1816–<br />

1830.<br />

Miranda, A.C.M.L, Loureiro, M.R.B. and Cardoso, J.N. 1999.Aliphatic and Aromatic<br />

hydrocarbons in Candiota col samples: novel series <strong>of</strong> bicyclic compounds.<br />

Organic Geochemistry 30:1027-1028.<br />

Mitterer, R.M. 1993.The diagenesis <strong>of</strong> proteins and amino acids in fossil shells. In:<br />

Engel, M.H., Macko, S.A., editors:Organic <strong>geochemistry</strong>, principles and<br />

applications. New York: Plenum Press.739-753.<br />

Moldowan, J.M., Seifert, W.K. and Gallegos, E.J. 1983.Idenification <strong>of</strong> an extended<br />

series <strong>of</strong> tricyclic terpanes in petroleum. Geochimica et. Cosmochimica Acta<br />

47:1531-1534.<br />

Moldowan, J.M., Seifert, W.K. and Gallegos, E.J. 1985.Relationship between<br />

petroleum composition and depositional environment <strong>of</strong> petroleum source<br />

rocks. American Association <strong>of</strong> Petroleum Geologists Bulletin 69:1255-1268.<br />

Moldowan, J.M., Dahl, J., Huizinga, B.J., Fago, F.J., Hickey, L.J., Peakman, T.M. and<br />

Taylor, D.W. 1994.The molecular fossil record <strong>of</strong> oleanane and its relation to<br />

angiosperms. Science 265:768-771.<br />

Moumouni, A., Obaje, N.G., Nzegbuna, A.I. and Chaanda, M.S. 2007. Bulk<br />

geochemical parameters and biomarker characteristics <strong>of</strong> <strong>organic</strong> matter in<br />

two wells (Gaibu-1 and Kasade-1) from the Bornu basin: Implications on the<br />

hydrocarbon potential. Petroleum Science and Engineering 58.1: 275-282.<br />

Murray, A.P., Summons, R.E., Boreham, C.J. and Dowling, L.M. 1994.Biomarker<br />

and normal alkane isotope pr<strong>of</strong>iles <strong>of</strong> Tertiary oils: Relationship to source rock<br />

depositional setting. Organic Geochemistry 22:521-542.<br />

Nao, K., Jun-ichiro, H., Chun-Zhu, L., Chirag, S. and Tadatoshi, C. 2004.Evidence <strong>of</strong><br />

polycondensed aromatic rings in a Victorian brown <strong>coal</strong>. Fuel 83:97-107.<br />

192


Norgate, C. M., Boreham, C. J. and Wilkens, A. J. 1999. Changes in hydrocarbon<br />

maturity indices with <strong>coal</strong> rank and type, Buller Coalfield, New Zealand.<br />

OrganicGeochemistry 30:985–1010.<br />

Nyt<strong>of</strong>t, H.P., Bojesen-Koefed, J.A., Christiansen, F.G. and Fowler, M.G. 2002.<br />

Oleanane or lupane? Reappraisal <strong>of</strong> the presence <strong>of</strong> oleanane in Cretaceous-<br />

Tertiary oils and sediments. Organic Geochemistry 33:1225-1240.<br />

Nyt<strong>of</strong>t, H.P., Lutnaes, B.F. and Johansen, J.E. 2006.28-Nor-spergulanes, a novel<br />

series <strong>of</strong> rearranged hopanes. Organic Geochemistry 37:772-786.<br />

Obaje, N.G. 1994.Coal petrography, micr<strong>of</strong>ossils and palaeoenvironment <strong>of</strong><br />

Cretaceous <strong>coal</strong> measures in the Middle Benue Trough <strong>of</strong> Nigeria.Tübinger<br />

Mikropal. Mitt. 11:1-50.<br />

Obaje, N.G. 2000.Biomarker evaluation <strong>of</strong> the oil-generative potential <strong>of</strong> <strong>organic</strong><br />

matter in Cretaceous strata from the Benue Trough, Nigeria.Nigerian<br />

Association <strong>of</strong> Petroleum Explorationists Bulletin 15:1.<br />

Obaje, N.G., Wehner, H., Scheeder, O., Abubakar, M.B. and Jauro, H.<br />

2004.Hydrocarbon prospectivity <strong>of</strong> Nigeria’s inland basins: from the view<br />

point <strong>of</strong> <strong>organic</strong> <strong>geochemistry</strong> and <strong>organic</strong> petrology. American Association <strong>of</strong><br />

Petroleum Geologists Bulletin 87:325-353.<br />

Olvella, M.A., Gorchs, R. and de las Heras, F.X.C. 2006.Origin and distribution <strong>of</strong><br />

biomarkers in the sulphur rich Utrillas <strong>coal</strong> basin-Teruel minning district-<br />

Spain. Organic <strong>geochemistry</strong> 37:1727-1735.<br />

Orr, W.L. 1974.Changes in sulphur content and isotopic ratios <strong>of</strong> sulphur during<br />

petroleum maturation. Study <strong>of</strong> Big Horn Basin Paleozoic oils. American<br />

Association <strong>of</strong> Petroleum Geologists Bulletin 58:2295-2318.<br />

Otto, A., Walther, H. and Puttmann, W. 1997.Sesqui- and diterpenoid biomarkers in<br />

Taxodium-rich Oligocene oxbow lake clay in the Weiselster Basin,<br />

Germany.Organic Geochemistry 26:105-115.<br />

Otto, A. and Simoneit, B.R.T. 2001.Chemosystematics and diagenesis <strong>of</strong> terpenoids<br />

in fossil conifer species and sediment from the Eocene Zeitz formation,<br />

Saxony, Germany.Geochim et Cosmochim Acta 65:3505-3527.<br />

Otto, A. and Wilde, V. 2001.Sesqui-, di-, and triterpenoids as chemosystematics<br />

markers in extant conifers-a review. Botanical Review 67:141-238.<br />

193


Otto, A., Simoneit, B.R.T. and Rember, W.C. 2005.Conifer and angiosperm<br />

biomarkers in clay sediments and fossil plants from the Miocene clarkia<br />

Formation, Idaho, USA.Organic Geochemistry 36:907-922.<br />

Otto, A. and Simpson, M.J. 2006.Sources and composition <strong>of</strong> hydrolysable aliphatic<br />

lipids and phenols in soils from western Canada.Organic Geochemistry<br />

37:385-407.<br />

Ourrison, G., Albrecht, P. and Rohmer, M. 1979. The hopanoids: palaeo-chemistry<br />

and biochemistry <strong>of</strong> a group <strong>of</strong> natural products. Pure and Applied Chemistry<br />

51:709-729.<br />

Ourisson, G., Albrecht, P. and Rohmer, M. 1982.Predictive microbial biochemistryfrom<br />

molecular fossil to prokaryotic membranes. Trends in Biochemistry<br />

Sciences 7:236-239.<br />

Ozcelik, O. and Altunsoy, M. 2005.Organic Geochemical Characteristics <strong>of</strong> Miocene<br />

Bituminous Units in the Beypazari Basin, Central Anatolia, Turkey.The<br />

Arabian Journal <strong>of</strong> Science and Engineering 30.2A: 181-194.<br />

Pearson, M.J. and Obaje, N.G. 1999.Onocerane and other triterpenoids in Late<br />

Cretaceous sediments from the Upper Benue Trough, Nigeria: tectonic and<br />

palaeoenvironmental implications. Organic Geochemistry 30:583-592.<br />

Peters, K.E. and Moldowan, J.M. 1991.Effects <strong>of</strong> source, thermal maturity, and<br />

biodegradation on the distribution and isomerization <strong>of</strong> homohopanes in<br />

petroleum. Organic Geochemistry 17:47-61.<br />

Peters, K.E. and Moldowan, J.M. 1993.The Biomarker Guide.Interpreting Molecular<br />

Fossils in Petroleum and Ancient Sediments.New Jersey: Prentice-Hall,<br />

Englewood Cliffs.<br />

Peters, K.E. 2000.Petroleum tricyclic terpanes: predicted physiochemical behaviour<br />

from molecular mechanics calculations. Organic Geochemistry 31:497-507.<br />

Peters, K.E. and Fowler, M.J. 2005.Applications <strong>of</strong> Petroleum <strong>geochemistry</strong> to<br />

Exploration and Reservoir management. Organic Geochemistry 33:5-36.<br />

Peters, K.E., Walters, C.C. and Moldowan, J.M. 2005.The Biomarker<br />

Guide.Biomarkers and Isotopes in the Environment and Human History (I).<br />

Cambridge University Press.<br />

Peters, S.W. 1977.Middle-Cretaceous paleoenvironments and biostratigraphy <strong>of</strong> the<br />

Benue Trough, Nigeria.Geological Society <strong>of</strong> American Bulletin 89:151-154.<br />

194


Peters, S.W. 1982.Central West African Cretaceous-Tertiary benthic foraminifera and<br />

stratigraphy. Palaeontographica Abteilung A179: 1-104.<br />

Peters, S.W. and Ekweozor, C.M. 1982.Petroleum geology <strong>of</strong> Benue Trough and<br />

southeastern Chad Basin, Nigeria.American Association <strong>of</strong> Petroleum<br />

Geologists Bulletin 66:1141-1149.<br />

Petersen, H.I., Ansbjerg, J., Bojesen-Koefoed, J.A. and Nyt<strong>of</strong>t, H.P. 2000.Coalgenerated<br />

oil: source rock evaluation and petroleum <strong>geochemistry</strong> <strong>of</strong> the Lulita<br />

oil field, Danish North Sea.Petroleum Geology 23:55-90.<br />

Petersen, H.I., Nyt<strong>of</strong>t, H.P. 2006.Oil generation capacity <strong>of</strong> <strong>coal</strong>s as a function <strong>of</strong> <strong>coal</strong><br />

age and aliphatic structure. Organic Geochemistry 37:558-583.<br />

Philippi, G.T. 1965.On the depth, time and mechanism <strong>of</strong> petroleum generation.<br />

Geochimica et Cosmochimica Acta 29:1021-1049.<br />

Philp, R.P. 1985. Methods in Geochemistry and Geophysics, 23.Fossil Fuel<br />

Biomarkers.Applications and spectra. NY: Elsevier Science Publishing<br />

Company Inc.<br />

Philp, R.P. and Gilbert, T.D. 1986. Biomarker distributions in Australian oils<br />

predominantly derived from terrigenous source material. Organic<br />

Geochemistry 10: 73-84.<br />

Poinsot, J., Adam, P., Trendel, J.M., Connan, J. and Albrecht, P. 1995.Diagenesis <strong>of</strong><br />

higher plant triterpenes in evaporitic sediments. Geochimica et Cosmochimica<br />

Acta 59:4653-4661.<br />

Popp, B.N., Parekh, P., Tilbrook, B., Bidigare, R.P. and Laws, E.A. 1997.Organic<br />

carbon δ 13 C variations in sedimentary rocks as chemostratigraphic and<br />

paleoenvironmental tools. Palaeogeography, Palaeoclimatology,<br />

Palaeoecology 132:119-132.<br />

Powell, T.G. and McKirdy, D.M. 1973.Relationship between ratio <strong>of</strong> pristane to<br />

phytane, crude oil composition and geological environment in<br />

Australia.Nature 243:37-39.<br />

Powell, T.G. and Boreham, C.J. 1994.Terrestrially sourced oils: where do they exist<br />

and what are our limits <strong>of</strong> knowledge? In: Scott, A.C., Fleet, A.J. (Eds.), <strong>coal</strong><br />

and <strong>coal</strong>-bearing strata as oil prone source rocks? Geological society,<br />

London special publication 77:11-30.<br />

195


Prahl, F.G. and Wakeham, S.G. 1987.Calibration <strong>of</strong> unsaturation patterns in long<br />

chain ketone compositions for palaeotemperature assessment. Nature 330:<br />

367-369.<br />

Püttmann, W. and Villar, H. 1987.Occurrence and geochemical significance <strong>of</strong><br />

1,2,5,6-tetramethylnaphthalene. Geochimica et Cosmochimica Acta 51:3023-<br />

3029.<br />

Püttmann, W. and Bracke, R. 1995.Extractable <strong>organic</strong> compounds in the clay mineral<br />

sealing <strong>of</strong> a waste disposal site. Organic Geochemistry 23:43-54.<br />

Radke, M., Welte, D.H. and Wilsch, H. 1982.Geochemical study on a well in the<br />

Western Canada Basin: relation <strong>of</strong> the aromatic distribution pattern to maturity<br />

<strong>of</strong> <strong>organic</strong> matter. Geochimica et Cosmochimica Acta 46:1-10.<br />

Radke, M. and Welte, D.H. 1983.The methylphenanthrene index (MPI): a maturity<br />

parameter based on aromatic hydrocarbons. In: Bjoroy M, editor. Advances in<br />

Organic Geochemistry 10:51-63.<br />

Radke, M., Welte, D. and Willsch, H. 1986. Maturity parameters based on aromatic<br />

hydrocarbons: influence <strong>of</strong> the <strong>organic</strong> matter type. Organic Geochemistry<br />

10:51-63.<br />

Radke, M. 1987.Organic <strong>geochemistry</strong> <strong>of</strong> aromatic hydrocarbons. In: Brooks, J.,<br />

Welte, D.H. (Eds.). Advances in Petroleum Geochemistry, Vol.2.London:<br />

Academic Press.141-207.<br />

Radke, M., Welte, D.H. and Willsch, H. 1990.Methylated dicyclic and tricyclic<br />

aromatic hydrocarbons in crude oils from Handil Field (Indonesia). Organic<br />

Geochemistry 15:17-34.<br />

Radke, M., Horsefield, B., Littke, R. and Rullkotter, J. 1997.Maturation and<br />

Petroleum generation. In: Migration <strong>of</strong> hydrocarbons in sedimentary<br />

basins.B.Dohgez (Ed.). 649-665.<br />

Radke, M., Hilkert, A. and Rullkoter, J. 1998.Molecular stable carbon isotope<br />

compositions <strong>of</strong> alkylphenanthrenes in <strong>coal</strong>s and marine shales related to<br />

source and maturity. Organic Geochemistry 28.12: 785-795.<br />

Radke, M., Vriend, S. P. and Ramanampisoa, L. R. 2000. Alkyldibenz<strong>of</strong>urans in<br />

terrestrial rocks: Influence <strong>of</strong> <strong>organic</strong> facies and maturation. Geochimica et<br />

Cosmochimica Acta 64:275–286.<br />

196


Revill, A.T., Volkman, J.K., O’Leary, T., Summons, R.E., Boreham, C.J., Banks,<br />

M.R. and Denver, K. 1994.Hydrocarbon biomarkers, thermal maturity, and<br />

depositional setting <strong>of</strong> Tasmania, Australia.Geochimica et Cosmochimica Acta<br />

58:3803-3822.<br />

Requejo, A. G. 1994. Maturation <strong>of</strong> petroleum source rocks-II. Quantitative<br />

changes in extractable hydrocarbon content and composition associated with<br />

hydrocarbon generation. Organic Geochemistry 21:91–105.<br />

Rieley, G., Collier, R.J., Jones, D.M. and Eglinton, G. 1991.The bio<strong>geochemistry</strong> <strong>of</strong><br />

Ellesmere Lake, UK-I: source correlation <strong>of</strong> leaf wax inputs to the<br />

sedimentary lipid record. Organic Geochemistry 17:901-912.<br />

Rohmer, M. 1987.The hopanoids, prokaryotic triterpenoids and sterol surrogates. In:<br />

Surface Structures <strong>of</strong> Microorganisms and their interactions with the<br />

Mammalian Host (Edited by E. Schriner et.al.). Proceeding <strong>of</strong> Eighteen<br />

Workshop Conference. Hocchst, Schloss Ringberg: 227-242.<br />

Rontani, J.F., Giral, R.J.P., Baillte, G. and Raphel, D. 1992.”Bound” 6,10,14trimethylpentadecan-2-one:<br />

a useful marker for photodegradation <strong>of</strong><br />

chlorophylls with a phytol ester group in sea water. Organic Geochemistry<br />

18:139-142.<br />

Rooney, M.A., Vuletich, A.K. and Griffith, C.E. 1998.Compound-specific isotope<br />

analysis as a tool for characterizing mixed oils: an example from West <strong>of</strong><br />

Shetlands area. Organic Geochemistry 29:241-254.<br />

Rullkötter, J., Spiro, B., and Nissenbaum, A. 1985.Biological marker characteristics<br />

<strong>of</strong> oils and asphalts from carbonate source rocks in a rapidly subsiding graben,<br />

Dead Sea, Israel.Geochimica et Cosmochimica Acta 49:1357-1370.<br />

Rullkötter, J., Peakman, T.M. and ten Haven, H.L. 1994. Early diagenesis <strong>of</strong><br />

terrigenous triterpenoids and its implications for petroleum <strong>geochemistry</strong>.<br />

Organic Geochemistry 21:215-233.<br />

Sabel, M., Bechtel, A., Püttmann, W. and Hoernes, S. 2005.Palaeoenvironment <strong>of</strong> the<br />

Eocene Eckfeld Maar lake (Germany): implications from geochemical<br />

analysis <strong>of</strong> the oil shale sequence. Organic Geochemistry 36:873-891.<br />

Sache, D., Radke, J. and Gleixner, G. 2004.Hydrogen isotope ratios <strong>of</strong> recent<br />

lacustrine sedimentary n-alkanes record, modern climate variability.<br />

Geochimica et Cosmochimica Acta 68:4877-4889.<br />

Sachsenh<strong>of</strong>er, R.F., Kogler, A., Polesny, H., Strauss, P. and Wagreich, M. 2000a.The<br />

197


Neogene Fohnsdorf Basin: basin formation and basin inversion during lateral<br />

extrusion in the Eastern Alps.International Journal <strong>of</strong> Earth Sciences 89:415-<br />

430.<br />

Santos Neto, E.V. and Hayes, J.M. 1999.Use <strong>of</strong> hydrogen and carbon stable isotopes<br />

characterizing oils from the Potiguar Basin (onshore), northeastern<br />

Brazil.American Association <strong>of</strong> Petroleum GeologistsBulletin 83:496-518.<br />

Sari, A. and Bahtiyar, I. 1999.Geochemical evaluation <strong>of</strong> the Besikli oil field, Kahta,<br />

Adiyaman, Turkey.Marine and Petroleum Geology 16:51-164.<br />

Scalan, R.S. and Smith, J.E. 1970.An improved measure <strong>of</strong> the odd-even<br />

predominance in the normal alkanes <strong>of</strong> sediment extracts and petroleum.<br />

Geochimica et. Cosmochimica Acta 34:611-620.<br />

Schimmelmman, A., Lewan, M.D. and Wintsch, R.P. 1999.D/H isotope ratio <strong>of</strong><br />

kerogen, bitumen, oil, and water in hydrous pyrolysis <strong>of</strong> source rocks<br />

containing kerogen types I, II and III.Geochimica et. Cosmochimica Acta<br />

63:3751-3766.<br />

Schimmelmann, A., Sessions, A.L., Boreham, C.J., Edwards, D.S., Logan, G.A. and<br />

Summons, R.E. 2004.D/H ratios in terrestrially sourced petroleum systems.<br />

Organic Geochemistry 35:1169-1195.<br />

Schouten, S., Eglinton, T.E., Sinninghe Damsté, J.S. and de Leeuw, J.W.<br />

1995a.Influence <strong>of</strong> sulphur-crosslinking on the molecular size distribution <strong>of</strong><br />

sulphur-rich macromolecules in bitumen. In: Vairavamurthy M.A., Schoonen,<br />

M.A.A (Eds.), Geochemical Transformations <strong>of</strong> Sedimentary<br />

Sulphur.American Chemical Society symposium Series 612:80-92.<br />

Schouten, S., Hoefs, M.J.L. and Sinninghe Damste, J.S. 2000.A molecular and stable<br />

carbon isotopic study <strong>of</strong> lipids in late Quartenary sediments from the Arabian<br />

Sea.Organic Geochemistry 31:509-521.<br />

Schouten, S., Rijpstra, W.I.C., Kok, M., Hopmans, E.C., Summons, R.E., Volkman,<br />

J.K. and Sinninghe Damste, J.S. 2001.Molecular Organc tracers <strong>of</strong><br />

biogeochemical processes in a saline meromictic lake (Ace Lake). Geochimica<br />

et Cosmochimica Acta 65,p.1629-1640.<br />

Schwab, V.F. and Spangenberg, J.E. 2007. Molecular and isotopic characterization <strong>of</strong><br />

biomarkers in the Frick Swiss Jura sediments: A palaeoenvironmental<br />

reconstruction on the northern Tethys margin. Organic Geochemistry 38:419-<br />

439.<br />

198


Schwark, L. Vliex, M. and Schaeffer, P. 1998.Geochemical characterization <strong>of</strong> Malm<br />

Zeta laminated carbonates from the Franconian Alb, SW-Germany (II).<br />

Organic Geochemistry 29:1921-1952.<br />

Seifert, W.K. and Moldowan, J.M. 1980. The effect <strong>of</strong> thermal stress on source rock<br />

quality as measured by hopanes stereochemistry. In: Douglas, A.G., Maxwell,<br />

J.R. (Eds.), Advances in Organic Geochemistry 1979. New York: Pergamon.<br />

229-237.<br />

Seifert, W.K. and Moldowan, J.M. 1981.Palaeoreconstruction by Biological markers.<br />

Geochimica et Cosmochimica Acta 45:783-794.<br />

Seifert, W.K. and Moldowan, J.M. 1986.Use <strong>of</strong> biological markers in petroleum<br />

exploration. In: Methods in Geochemistry and Geophysics vol.24 (R.B. Johns,<br />

ed.). Amsterdam: Elsevier. 261-290.<br />

Sessions, A.L., Burgoyne, T.W., Schimmelmann, A. and Hayes, J.M.<br />

1999.Fractionation <strong>of</strong> hydrogen isotopes in lipid biosynthesis. Organic<br />

Geochemistry 30:1193-1200.<br />

Sessions, A.L. 2006.Isotope-ratio detection for gas chromatography. Journal <strong>of</strong><br />

Separation Science 29:1946-1961.<br />

Silverman, S.R. and Epstein, S. 1958.Carbon isotopic composition <strong>of</strong> petroleums and<br />

other sedimentary <strong>organic</strong> materials. American Association <strong>of</strong> Petroleum<br />

Geologists Bulletin 42:998-1012.<br />

Simoneit, B.R.T. 1977.Diterpenoid compounds and other lipids in deep sea sediments<br />

and their geochemical sigificance. Geochimica et Cosmochimica Acta 41:463-<br />

476.<br />

Simoneit, B.R.T., Mazurek, M.A., Brenner, S., Crisp.P.T. and Kaplan, I.R.<br />

1979.Organic chemistry <strong>of</strong> recent sediments from Guaymas Basin, Gulf <strong>of</strong><br />

Califonia.Deep-Sea Research 26A: 879-891.<br />

Simoneit, B.R.T. 1986.Cyclic terpenoids <strong>of</strong> the geosphere. In: Johns, R.B., editor.<br />

Biological markers in the sedimentary records. Amsterdam: Elsevier Science<br />

Publishers. 43-99.<br />

Simoneit, B.R.T., Grimalt, J.O., Wang, T.G., Cox, R.E., Hatcher, P.G. and<br />

Nissenbaum, A. 1986.Cyclic terpenoids <strong>of</strong> contemporary resinous plant<br />

detritus and <strong>of</strong> fossil woods, amber and <strong>coal</strong>. Organic Geochemistry 10:877-<br />

889.<br />

199


Simoneit, B.R.T., Cox, R.E. and Standley, L.J. 1988.Organic matter <strong>of</strong> the<br />

troposphere-IV: lipids in harmattan aerosols <strong>of</strong> Nigeria.Atmospheric<br />

Environment 22:983-1004.<br />

Simoneit, B.R.T., Sheng, G.Y., Chen, X.J., Fu, J.M., Zhang, J. and Xu, Y.P.<br />

1991.Molecular marker study <strong>of</strong> extractable <strong>organic</strong> matter in aerosols from<br />

urban areas <strong>of</strong> China.Atmospheric Environment 25A: 2111-2129.<br />

Simoneit, B.R.T. 1998.Biomarker PAHs in the environment. In: Neilson, A.,<br />

Hutzinger, O., editors. The handbook <strong>of</strong> environmental chemistry. Berlin:<br />

Springer Verlag. 175-221.<br />

Simoneit, B.R.T. 2002.Molecular Indicators <strong>of</strong> Past Life.The Anatomical Record<br />

268:186-195.<br />

Sinninghe Damsté, J.S., Englinton, T.I., de Leeuw, J.W. and Schenck, P.A.<br />

1989a.Organic sulphur in macromolecular sedimentary <strong>organic</strong> matter. Part<br />

I.Structure and origin <strong>of</strong> sulphur-containing moieties in kerogen, asphaltene<br />

and <strong>coal</strong> as revealed by flash pyrolysis. Geochimica et Cosmochimica Acta<br />

53:873-889.<br />

Sinninghe Damsté, J.S., Kenig, F., Koopmans, M.P., Koster, I., Schouten, S., Hayes,<br />

J.M. and De Leeum, J.W. 1995.Evidence for gammacerane as an indicator <strong>of</strong><br />

water coloumn stratification. Geochim et Cosmochim Acta 59:1895-1900.<br />

Smith, J.W., George, S.C. and Batts, B.D. 1994.Pyrolysis <strong>of</strong> aromatic compounds as a<br />

guide to synthetic reactions in sediments. American Petroleum Explorationists<br />

Association Journal 30:504-512.<br />

Sonzogoni, C., Bard, E., Rostek, F., Dollfus, D., Roselle-Melle, A. and Eglinton, G.<br />

1997.Temperature and salinity effects on the alkenone ratios measured in<br />

surface sediments from the Indian ocean. Quartenary Research 47:344-355.<br />

Stahl, W.J. 1977.Carbon and Nitrogen isotopes in hydrocarbn research and<br />

exploration. Chemical Geology 20:121-149.<br />

Stefanova, M., Magnier, C. and Velivona, D. 1995.Biomarker assemblage <strong>of</strong> some<br />

Miocene-aged Bulgarian lignite lithotypes. Organic Geochemistry 23:1067-<br />

1084.<br />

Stefanova, M., Oros, D.R., Otto, A. and Simoneit, B.R.T. 2002.Polar aromatic<br />

biomarkers in the Miocene Maritza-East lignite, Bulgaria.Organic<br />

Geochemistry 33:1079-1091.<br />

200


Stefanova, M., Markova, K., Marinov, S. and Simoneit, B.R.T. 2005. Molecular<br />

indicators for <strong>coal</strong>-forming vegetation <strong>of</strong> the Miocene Chukurovo lignite,<br />

Bulgaria. Fuel 84:1830–1838.<br />

Sternberg, L., De Niro, M.J. and Keeley, J.E. 1984.Carbon, hydrogen and oxygen<br />

isotope ratios <strong>of</strong> cellulose from plants having intermediates photosynthetic<br />

modes. Plant Physiology 74:104-107.<br />

Strachan, M.G., Alexander, R. and Kagi, R.I. 1988.Trimethylnaphthalenes in crude<br />

oils and sediments: Effects <strong>of</strong> source and maturity. Geochimica et<br />

Cosmochimica Acta 52:1255-1264.<br />

Summons, R.E., Brassel, S.C., Eglinton, G., Evans, E., Horodyski, R.J., Robinson, N.<br />

and Ward, D. M. 1988.Distinctive hydrocarbon biomarkers from fossiliferous<br />

sediment <strong>of</strong> the Late Proterozoic Walcott Member, Chuar Group, Grand<br />

Canyon, Arizona.Geochimica et. Cosmochimica Acta 52:2625-2637.<br />

Sun, Y., Sheng, G., Peng, P. and Fu, J. 2000. Compound-specific stable carbon<br />

isotope analysis as a tool for correlating <strong>coal</strong>-sourced oils and interbedded<br />

shale-sourced oils in <strong>coal</strong> measures: an example from Turpan basin, north-<br />

western China. Organic Geochemistry 31:1349-1362.<br />

Sykes, R., Snowdon, L.R. and Jahansen, P.E. 2004.Leaf biomass-a new paradigm for<br />

sourcing terrestrial oil <strong>of</strong> Taranaki basin. In: Boult, P.J., Johns, D.R., Lang,<br />

S.C., editors. Eastern Australia basin symposium II.Petroleum exploration<br />

society <strong>of</strong> Australia, Special Publication.553-574.<br />

Tanner, B.R., Uhle, M.E., Kelley, J.T. and Mora, C.I. 2007. C3/C4 variations in saltmarsh<br />

sediments: An application <strong>of</strong> compound specific isotopic analysis <strong>of</strong><br />

lipid biomarkers to late Holocene paleoenvironmental research. Organic<br />

Geochemistry 38: 474-484.<br />

Teichmuller, M. 1958.Metamorphism du carbon et propection du petrole. Review <strong>of</strong><br />

Industrial Minerals, Special Issue: 1-15.<br />

Ten Haven, H.L., de Leeuw, J.W., Rullkoter, J. and Sinninghe Damste, J.S.<br />

1987.Restricted utility <strong>of</strong> the pristane/phytane ratio as palaeoenvironmental<br />

indicator. Nature 330:641-643.<br />

Ten Haven, H.L. and Rulkotter, J. 1988.The diagenetic fate <strong>of</strong> taraxer-14-ene and<br />

oleanene isomers. Geochimica et Cosmochimica Acta 52:2543-2548.<br />

Thode, H.G. 1981.Sulfur isotope ratios in petroleum research and exploration:<br />

201


Williston basin. American Association <strong>of</strong> Petroleum Geologists Bulletin:<br />

1527-1535.<br />

Thomas, B. R. 1970. Phytochemical Phylogeny.In: Modern and Fossil Plant Resins.<br />

Academic Press. 59-79.<br />

Tissot, B., Pelet, R. and Roucache, J. 1977.Alkanes as geochemical fossils indicators<br />

<strong>of</strong> geological environments. In: Campus, R., Goni, J. (Eds.), Advances in<br />

Organic Geochemistry 1975,Enadimsa, Madrid. 117-154.<br />

Tissot, B.T. and Welte, D.H. 1984. Petroleum Formation and Occurrences.Second<br />

Edition. Berlin: Springer-Verlag.<br />

Treibs, A. 1934.The occurrence <strong>of</strong> Chlorophyll derivatives in an oil shale <strong>of</strong> the upper<br />

Triassic.Annalen 517:103-114.<br />

Tuo, J., Wang, X. and Chen, J. 1999.Distribution and evolution <strong>of</strong> tricyclic terpanes in<br />

lacustrine carbonates. Organic Geochemistry 30:1429-1435.<br />

Tuo, J. and Li, Q. 2005.Occurrence and distribution <strong>of</strong> long-chain acyclic ketones in<br />

immature <strong>coal</strong>s. Applied Geochemistry 20:553-568.<br />

Tuo,J. and Philp,R.P. 2005.Saturated and Aromatic diterpenoids and triterpenoids in<br />

Eocene <strong>coal</strong>s and mudstones from China. Applied Geochemistry 20:367-381.<br />

Tuo, J., Zhang, M., Wang, X. and Zhang, C. 2006.Hydrogen isotopic ratios <strong>of</strong><br />

aliphatic and diterpenoid hydrocarbons in <strong>coal</strong>s and carbonaceous mudstones<br />

from the Liaohe Basin, China.Organic Geochemistry 37:165-176.<br />

Tuo, J. Ma, W., Zhang, M. and Wang, X. 2007.Organic <strong>geochemistry</strong> <strong>of</strong> the<br />

Dongsheng sedimentary uranium ore deposits, China.Applied Geochemstry<br />

22:1949-1969.<br />

Tyson, R.V.1995. Sedimentary Organic Matter.In:Organic Facies and<br />

Palyn<strong>of</strong>acies.New York: Chapman and Hall.309-315.<br />

Van Aarssen, B.G.K., Cox, H.C., Hoogendoom, P. and de Leeuw, J.W. 1990.A<br />

cadinene biopolymer present in fossil and extant dammar resins as source for<br />

cadinanes and bicadinanes in crude oils from South East Asia.Geochimica et<br />

Cosmochimica Acta 54:3021-3031.<br />

Van Aarssen, B.G.K., Hessels, J.K.C., Abbink, O.A. and de Leeuw, J.W. 1992.The<br />

occurrence <strong>of</strong> polycyclic sesqui-, tri- and oligoterpenoids derived from<br />

resinous polymeric cadinene in crude oils from South East Asia.Geochimica et.<br />

202


Cosmochimica Acta 56: 1231-1246.<br />

Van Aarssen, B.G.K., Bastow, T.P., Alexander, R. and Kagi, R.I. 1999.Distribution <strong>of</strong><br />

methylated naphthalenes in crude oils: indicators <strong>of</strong> maturity, biodegradation<br />

and mixing. Organic Geochemistry 30:1213-1227.<br />

Van Aarssen, B.G., Alexander, R. and Kagi, R.I. 2000.Higher plant biomarkers reflect<br />

palaeovegetation changes during Jurassic times. Geochimica et Cosmochimica<br />

Acta 64:1417-1424.<br />

Van Kaam-Peters, H.M.E., Rijipstra, W.I.C., de Leeuw, J.W. and Sinninghe Damsté,<br />

J.S. 1998.A high resolution biomarker study <strong>of</strong> different lith<strong>of</strong>acies <strong>of</strong> <strong>organic</strong><br />

sulphur- rich carbonate rocks <strong>of</strong> a Kimmeridgian lagoon (French Southern<br />

Jura). Organic Geochemistry 28. (3-4): 151-177.<br />

Van Krevelen,D.W. 1961.Coal: Typology-chemistry-physics-constitiution.Amsterdam:<br />

Elsevier Science.<br />

Vlado and Valkovic 1983.Trace Elements in Coal I.Boca Raton, Florida, USA: CRC<br />

Press Inc.<br />

Volkman, J.K., Johnson, R.B., Gillian, F.T., Perry, G.T. and Bavor, H.J.<br />

1980.Microbial lipids <strong>of</strong> intertidal sediment-I. Fatty acids and hydrocarbons.<br />

Geochimica et Cosmochimica Acta 44:1133-1143.<br />

Volkman, J.K., Farrington, J.W., Gagosian, R.B. and Wakeham, S.G. 1983.Lipid<br />

composition <strong>of</strong> coastal marine sediments from the Peru upwelling region. In<br />

Advances in Organic Geochemistry, 1981,eds.M.Bjoroy et.al. Chichester:<br />

Wiley. 228-240.<br />

Volkman, J.K. and Maxwell, J.R. 1986.Acyclic isoprenoids as biological markers. In:<br />

Johns, R.B. (Ed.), Biological Markers in Sedimentary Record. Amsterdam:<br />

Elsevier .1-42.<br />

Volkman, J.K. 1988. The biological marker compounds as indicators <strong>of</strong> the<br />

depositional environments <strong>of</strong> petroleum source rocks. In: Fleet, A.J., Kelts, K.,<br />

Talbot, M.R. (Eds.), Lacustrine Petroleum Source Rocks. Geological Society<br />

Special Publication 40:103-122.<br />

Volkman, J.K., Barret, S.M., Blackburn, S.I. and Sikes, E.L. 1995.Alkenones in<br />

Gephyrocapsa oceanica: implications for studies <strong>of</strong> paleoclimate. Geochimica<br />

et Cosmochimica Acta 59:513-520.<br />

203


Volkman, J.K., Barrett, S.M., Blackburn, S.I., Mansour, M.P., Sikes, E.L. and Gelin,<br />

F. 1998. Microalgal biomarkers: a review <strong>of</strong> recent research developments.<br />

Organic Geochemistry 29:1163-1179.<br />

Volkman, J.K., Barrett, S.M. and Blackburn, S.I. 1999.Eustigmatophyte microalgae<br />

are potential sources <strong>of</strong> C29 sterols, C22-C28 n-alcohols and C28-C32 nalkyldiols<br />

in fresh water environments. Organic Geochemistry 30:307-318.<br />

Volkman, J.K., Rohjans, D., Rullkotter, J., Scholz-Bottcher, G. and Liebezeit, G.<br />

2000.Sources and diagenesis <strong>of</strong> <strong>organic</strong> matter in tidal flat sediments from the<br />

German Wadden sea. Continental Shelf Research 20:1139-1158.<br />

Volkman, J.K. 2003. Sterols in microorganisms. Applied Microbiology and<br />

Biotechnology 60:495-506.<br />

Volkman, J.K. 2005.Sterols and other triterpenoids: source specificity and evolution<br />

<strong>of</strong> biosynthetic pathways. Organic <strong>geochemistry</strong> 36:139-159.<br />

Wakeham, S.G., Schaffner, C. and Giger, W. 1980. Polycyclic aromatic hydrocarbons<br />

in recent lakesediments. II. Compounds derived from biogenic precursors<br />

during early diagenesis. Geochimica et Cosmochimica Acta, 44.13:415-429.<br />

Wang, P., Li, M. and Larter, S.R. 1996.Extended hopanes beyond C40 in crude oils<br />

and source rocks extracts from Liaohe Basin, N.E.China.Organic<br />

Geochemistry 24.5:547-551.<br />

Wang, R.L., Brassell, S.C., Scarpitta, S.C., Zheng, M.P., Zhang, S.C., Hayde, P.R. and<br />

Muench, L.M. 2004b.Steroids in sediments from Zabuye Salt Lake, Western<br />

Tibet: diagenetic, ecological or climatic signals? Organic Geochemistry<br />

35:157-168.<br />

Wang, T.G. and Simoneit, B.R.T. 1990.Organic <strong>geochemistry</strong> and <strong>coal</strong> petrology <strong>of</strong><br />

Tertiary brown <strong>coal</strong> in the Zhoujing mine, Baise basin, south<br />

China.2.Biomarker assemblage and significance. Fuel 69:12-20.<br />

Waples, D.W. and Machiara, T. 1990.Application <strong>of</strong> Sterane and Triterpane<br />

Biomarkers in Petroleum Exploration. Bulletin <strong>of</strong> Canadian Petroleum<br />

Geology 38:357-380.<br />

Welte, D.H., Hageman, H.W., Hollerbach, A., Leythauser, D. and Stahl, W.<br />

1975.Correlation between petroleum and source rock. Proceedings 9 th World<br />

Petroleum Congress.London: Applied Science Publishers II.179-191.<br />

204


Wen, Z., Ruiyog, W., Radke, M., Qingyu, W., Guoying, S. and Zhili, L. 2000.Retene<br />

in pyrolysate <strong>of</strong> algal and bacterial <strong>organic</strong> matter. Organic Geochemistry<br />

31:757-762.<br />

Wilkes, H., Ramrath, A. and Negendank, J. F. W. 1999. Organic geochemical<br />

evidence for environmental changes since 34,000 yrs BP from Lago di<br />

Mezzano, central Italy. Journal <strong>of</strong> Paleolimnology 22:349–365.<br />

Whiteman, A.J. 1982.Nigeria:Its Petroleum Geology, Resources and Potential, 1 and<br />

2.London: Graham and Trotman.394.<br />

Winters, J.C., Williams, J.A. and Lewan, M.D. 1983.A laboratory study <strong>of</strong> petroleum<br />

generation by hydrous pyrolysis. In: Bjorøy, M. et.al (Eds.), Advances in<br />

Organic Geochemistry 1981. Chichester: Willey.524-533.<br />

Xie, S., Yi, Y., Huang, J., Hu, C., Cai, Y., Collins, M. and Baker, A., 2003c. Lipid<br />

distribution in a subtropical southern China stalagmite as a record <strong>of</strong> soil<br />

ecosystem response to palaeoclimatic change. Quaternary Research 60: 340-<br />

347.<br />

Xiong, Y., Geng, A., Pan, C., Liu, D. and Peng, P. 2005.Characterization <strong>of</strong> the<br />

hydrogen isotopic composition <strong>of</strong> individual n-alkanes in terrestrial source<br />

rocks. Applied Geochemistry 20:455-464.<br />

Yangming, Z., Huanxin, W., Aiguo, S., Digang, L. and Dehua, P. 2005.Geochemical<br />

characteristics <strong>of</strong> Tertiary saline lacustrine oils in the Western Qaidam Basin,<br />

northwest China.Applied Geochemistry 33:1225-1240.<br />

Yessalina, S., Suzuki, N., Nishita, H. and Waseda, A. 2006.Higher plant biomarkers<br />

in Paleogene crude oils from the Yufutsu oil- and gas field and <strong>of</strong>fshore<br />

wildcats, Japan.Journal <strong>of</strong> Petroleum Geology 29.4:327-336.<br />

Younes, M.A. and Philp, R.P. 2005.Source rock characterization based on biological<br />

marker distributions <strong>of</strong> crude oils in the southern Gulf <strong>of</strong> Suez, Egypt.Journal<br />

<strong>of</strong> Petroleum Geology 28.3:301-317.<br />

Yuang, H. and Huang, Y. 2003.Preservation <strong>of</strong> lipid hydrogen isotope ratios in<br />

Miocene lacustrine sediments and fossils at Clarkia, northern Idaho,<br />

USA.Organic Geochemistry 34:413-423.<br />

Zaborski, P.M. 2000.The Cretaceous and Palaeocene trangression in Nigeria and<br />

Niger. Journal <strong>of</strong> Mineral Geology 36.2:153-173.<br />

205


APPENDIX<br />

206


Fig. 4.8 : m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Awgu samples showing the<br />

distribution <strong>of</strong> n-Alkanes.<br />

207


Fig. 4.8 (contd.): m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Awgu samples<br />

showing the distribution <strong>of</strong> n-Alkanes.<br />

208


Fig. 4.9 : m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu Formation samples<br />

(Okaba) showing the distribution <strong>of</strong> n-Alkanes.<br />

209


Fig. 4.9 (contd.): m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu Formation<br />

samples (Okaba) showing the distribution <strong>of</strong> n-Alkanes.<br />

210


Fig. 4.10 : m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu Formation samples<br />

(Onyeama) showing the distribution <strong>of</strong> n-Alkanes.<br />

211


Fig. 4.10 (contd.): m/z 85 Mass chromatograms <strong>of</strong> aliphatic fractions <strong>of</strong> Mamu Formation<br />

samples (Onyeama) showing distribution <strong>of</strong> n-Alkanes.<br />

212


Fig. 4.13: m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in Awgu<br />

samples (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

213


Fig 4.13 (contd.): m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in<br />

Awgu samples (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

214


Fig. 4.14: m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in Mamu<br />

samples (Okaba) (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

215


Fig. 4.15: m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in Mamu<br />

samples (Onyeama) (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

216


Fig. 4.15 (contd.): m/z 74 mass chromatogram showing the distribution <strong>of</strong> n-fatty acids in<br />

Mamu samples (Onyeama) (Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

(Numbers refer to carbon chain lengths <strong>of</strong> n-fatty acids).<br />

217


Fig. 4.16: m/z 58 mass chromatograms showing the distributions <strong>of</strong> alkan-2-ones in Awgu<br />

samples (Numbers refer to carbon chain lengths <strong>of</strong> alkan-2-ones).<br />

218


Fig. 4.17: m/z 58 mass chromatograms showing the distributions <strong>of</strong> alkan-2-ones in Mamu<br />

samples (Okaba) (Numbers refer to carbon chain lengths <strong>of</strong> alkan-2-ones).<br />

219


Fig. 4.18: m/z 58 mass chromatograms showing the distributions <strong>of</strong> alkan-2-ones in Mamu<br />

samples (Onyeama) (Numbers refer to carbon chain lengths <strong>of</strong> alkan-2-ones).<br />

220


Fig. 4.19 :m/z 191 showing the distribution <strong>of</strong> tricyclic and tetracyclic terpane in<br />

Awgu samples.<br />

221


Fig. 4.20: m/z 191 showing the distribution <strong>of</strong> tricyclic and tetracyclic terpane in<br />

Mamu Formation samples (Okaba).<br />

222


Fig. 4.21 : m/z 191 showing the distribution <strong>of</strong> tricyclic and tetracyclic terpane in<br />

Mamu Formation samples (Onyeama).<br />

223


Fig. 4.22: m/z 191 Mass chromatogram showing the distribution <strong>of</strong> hopanes in<br />

Awgu samples.<br />

224


Fig. 4.23 : m/z 191 Mass chromatogram showing the distribution <strong>of</strong> hopanes and<br />

benzohopanes in Mamu sample (Okaba).<br />

225


Fig. 4.24 : m/z 191 mass chromatogram showing the distribution <strong>of</strong> hopanes and<br />

benzohopanes in Mamu samples (Onyeama).<br />

226


Fig. 4.28 : m/z 217 mass chromatograms showing the distribution <strong>of</strong> steranes and<br />

diasteranes in Awgu samples.<br />

227


Fig. 4.29: m/z 217 mass chromatograms showing the distribution <strong>of</strong> steranes and<br />

diasteranes in Mamu samples (Okaba).<br />

228


Fig. 4.30 : m/z 217 mass chromatograms showing the distribution <strong>of</strong> steranes and diasteranes<br />

in Mamu samples (Onyeama).<br />

229


Fig. 4.34 : m/z 156, 170 mass chromatograms showing the distribution <strong>of</strong> naphthalene and<br />

alkylnaphthalenes in Awgu Samples.<br />

230


Fig. 4.34 (contd.): m/z 156, 170 mass chromatograms showing the distribution <strong>of</strong> naphthalene<br />

and alkylnaphthalenes in Awgu Samples.<br />

231


Fig. 4.35 : m/z 156, 170 mass chromatograms showing the distribution <strong>of</strong> naphthalene and<br />

alkylnaphthalenes in Mamu samples (Okaba).<br />

232


Fig. 4.36 : m/z 156, 170 mass chromatograms showing the distribution <strong>of</strong> naphthalene and<br />

alkylnaphthalenes in Mamu Samples (Onyeama).<br />

233


Fig. 4.37 : m/z 178, 192, 206 mass chromatograms showing the distribution <strong>of</strong> phenanthrene and<br />

alkylphenanthrenes in Awgu Samples.<br />

234


Fig. 4.38 : m/z 178, 192, 206 mass chromatograms showing the distribution <strong>of</strong> phenanthrene<br />

and alkylphenanthrenes in Mamu Samples (Okaba).<br />

235


Fig. 4.39 : m/z 178, 192, 206 mass chromatograms showing the distribution <strong>of</strong> phenanthrene and<br />

alkylphenanthrenes in Mamu Samples (Onyeama).<br />

236


Fig. 4.41: m/z 184,198,212,226 mass chromatograms showing the distributions <strong>of</strong><br />

dibenzothiophene and alkyldibenzothiophenes in Awgu samples.<br />

237


Fig. 4.41 (contd.): m/z 184,198,212,226 mass chromatograms showing the distributions <strong>of</strong><br />

dibenzothiophene and alkyldibenzothiophenes in Awgu samples.<br />

238


239

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!