You are on page 1of 56

Palaeoenvironmental analysis from organic

geochemistry of samples from the South Central


Pyrenees, Spain

James Crosby BSc Hons. Geochemistry

An organic and inorganic geochemical report submitted as part of the EART30100


Independent Mapping and Geochemistry Project in partial fulfilment of the degree of BSc
Hons. Geochemistry at the University of Manchester.

May 2016

Word Count: 8,822

1
Figure showing a top image of lithology within the mapped area as part of the project and a bottom image of soxhlet organic extraction as
part of the lab analysis aspect of the project

2
Abstract
In this study palaeoenvironment and source rock suitability of samples obtained from the
southern central Pyrenees have been determined based on isotopic, mineralogy, organic
and inorganic composition. Isotopic values of 13 and 18O have been investigated and
compared with typical composition to determine disequilibrium and organic input.
Mineralogy and inorganic information have been obtained using X-Ray diffraction.
Biomarkers (n-alkane, n-alcohol, n-fatty acid measured using GC-MS) have been utilised to
determine environment of deposition and subsequent post-depositional diagenesis. Bulk
water, organic and inorganic compositions have been determined using loss on ignition
experiment.

Palaeoenvironment deduction based on GC-MS and X-Ray Diffraction analysis of the


samples show an immature fluvial and marine environment of deposition based on
biomarker content in addition to n-fatty acid and n-alcohol distribution. Evidence from n-
alkane composition and distribution indicate post-depositional migration of a mature n-
alkane petroleum source into the samples.

Bulk organic analysis through isotopic equilibrium relationship, GC-MS and loss on ignition
shows varying organic quantities within samples. Samples showing the strongest isotopic
disequilibrium of 13C, highest total lipid extract and greatest bulk organic content indicate
greatest petroleum source rock potential.

3
1. Table of Contents

Section Page(s)
TITLE PAGE ...................................................................................................................................................... 1

FRONTISPIECE ................................................................................................................................................. 2

ABSTRACT ........................................................................................................................................................ 3

1.0 TABLE OF CONTENTS ................................................................................................................................. 4

2.0 INTRODUCTION ......................................................................................................................................... 5

3.0 REGIONAL GEOLOGICAL HISTORY ............................................................................................................. 6 - 9

4.0 SAMPLES .. 10-16

5.0 PREVIOUS WORK IN THE SPECIFIC AREA AND A REPORT OF DATA ALREADY EXISTS ............................... 17-18
5.1 X-Ray Diffraction Analysis ......................................................................................................................... 17
5.2 Stable Isotope Analysis ....................................................................................................................... 17
5.3 GC-MS Analysis ......................................................................................................................................... 17-18
7
5.4 Loss on Ignition Analysis ........................................................................................................................... 18

6.0 METHODOLOGY AND JUSTIFICATION OF METHODS AND ANALYSIS USED TO OBTAINED AND MANIPULATE
DATA................................................................................................................................................................19-23
6.1 X-Ray Diffraction Analysis ......................................................................................................................... 19
6.2 Stable Isotope Analysis .......................................................................................................................... 19-20
6.3 GC-MS Analysis ......................................................................................................................................... 20-22
6.4 Loss on Ignition Analysis ........................................................................................................................... 22-23

7.0 RESULTS AND ANALYSIS ............................................................................................................................24-37


7.1 X-Ray Diffraction Analysis ......................................................................................................................... 24-27
7.2 Stable Isotope Analysis ........................................................................................................................ 28-29
7.3 GC-MS Analysis ......................................................................................................................................... 30-36
7.4 Loss on Ignition Analysis ........................................................................................................................... 37

8.0 DISCUSSION ............................................................................................................................................. 38-42


8.1 Palaeoenvironment Deductions .............................................................................................................. 38-41
8.2 Source Rock Discussion ........................................................................................................................... 41-42

9.0 CONCLUSIONS ......................................................................................................................................... 43-45

10.0 ACKNOWLEDGEMENTS ........................................................................................................................... 46

11.0 REFERENCES ............................................................................................................................................ 47-52

12.0 APPRENDIX ..............................................................................................................................................52-54

Table 1.0 showing the table on contents

4
2. Introduction
This report is to investigate the depositional environment of samples obtained from the
southern central Pyrenees and what post-depositional diagenesis has affected their
composition. In addition, this report will also determine which of the samples collected from
this region present the greatest petroleum source rock potential.

The samples have been selected from various different locations within the stratigraphic
sequence in order to determine the palaeoenvironment, post-depositional influences and
petroleum source rock suitability of different depositional environments from varying
geological time periods.

The techniques used have been targeted to reveal a variety of characteristics from each
sample. X-Ray Diffraction provides fundamental mineralogical information which can
provide information about depositional environment. Stable isotope analysis of 13CVPDB
and 18OVPDB was a key aspect in determining source rock potential based on 13C
disequilibrium and post-depositional influence based on 18O disequilibrium. Gas
Chromatograph Mass Spectroscopy (GC-MS) revealed the organic composition within the
sample to enable determination of palaeoenvironment and post-depositional influences.
Finally, Loss on Ignition (LOI) was a key parameter is determining petroleum source rock
suitability.

The samples were obtained from various different locations and lithologies within
Cretaceous sediment of various lithologies. The samples were all obtained within the Pobla
basin and the Tremp basin of the Southern Pyrenean thrust wedge (Figure 3.0).

5
3. Regional Geological History
The mapped area is located within the Southern Central Pyrenees within the North-eastern
section of the Iberian Peninsula of Spain. This area was formed during the Pyrenean
orogeny, an asymmetrical, double-wedged continental belt formed by oblique convergence
and subduction. The Iberian plate subducted beneath the European plate as a result of this
convergence (ECORS-Pyrenees Team, 1988; Choukroune et al., 1989; Roure at al., 1989;
Munoz, 1992) from the Late Cretaceous to the Miocene (Puigde-fabregas and Qouquet,
1986).

The mapped area is located in the thrust front of the Boixols thrust which resulted from the
inversion in the Late Cretaceous of earlier extensional basins (Bond and McClay, 1995;
Munoz, 2002). The Boixols thrust is the northernmost of the three upper thrust sheets;
the other two being the Montasec and the Sierras Marginales which over-thrusted the
Palaeogene rocks of the Ebro Foreland basin (Seguret, 1972; Garrido, 1973; Berastegui et al.,
1993).The basins were part of a rift system developed in the Cretaceous by the divergence
of the Iberian and European plates creating a series of synrift marine sediment on the
hanging walls of the extensional faults in the basins of the south (Garcia-Senez, 2002).

Figure 3.0: Cross section through South Pyrenean Thrust belt to show different thrust sheets (Sinclair et al., 2005)

The mapped area observed geology can be broadly divided up into 3 main stratigraphic
units (Sim, 1986; Cuevas, 1992; Garca-Senz, 2002 and Beamud et al., 2003):

6
1. Post-folding conglomeritic deposits from the Oligocene and Quaternary.
2. Syn-folding quartzitic and calcareous sandstones, marls, clays and turbidites from the
Upper Santonian to Maastrichtian.
3. Pre-folding and post-rift calcarenites, limestones and marls from the Upper Cenomanian
to Lower Santonian.

The Sant Corneli Anticline dominates the central part of the mapped area. This area shows
Mesozoic succession of pre-, syn and postrift carbonates (Mencos et al., 2011). The
compressive regime began during the Late-Cretaceous (Santonian/Campanian) and resulted
in with the formation of thrusts carrying piggy-back basins. This allowed the creation of
accommodation space for sediment formation and formation of a piggy back foreland basin
(Pomar et al., 2005).

The oldest mapped formations of the Aramunt Vell and Congost formations represent
postrift sediments (Simo, 1986) arising from the post-rift of the Iberian and European plate
in the Cretaceous (Garcia-Senez, 2002). The continued opening of the Bay of Biscay (Reis,
1978) resulted with the formation of limestone outcropping within the Southern Pyrenees.

Amplification of the Boixols thrust during the Mid-Cretaceous led to the generation of this
thrust related anticline striking NE-SW (Tavani et al., 2011) and formation of syn-rift
sedimentary sequences, allowing the formation of the Vallacagra to the Tremp syn
compressional sedimentary sequences. The main folding stage was during the late
Cretaceous, when the forelimb reached a near-vertical attitude (Tavani et al., 2011).
Regional regression due to crustal shortening from Iberian convergence led to a transition
from Vallacagra clays and marls at the base of Sant Corneli into the Aren floodplane,
lagoonal and continental deposits (Nagtegaal et al., 1983) in the southern boundary with
third order cycles transgressions (Sim et al., 1986). The South of the mapped area shows
turbidite deposits and continental sands crossed bedding outlines a prograding shoreline
towards the North West (Nagtegaal et al., 1983). This lagoonal environment is confirmed
from the observation of rudists Hippurites castrol and Agriopleura moroi (Liebau, 1973;
Pons, 1977). Sedimentary dips decreasing from 40 to 12 up through the succession
represent tectonic synsedimentary growth with decreasing tectonic activity and relaxation
of crustal shortening of Sant Corneli Anticline.

The youngest observed lithology from the Oligocene and Quaternary show conglomerate
deposits beyond the north limb of the Sant Corneli Anticline. Unconformable dips of 20 are
recorded in the north of the mapping area, which exhibits an unconformable post-orogenic
nature (Tavini et al., 2011; Shackleton et al. 2011). These conglomerates represent more
than 20 alluvial fan lobes developed within an intramontane basin (Mellere, 1993).

7
Figure 3.1 Geological map of the Sant Corneli Anticline and the proximal areas obtained by previous geological investigation and research
(Mencos et al., 2011).

8
Figure 3.2: Geological map of area and cross section

9
4. Samples
The samples that were collected for the analysis were all deposited in the Southern Central
Pyrenees within close proximity to the mapped area. The analysed samples vary in
composition from mudstones to platform carbonates and vary in age from Albian to Eocene.

The oldest sampled formation titled the Albian formed approximately 113Ma during the
Albian. It is a marl sediment abundant in calcareous mud. This formation can be observed
directly below the Santa Fe formation and Congost Formations due to a time gap in
sedimentation related to a reactivation on inversion of the extensional fault system pushing
syn-extension wedge carbonates out of the half graben, showing an unconformity
representing the end of rifting (Bond & McClay, 1995). Formation shown to be over 700m
thick (Dinars-Turell, & Garcia-Senz, 2000). Sample obtained from the Flamisell River valley
along the L-503 located due north of La Pobleta de Bellvei and North East of the N-260.
GPS co-ordinates: 31IT 0329868 4685492

Sample titled the Pardina comes from the Cenomanian-Turonian Pardina Formation
(~94Ma) stratigraphically underlying the Reguard Formation in a dysoxic-anoxic conditions.
Within this formation abundant sponge spicules and nodular cherts have been observed
(Sanchez de la Torre, 2014). Microscope observations showed it with a grainstone texture
abundant in skeletal clasts and pyrite nodules. Sample obtained from obtained from the
Flamisell River valley along the L-503 located due north of La Pobleta de Bellvei and North
East of the N-260.
GPS co-ordinates: 3IT 0329832 4684861

Samples Reguard 1 & Reguard 4 belong to the Turonian-Coniacian Reguard Formation


(~88Ma). The Reguard Formation is the slope and basinal equivalent of the Congost
carbonate platform, into which it grades laterally and vertically. It is a thick continuous
limestone varying from very fine-grained wackestone up to a fine-grained packstone with
abundant biota including sponge spicules, planktonic foraminifera and shell debris. It is
identified as a sediment-starved, deeper-water facies (Booler & Tucker, 2002). Reguard 1
obtained from Flamisell River valley along the L-503 located due north of La Pobleta de
Bellvei and North East of the N-260.
GPS co-ordinates: 3IT 0329975 4684183
Regaurd 4 obtained from locality within 1km of Abella de la Conca.
GPS co-ordinates: 3IT 0342495 4669524

The sample Salas is a marl from the Salas marl Formation, deposited in the Upper
Campanian and lower Maastrichtian boundary approximately 75Ma. This marine deposit
contains abundant benthonic foraminifera fauna indicating a bathyal environment, such as
Chilostomella. This formation exhibits a true slope character with slumping and flowage
structures (Nagtegaal et al., 1983; Puigde-fabregas & Qouquet, 1986). Formation over 500m
thick (Nagtegaal, 1972). Sample obtained from Flamisell River valley along the L-503 located

10
due north of La Pobleta de Bellvei and North East of the N-260.
Co-ordinates: 3IT 0329842 4685102

The youngest sample analysed was the Montllobar mudstone, which belongs to the upper
Ager Group (45Ma) deposited during the Lower Eocene. The Ager comprises mudstone,
marls and marine limestones. The mudstone formed intermittedly as point bar deposits
(Meulen, 1986). The Montllobar sample was obtained alongC-1311 road due west of Figols
de Tremp.
Co-ordinates: N420917and E04815

When each sample was deposited, the Iberean plate would have been located in a mid
palaeolatitude.

Figure 4.0 Map of Southern Pyrenean area with locations for obtaining samples.

11
Montllobar

Salas

Pardina
Reguard

Albian

Figure 4.1 Sample age sequence of analysed samples with respect to a geological time scale. (Figure from The British Geological survey.
2012)

12
Figure 4.2: Stratigraphic column of the Albian marls and the relationship it had with surrounding lithology (Dinars-Turell & Garcia-Senz,
2000).

13
Figure 4.3 Stratigraphic column of the Sant Corneli sequence. Ages and relationship of Reguard (R) and Pardina (P) sample show (Caus et
al., 2013).

Figure 4.4: Stratigraphic column of Pobla de Segur Tremp area showing the Salas Marls age relationship. (Nagtegaal, 1972).

14
Figure 4.5. Stratigraphic column of Ypresian stratigraphy of the present-day Tremp-Graus and Ager Basins with a focus on the Ager Group
(Martinius, 2012).

Figure 4.6. Simple Stratigraphic column summarising age, lithology and fragments of each sample

15
16
C B G W P M

Clay
Silt
VF Sand
F Sand
M Sand
C Sand
Cong
5. Previous work in the specific area and a report of data that
already exists
5.1 X-Ray Diffraction Analysis
Comparison of mineralogy obtained in X-Ray Diffraction with various different papers allows
palaeoenvironment determination assuming insitu production. Work from Bairstow et al.,
2009 indicated Saponite forms in high pH (>9) environments, in addition work from Berner,
1970 shows pyrite forms in high organic lithology from sulphate reduction from bacteria.
Finally, Clinochlore is normally deposited in environment of high lacustrine inputs based on
information from Hillier, 1993. Comparison of mineralogy with this reference material
allows ideal deduction of palaoenvironment and environmental influences on inorganic
production.

5.2 Stable Isotope Analysis


There has been a wide range of analysis undertaken in Gradstein et al., 2004 showing
isotopic bulk rock values of 13C and 18O of global bulk rocks, brachiopods, belemnites,
planktonic foraminifera and trilobites from low latitudes from around the world of Cambrian
to Tertiary times. These provided ideal bulk rock comparison to samples to determine
equilibrium relationship and potential organic input. This is because work from Deines, 1980
indicates plants have a heavily depleted 13C which would result in disequilibrium.

5.3 GC-MS
Previous work undertaken by Sokolov et al., 1969; Shibaoka et al., 1973; Tissot et al., 1974;
Vassoevich et al., 1974; Dow, 1977 describe mechanisms of how hydrocarbons formed at
higher thermal alteration regimes can become expelled and rise into overlying more
immature sediment and become incorporated into the matrix. As a result, this can make the
immature overlying sediment display a mature n-alkane distribution. This work is key in
relating the immature fatty acid (methylester), n-alochol and other biomarker readings to
mature n-alkane distribution patterns.

Biomarker analysis of aromatic hydrocarbons by Hughes et al., 1995 provides key


information about phenanthrene showing it to form in sulphic carbonate rocks. In addition,
aromatic hydrocarbon analysis undertaken by Schoell et al., 1992 and Moldowan et al., 1984
showing norhopane formed from marine algal bacteria and bisnorhopane formed from
algae in a reducing environment were key sources of information to determine sediment
source and recreate palaeoenvironment.

Finally, biomarker analysis of triterpenes from J. D. Bu'Lock, 1989 shows compounds such as
botulin, oleane and lupanone to be indicative of higher plants and therefore derived from a
terrestrial environment. This investigation allows determination of sediment source and
palaeoenvironment recreation.

17
5.4 Loss of Ignition
Work undertaken by Chinn, 1991 provides information on average source rock organic
content for varying lithologies. In addition, work from Tissot et al., 1974 provides details on
kerogen characteristics and formation environments. Based on this obtained information
the total organic content % of each sample can be qualified as to whether or not they can
provide ideal source rocks and the potential environment of formation for the hydrocarbon
content.

18
6. Methodology and Justification of Methods and Analyses
Used to Obtain and Manipulate Data
The 6 samples used for analysis within this report were collected from various outcrops
within the Southern Central Spanish Pyrenees. To ensure minimal alteration to the samples
they were obtained from fresh, unexposed surfaces with minimal/no fracture planes using a
rock hammer under the supervision of Stefan Schroeder. Upon returning to Manchester
(UK) these samples were milled and dried and stored in containers in anticipation of
analysis.

6.1. X-Ray Diffraction Analysis:

Mineralogical determination was undertaken using X-Ray Powder Diffraction (XRD). By


obtaining inorganic data deductions can be made about formation environment, potential
alteration and metamorphic effect post-depositionally.

6.1.1. Preparation and analysis:


Samples further milled using pestle and mortar and treated with Amyl Acetate to
homogenise and adhese sample to 2cm diameter mount. Samples simultaneously analysed
for approximately 11 minutes each. Sample rotated from 5 - 70 and identification of
minerals obtained based on reference system within computer software. Peak shift adjusted
for based on presence and location of Calcite and Quartz.

6.1.2. Equipment:
X-Ray Powder Diffraction was undertaken by a Bruker D8Advance fitted with a Gbel mirror.
The machine ran at 40Kv and 40Ma emitting CuK radiation (0.15418 nm wavelength). It
has a step size of 0.0197 - 0.02, operating in room temperature and ambient humidity.
Approximately 0.1g of sediment was used per sample (The University of Manchester, 2013).

6.2. Stable Isotope Analysis:

13C and 18O analysis have been undertaken on the inorganic carbonate fraction of bulk
samples. The samples will then have their isotopic compositions compared to typical whole
rock data of mid-latitude isotopic compositions form Gradstein et al., 2004 to establish
whether they could have been in equilibrium with seawater. Mid-latitude values are used
because the samples were deposited on the Iberian plate at mid-latitude sea-levels. If they
are within 2 values of the sample it will be considered within equilibrium, however
outside of this value shows a significant organic input or diagenetic resetting; this is due to
13C of organic derived carbon being significantly lower, with C4 higher plants values
varying from -17 to -9 (average of 13%) and C3 plants show delta values ranging from -
32 to -20 (average value of -27) (Stable Isotope Ecology Laboratory, 1997) in
comparison to inorganic carbon values ranging from -4 to 4 since the Jurrasic
(Gradstein et al., 2004).

19
6.2.1. Methodology:
Undertaken by the University of Liverpool where powdered samples were weighed into
acid-cleaned glass tubes and CO2 for mass spectrometric measurement of carbon (13C/12C)
and oxygen (18O/16O) isotope ratios was prepared by reaction (to completion) of each
sample with phosphoric acid (specific gravity 1.91-1.92), at constant temperature (calcite at
25oC; dolomite at 60oC), following the classical sealed vessel method described by McCrea
(1950) and Swart et al. (1991). The evolved CO2 was recovered under vacuum using
standard cryogenic separation procedures (to remove H2O and N2) and the amounts of the
component masses (m/z 44, m/z 45, m/z 46) were measured with respect to a comparison
45/44 46/44
(reference) CO2. Resultant delta values (CO2 and CO2 ) were corrected for 17O
contributions (Craig, 1957) and calibration to the international Vienna Pee Dee Belemnite
(VPDB) carbon and oxygen isotope scales was achieved by measurement of CO2 prepared
(under identical conditions) from multiple aliquots of a laboratory quality control material
(Carrara marble calcite) with each batch of unknowns. Isotope compositions are reported as
conventional delta values ( 13CVPDB and 18OVPDB) with respect to VPDB, e.g.

nCO2 C n C n C n
13 12 13
C
12

CO2 (13 C) VPDB


CO 2 VPDB VPDB

n C n C
VPDB
13
VPDB
12

nCO2 (13 C) nCO2 (12 C)


where: and are, respectively, the amounts of 13C and 12C in sample CO2 in units of moles, and

similarly for VPDB


(nVPDB ) .

Oxygen values were subsequently adjusted for temperature-dependent kinetic isotope


fractionation effects associated with the (mineral-specific) carbonate-phosphoric acid
reaction using appropriate fractionation factors taken from published (Friedman & ONeil,
1977; Rosenbaum & Shepherd, 1986) sources.

Also obtained is wt % of carbonate content per sample obtained semi-quantitatively by


dissolving HCl to determine the weight.

6.2.2 Equipment:
Analysis was undertaken using a using a VG Sira 10 dual-inlet, stable isotope ratio mass
spectrometer. Analytical precision (1) is estimated to be better than 0.1 for both
isotope ratios based on the replicate measurement of well-characterized, laboratory
carbonate quality control materials.

6.3. Gas Chromatography Mass Spectrometry analysis:

The primary method used for organic geochemical analysis was Gas Chromatography - Mass
Spectrometry analysis. This method was selected because it allows the separation,
identification and quantification of complex mixtures of organic compounds. The Gas

20
Chromatogram separates the different compounds based on their affinity to the stationary
phase before the sample is injected in the Mass Spectrometer which identifies the
compound based on the mass/charge ratio.

6.3.1 Extraction and fractionation:


Methodology used adapted from Al Lawati et al, 2012. 4 samples (Montllobar, Pardina, Salas
and Albian) samples ranging in age from Eocene to Cretaceous were extracted using Soxhlet
apparatus with dichloromethane/methanol (DCM/MeOH, 2:1, v/v) for 24 hours. The total
lipid extracts (TLEs) were concentrated using rotary evaporation. The remaining dried lipids
were washed with DCM/MeOH, 2:1, v/v to transport lipids which were then filtered through
a column of glass wool and dilute NaSO4 crystals to remove remaining sediment and absorb
water. The samples were transferred into 3ml vial, blowdown and treated with activated (2
N HCl) copper to remove elemental sulfur. Internal standards of 540ng of tetracosane and
5400ng of 2-hexadefanol are then added. The sample were methylated by reacting with
200l of BF3 for 1 hour at 70C and MeOh to convert acids into methyl esters and filtered
through column of glass wool to remove very polar fraction. After the samples were Silyated
by addition of 25l with bis(trimethylsilyl)trifluroacetamide (BSTFA) and heated at 70C for 1
hour to convert alcohols to trimethylsilyl ethers. Samples were concentrated to improve
sensitivity under GC-MS through addition of hexane.

6.3.2 Equipment:
The Gas Chromatogram Mass Spectrometry model (GC/MS) was performed using a Aligent
5975C GC/MSD fitted with a programmable temperature vaporising (PTV), on-column
injector and can operates both in electron ionisation (EI) and chemical ionisation (CI) mode
(University of Manchester). A fused silica capillary column (30m x 0.12m) coated with CP-
Sil-5 was used with Helium carrier gas. The samples were injected at 50C and the oven was
programmed to 20 C/min to 130 and then 4 C/min to 320 C and held isothermally for the
remaining of the analysis. Compounds were identified by comparison of mass spectra and
retention times with those reported in the literature.

6.3.3. Calculations:
Methodology for Total n-alkane amount in sediment detection:
[ ]
Step 1: [ ] [ ] = X


Step 2: = Y

Step 3: =

Step 4: = Total n-alkane in sediment

Methodology for calculating Carbon Preference Index (CPI), (E. Schefub et al. 2003):

21
1 (() + ( + 2) + ( + 4) + ( + 6) + ( + 8) + + 10))
= ( )+
2 (( 1) + ( + 1) + ( + 3) + ( + 5) + ( + 7) + ( + 9))

1 (() + ( + 2) + ( + 4) + ( + 6) + ( + 8) + ( + 10))
( )
2 (( + 1) + ( + 3) + ( + 5) + ( + 7) + ( + 9) + ( + 11))
The Cn value range varied based on compound:
n-alkanes Cn = C23,
n-alcohols Cn = 13 .
n-Fatty acids Cn value varied based on sample: Montllobar Cn=15, Pardina Cn=17, Salas
Cn=15.

Methodology for calculating Average Carbon Chain Length (ACL), (E. Schefub et al. 2003):

()
=
()
Where: X = Area and i = Range from 21-31, however this can change depending on results
obtained from different samples.

Methodology for calculating 31/29+31, (Schefub, E et al. 2003):


(31)
31/29+31 = (29+31)

6.4 Loss on Ignition

The final technique used in for organic and inorganic analysis is loss on ignition. The purpose
of this analysis is because water, organic matter, carbonate material and siliciclastic content
can be sequentially measured after heating at selected temperatures. At 105C conditions
within the furnace will be hot enough to remove water from the sample due to the water
boiling point of 100C. Organic compound usually ignite at around 200C and upward,
therefore 550C conditions within the furnace are used to ignite organic carbon and remove
it from the sample. Inorganic materials such as carbonates are destroyed at significantly
higher tempretures of 700C-850C, therefore 1000C the environments within the furnace
are used to ignite inorganic carbon and remove it from the sample (Santisteban, 2004).

6.4.1 Methodology
The method applied is adapted from Santisteban, 2004. Ensure sample is completely dried
and milled into a fine powder and measure the weight of each crucible before it has had
sample added and after it has had sample added to it (WS), adding roughly 1g of sample into
the crucible. Add the crucible into the furnace at 105C for 2 hours and calculate the loss on
ignition at 105C (LOI105) as LOI105 = 100(WS - DW105)/WS, where WS is the weight of the air-
dried sample and DW105 is the dry weight of the sample heated at 105C. The next stage is
to heat the sample at 550C for 2 hours. Calculate LOI550 = 100(DW105 - DW550)/WS, where
LOI550 is the percentage of loss on ignition at 550C and DW550 is the weight of the sample

22
after heating at 550C. The final stage is to heat the sample to 1000C for 2 hours. Calculate
the LOI1000 = 100(DW550 DW1000)/WS, where LOI1000 is the percentage loss on ignition at
1000C and DW1000 is the weight of the sample after heating at 1000C. The remaining
sample after heating to 1000C is the residuum (LOIres).

This was coupled with GC-MS to calculate % kerogen content of each sample using the
calculation:
(%) = (%) (%)

5.4.1 Equipment
Samples were placed inside of circular ceramic bowls which were then placed inside of a
large glass hold and put into a Carbolite CWF 1200 model furnace for the allotted times
required for analysis.

23
7. Results and Analysis
7.1 X-Ray Diffraction
The following data was obtained by analysing all 6 samples using the X-Ray Diffraction
method.

The X-Ray Diffraction results displaying inorganic mineralogy content of the 6 samples varies
greatly amongst the samples.

Sample Mineralogy
Albian Quartz, Calcite, Muscovite
Pardina Quartz, Calcite
Reguard 1 Quartz, Calcite, Muscovite, Clinochlore, Pyrite
Reguard 4 Quartz, Calcite, Muscovite, Clinochlore, Pyrite
Salas Quartz, Calcite, Muscovite, Clinochlore
Montllobar Quartz, Calcite, Ankerite, Kaolinite, Saponite, Pyrite
Table 7.1 Dominant mineralogy of each sample

Quartz and Calcite are found in all samples. Calcite is the most abundant mineral in the
Albian, Pardina, Reguard 1, Reguard 4 and the Salas. Quartz is the most abundant mineral in
the Montllobar. Muscovite presence and abundance varies from each sample. Clinochlore is
found in Reguard 1, Reguard 4 and the Salas. Pyrite is found in the Reguard 1, Reguard 4 and
the Montllobar. Kaolinite is only found in the youngest sample; the Montllobar.

24
1800

PDF 00-046-1045 Si O2 Quartz, syn


PDF 00-005-0586 Ca C O3 Calcite, syn
PDF 01-076-0928 K ( Al1.5 Mg.5 ) ( Si3.5 Al.5 ) O10 ( O H )2 Muscovite-2M1
PDF 00-029-0701 ( Mg , Fe )6 ( Si , Al )4 O10 ( O H )8 Clinochlore-1MIIb, ferroan
1600

Salas
1400
1200
1000
Counts

800
600
400
200
0

10 20 30 40 50 60
2Theta (Coupled TwoTheta/Theta) WL=1.54060
1400

PDF 00-046-1045 Si O2 Quartz, syn


PDF 00-005-0586 Ca C O3 Calcite, syn
PDF 00-041-0586 Ca ( Fe +2 , Mg ) ( C O3 )2 Ankerite
1200

PDF 00-014-0164 Al2 Si2 O5 ( O H )4 Kaolinite-1A


PDF 00-030-0789 ( Mg2 Al ) ( Si3 Al ) O10 ( O H )2 4 H2 O Saponite-15A, aluminian
PDF 00-042-1340 Fe S2 Pyrite

Montllobar
1000
800
Counts

600
400
200
0

10 20 30 40 50 60
2Theta (Coupled TwoTheta/Theta) WL=1.54060

Figure 7.2.0 X-Ray Diffraction results of the Salas and Montllobar

25
2200
PDF 00-046-1045 Si O2 Quartz, syn
PDF 00-005-0586 Ca C O3 Calcite, syn
2000 PDF 01-076-0928 K ( Al1.5 Mg.5 ) ( Si3.5 Al.5 ) O10 ( O H )2 Muscovite-2M1
PDF 00-029-0701 ( Mg , Fe )6 ( Si , Al )4 O10 ( O H )8 Clinochlore-1MIIb, ferroan
PDF 00-042-1340 Fe S2 Pyrite
1800

Reguard 1
1600
1400
1200
Counts

1000
800
600
400
200
0

10 20 30 40 50 60
2Theta (Coupled TwoTheta/Theta) WL=1.54060
2200

PDF 00-046-1045 Si O2 Quartz, syn


PDF 00-005-0586 Ca C O3 Calcite, syn
2000

PDF 01-076-0928 K ( Al1.5 Mg.5 ) ( Si3.5 Al.5 ) O10 ( O H )2 Muscovite-2M1


PDF 00-029-0701 ( Mg , Fe )6 ( Si , Al )4 O10 ( O H )8 Clinochlore-1MIIb, ferroan
PDF 00-042-1340 Fe S2 Pyrite
1800

Reguard 4
1600
1400
1200
Counts

1000
800
600
400
200
0

10 20 30 40 50 60
2Theta (Coupled TwoTheta/Theta) WL=1.54060

Figure 7.2.1 X-Ray diffraction results of the Reguard 1 and Reguard 4.

26
2600 PDF 00-046-1045 Si O2 Quartz, syn
PDF 00-005-0586 Ca C O3 Calcite, syn

Pardina
2200
2000
1800
1600
1400
Counts

1200
1000
800
600
400
200
0

10 20 30 40 50 60
2Theta (Coupled TwoTheta/Theta) WL=1.54060
1800

PDF 00-046-1045 Si O2 Quartz, syn


PDF 00-005-0586 Ca C O3 Calcite, syn
PDF 01-076-0928 K ( Al1.5 Mg.5 ) ( Si3.5 Al.5 ) O10 ( O H )2 Muscovite-2M1
1600

Albian
1400
1200
1000
Counts

800
600
400
200
0

10 20 30 40 50 60
2Theta (Coupled TwoTheta/Theta) WL=1.54060

Figure 7.2.2 X-Ray Diffraction result of the Pardina and Albian.

27
7.2. Stable Isotope analysis

This data shows the variation in isotopic values through different samples deposited during
different ages in comparison with data collected for mid-latitude bulk rock at different ages
form Gradstein et al., 2004. The relationship between bulk standard isotopic values form
Gradstein et al., 2004 and the isotopic values vary greatly between samples.

All observed samples display a disequilibrium relationship in both the 13C or the 18O.

7.2.1 Predominant 13C disequilibrium:


The Salas and Montllobar samples show the greatest difference in 13C compared to all
other samples. The Salas has 3.4413C value in comparison to -1 to 113C for the
Campanian reference, showing a great difference. The Montllobar also shows a great
disequilibrium with a -2.7 13C in comparison to the 1 to 3 13C Eocene reference.

7.2.2 Predominant 18O disequilibrium:


Albian, Reguard 1, Reguard 4 and the Pardina samples show the greatest difference in
18O. The Albian has a -5.8318O in comparison to the Albian reference of -1 to 1
18O. The Reguard 1 has a -3.5518O and Reguard 4 has a -3.4418O in comparison to
the Turnonian reference of -1 to -318O. The Pardina has a -3.9818O in comparison to
the Turnonian reference of 1 to -318O.

Table 7.1.0 Isotopic data obtained from analysis of samples. Calcite notes obtained using XRD in section 6.1

Sample Location notes wt% carbonate approx error in carb 13CVPDB 18OVPDB
Montllobar Montllobar calcite 43 +/- 1wt% -2.70 -6.10
Albian Albian marls calcite 61 +/- 1wt% 1.99 -5.83
Reguard 1 Reguard 1 calcite 92 +/- 2wt% 2.90 -3.55
Reguard 4 Reguard 4 calcite 51 +/- 2wt% 2.50 -3.44
Salas Salas marls calcite 46 +/- 5wt% 3.44 -4.35
Pardina Pardina Lst calcite 67 +/- 1wt% 2.44 -3.98

Table 7.1.1 Bulk rock isotopic values obtained from Gradstein et al., 2004

Gradstein
Sample Name and Age Sample Age Gradstein 18O
13C
Montllobar Eocene 40Ma 1-3 0-3
Albian Albian 113Ma 13 -1 to +1
Reguard 1 Turnonian 90Ma 13 -1 to -3
Reguard 4 Turnonian 90Ma 13 -1 to -3
Salas Campanian 72Ma -1 to +1 -2 to -5
Pardina - Turnonian 94Ma 1- 3 -1 to -3

28
4.00

3.00

2.00

1.00

18O 0.00
-7.00 -6.00 -5.00 -4.00 -3.00 -2.00 -1.00 0.00 1.00 2.00
-1.00

Albian Reference Albian


-2.00
Eocene Reference Montllobar
Campanian Reference Salas -3.00

Turnonian Reference Reguard 1 & 4


-4.00
13C
Pardina

Figure 7.1. Global marine bulk-rock isotopic values for the study intervals following from Gradstein et al., 2004 for different ages from
table 7.1.0.

29
7.3. GC-MS analysis
The following data was obtained for the Albian, Montllobar, Pardina and Salas samples.

7.3.0. Total Lipid Fraction:


These calculations are of the n-alkanes, n-fatty acids and n-alcohols that have becomes
extractable in acidic solution. Separation and calculation allows relative portions of these
contents to be observed.

The total lipid extract varies between the different samples, with a range of 25.1mg to
15.2mg (table 6.3), with the Albian having the greatest lipid extract. The total lipid extract %
shows a range from 0.042% to 0.002%. There are 2 clear distinctions, lipid rich samples of
the Albian (25.1mg and 0.041%) and Salas (15.2g and 0.024%) and the lipid poor samples of
the Montllobar (4.1g and 0.0061%) and the Pardina (1.2g and 0.002%).

The bulk n-alkane extraction ranges from 12.1 -1.011g with a percentage variation of
84.25% n-alkane content of the Pardina to 36.59% of the Montllobar. The n-fatty acid
content varies from 6.6g to 0.17g with a percentage variation of 57.50% in the Salas to
14.17% in the Pardina. The n-alcohol bulk value ranges from 5.3g to 0.015g with a
percentage range of 21.12% in the Albian to 1.25% in the Pardina (Table 7.3).
The samples are all contain a large abundance and wide distribution of a homologous series
of n-alkanes, ranging from C14 C37. The most abundant homologue varies among the
samples; however it is most commonly C25 n-alkane in Albian, Montllobar and Salas (Table
6.3); the Pardina also shows a high abundance of C25 (Table 15.4 of Appendix). Based on CPI
values for n-alkanes of sediments analysed, there is no carbon chain length predominance
with values ranging from 1.27 2.93 (Figure 7.3.3). C31 ratios are similar within samples
ranging from 0.33-0.5, the oldest samples of Albian, Pardina and Salas exhibit a range of
0.33-0.36 while the younger Montllobar sample shows a higher value of 0.5. The average
chain length (ACL) varies widely among samples from 22.18 27.23. Observed in the
Montllobar, Pardina and Salas peaks are large areas of unresolved complex mixture.

The analysed samples show varying distributions of n-alcohol (Table 7.3); with the Pardina
(C12-C24) and Salas (C12-C18) showing an abundant and well distributed homologous series.
While samples of Pardina and Salas show a very limited abundance and distribution of n-
alcohols (figure 7.3.4). The CPI value that could be measured for the Albian shows a strong
predominance of even numbered homologues. The average chain length (ACL) of the n-
alcohols shows similar values with the Albian, Montllobar and Salas having a low value range
form 13.8-17.3. Cmax values vary between the samples with low value range of C13-C18.

All analysed samples show a homologous series distribution of n-fatty acids; with the
Montllobar (C12-C32), Pardina (C12-C29) and the Salas (C12-C33) showing the greatest
abundance and distribution (figure 7.3.5). The CPI values that could be obtained for certain
samples show low values ranging from 0.12-0.7, indicating a strong even-over-odd
predominance. Average chain lengths (ACL) for n-fatty acids show a range from 15.6-21.2.

30
The Montllobar, Pardina and Salas show similar values ranging from 20.0-21.2 while the
Albian shows a considerable lower ACL value of 15.

Hopanoids, Triterpenes and Aromatic Hydrocarbons at 191m/z:


Hopanoids and Tritenoids have been observed in various samples. The Albian sample
showed limited inputs of bisnorhopane. The Montllobar sample showed triterpenoid
biomarkers oleanene and Betulin. The Pardina showed limited inputs from fluorenone,
phenanthrene and anthracene. The Salas sample showed the most intense readings with
bisnorhopane, norhopane, lupanone, oleanol and phenanthrene.

31
Albian

Montllobar

Figure 7.3.0. Total Ion Chromatograms for total lipid fraction of sample: Albian and Montllobar. These illustrate overall differences in
biomarker compositions. , n-alkanes; +, n-fatty acids; , n-alcohols; , hopanes and triterpenes. Carbon chain length indicated by the
numbers.

32
Pardina

Salas

Figure 7.3.1. Total Ion Chromatograms for total lipid fraction of sample: Pardina and Salas. These illustrate overall differences in biomarker
compositions. , n-alkanes; +, n-fatty acids; , n-alcohols; , hopanes and triterpenes. Carbon chain length indicated by the numbers.

33
Montllobar

Albian

Figure 7.3.2 m/z 57 chromatogram for Montllobar and Albian samples for n-alkane; , distribution outlining no odd-over-even
predominance between CPI =1.83 of Montllobar in comparison to CPI = 2.93 value for Albian.

Albian

Pardina

Figure 7.3.3 m/z 103 chromatogram of Albian and Pardina samples for n-alcohol; , distribution

34
Salas

Albian

Figure 7.3.4 m/z 74 chromatogram of Salas and Albian sample for n-fatty acids; +, distribution.

Salas

Figure 7.3.5 m/z 191 chromatogram of Salas sample showing distribution of hopanes in the.

35
Table 7.3 Summary of total lipid extract (mg & %), n-alkane mass (mg) & %, n-fatty acid mass (mg) & %, n-alcohol mass (mg) & %, n-alkane
distribution through (Cmax, 31/29+31, ACL and CPI), n-fatty acid distribution (Cmax and ACL) and n-alcohol distribution (CL and CPI).

Mass of Sample Total Lipid Extract


Sample Total Lipid Extract (%)
(g) (mg)
Albian 60.39 25.1 0.04156
Montllobar 67.55 4.1 0.006069
Pardina 59.85 1.2 0.002039
Salas 62.85 15.2 0.02418
Total Lipid Extract Bulk n-alkane Bulk fatty acids Bulk n-
Sample
(mg) (mg) (methylesters) alcohols
Albian 25.1 12.1 6.6 5.3
Montllobar 4.1 1.5 2.16 0.44
Pardina 1.2 1.011 0.17 0.015
Salas 15.2 5.85 8.74 0.58
Fatty acids (methylester) n-Alcohol
Sample n-Alkane %
% %
Albian 48.21 26.29 21.12
Montllobar 36.59 52.68 10.73
Pardina 84.25 14.17 1.25
Salas 38.49 57.5 3.82
Sample n-Alkenes
Cmax 31/(29+31) ACL CPI
Albian C25 0.36 27.23 2.93
Montllobar C25 0.5 25.33 1.83
Pardina C17 0.33 22.18 1.27
Salas C25 0.37 24.1 2.42
Samples Fatty acids (methylesters) n-Alcohols
Cmax ACL CPI Cmax ACL CPI
Albian C15 15.6 N/A C18 17.3 0.185
Montllobar C16 21.2 0.12 C15 14.7 N/A
Pardina C22 20 0.23 C12 N/A N/A
Salas C22 20.5 0.7 C13 13.8 N/A

36
7.4 Loss on Ignition
The following data was obtained after all 6 samples were analysed to see the inorganic carbon
content and the organic carbon content.

7.4.1 Mass of water (%)


The mass of water within different samples vary greatly in the analysed sampled with a
range from 0.233% to 2.369%. The sample with the largest % water is the Salas with 2.369%
of water within the lithology while the sample with the lowest % water within its structure is
the Pardina with 0.239%.

7.4.2 Mass of Organic Carbon (%)


The mass of organic carbon within each sample varies greatly, with a range of 1.885%
organic carbon to 0.102%. The Salas (1.885%), Albian (1.744%) and Montllobar (1.540%)
have a very enriched organic carbon lithology while the Reguard 4 (0.522%), Reguard 1
(0.410%) and Pardina have a very organic carbon depleted lithology. The Reguard 1 and
Reguard 4 show a variation in organic content.

7.4.3 Mass of Inorganic Carbon (%)


The mass of Inorganic carbon within the analysed samples are very high, with a range of
42.374% to 17.196%. The highest value of 42.374% inorganic carbon is the Pardina sample,
with the Reguard 1 (38.095%) and Reguard 4 (36.063%) also showing high values. The most
depeated sample is the Montllobar with 17.196% inorganic carbon with Salas (25.010%) and
Albian (27.851%) showing intermediate values.

7.4.4 Kerogen total (%)


Kerogen content varies from 1.861% in the Salas to as low as 1.534% in the Montllobar. All
indicates a high kerogen content of each organic rich sample.
Table 7.4.0 Data from Loss on Ignition experiment that allow the Mass of Water (%), Mass of Organic Carbon (%) and Mass of Inorganic
Carbon (%) to be determined (See figure 14.5 in appendix for calculations)

Sample Mass of Water (%) Mass of Organic Carbon (%) Mass of Inorganic Carbon (%)
Reguard 1 0.544 0.410 38.095
Pardina 0.233 0.102 42.374
Salas 2.369 1.885 25.010
Albian 1.229 1.744 27.851
Montollobar 1.049 1.540 17.196
Reguard 4 0.648 0.522 36.063

Table 7.4.1 % Kerogen based on total organic carbon (TOC(%)) from table 9.4.0 and total lipid extract (TLE(%)) from table 9.3.0

Sample TOC (%) TLE (%) % Kerogen


Salas 1.885 0.024 1.861
Albian 1.744 0.042 1.702
Montllobar 1.54 0.006 1.534

37
8. Discussion
8.1 Palaeoenvironment deductions

8.1.0 X-Ray Diffraction Data


The Montllobar sample shows abundant mineralogy. Based on mineraology of dominant
calcite with less quartz and kaolinite it shows a lime-mud micrite sample indicative of high
biological activity. The Reguard 1, Reguard 4, Pardina and Albian also shows abundant
calcite with minor quartz and muscovite, indicative of a bioactive environment during
deposition. The Pardina and Albian show a mature sample with limited mineralogy. The
Salas sample shows a quartz dominated sample with less calcite and muscovite, indicating a
less bioactive zone and higher terrestrial input. The presence of Saponite in the Montllobar
indicates diagenesis of pH>9 and formed formed in alkaline conditions (Bristow et al., 2009),
however it is possible most Saponite is detrital and input from an external source due to
lack of evidence in biomarkers supporting this observation. The presence of pyrite in
Montllobar, Reguard 1 and Reguard 4 indicates diagenesis from an organic rich rock
produced via bacterial sulfate reduction (Berner, 1970). Data in GC-MS analysis and loss on
ignition analysis shows that these are organic poor samples; there this evidence suggests
that these once organic rich samples have undergone organic migration. The presence of
clinochlore in the Salas, Reguard 1 and Reguard 4 samples shows a strong lacustrine input
(Hillier, 1993). In addition, clinochlore is common low-grade metamorphic mineral; this is
likely to be from a detrital source and washed in from the axial zone of the Pyrenees which
has experienced high metamorphic activity.

8.1.1 Isotopic Data


Graph 7.1.0 shows the distribution of 13C and 18O signatures for the 6 analysed samples in
comparison with 13C and 18O bulk rock values during the time they were deposited. There
are 2 clear differences, with the Montllobar and and Salas showing a 13C disequilibrium
relationship, which; indicates significant organic input during deposition because of carbon
isotopes becoming lighter along the direction of natural gas migration (Zhang, 2013). During
formation of the Montllobar samples the inorganic bulk isotopic values would have matched
or been very similar to PDB (Kendall et al., 2004). However, algal and marine plants (C4) fix
carbon through the Calvin-Benson cycle (Chikaraishi & Naraoka, 2001). Carbon derived from
this process has a 13C of -6 to -19 (Deines, 1980). Incorporation of this organic carbon
would allow the Montllobar sample to exhibit the negative 13C disequilibrium nature it
shows. Therefore, this data indicates this sample has been diagentically altered from post-
depositionally from the initial environment through incorporation of organic carbon. In
addition, low 13C can be caused by alteration and oxidation of organic content within the
inorganic carbon content of each sample (Gross, 1964). The Salas displays a positive
equilibrium most likely through post depositional alteration from preferential degradation
of labile organic compounds (Hatch & Leventhal, 1997). The other samples of Reguard 1,
Reguard 4, Pardina and Albian show isotopic 18O signatures in disequilibrium to the 18O

38
bulk rock values of the Turonian and Albian reference. Although the 18O values of each the
sample is within equilibrium with minimal amount of organic input from C3 or C4 plant
input, there is evidence of thermal resetting in 18O. This resetting can occur due to an
increase in temperature or influx of fluids with different isotopic compositions (Garzione et
al., 2004). Therefore this alteration may have occurred through thermal resetting during
burial of the lithologys or from meteoric water influx into pores of lithology.

There are factors which may complicate the interpretation of isotopic 13C and 18O
signatures stated. There is the possibility that the sample has been altered due to
interaction with dissolved inorganic carbon from subsurface water which can have 113C
signatures of -5 to -25. In addition, the organic contents may be washed in from a
terrestrial environment containing C3 plants with 113C values of -24 to -34 (United
States Geological Survey, 2004) which fix carbon through the Hatch-Slack cycle (Chikaraishi
& Naraoka, 2001). However, to investigate this it would need to be verified with organic
carbon 13C analysis.

8.1.2 GC-MS
Organic geochemical analysis of the sediments revealed a predominance of n-alkanes, fatty
acid (methylesters) and n-alcohols. The n-alkanes show a low CPI ranging from 1.27 2.93
with no odd-over-even carbon chain length predominance (Table 7.3), indicating these
samples are mature (Bray & Evans, 1961). The n-fatty acids and n-alcohols where display a
strong even over odd predominance (Table 7.3), indicating a higher plant source. Long
chains of n-alkanes (C25-C35), n-fatty acids (C24-C28) and n-alcohols (C23-C34) are become
biosynthesised by higher plants as part of leaf wax (Eglinton and Hamilton, 1967; Tulloch,
1976). Abundant minor products have also been found such as triterpoids; oleananes and
lupanes indicative of higher plant material (Ekweozor & Udo, 1988), hopanes such as
bisnorhopane, norhopane indicative of bacteria from a reducing environment (Schoell et al.,
1992; Moldowan et al., 1984) and phenanthrene indicative of a high-sulfur carbonate source
rock (Hughes et al., 1995).

Based on these observations, these samples show that there has been a mature n-alkane
rich hydrocarbon source that has migrated into relatively immature, terrestrially derived
sediment. Kerogen and other oil products formed at deeper, higher thermal regimes and
are expelled from the source rock rise through overlying layers at varying rates depending
on porosity and permeability (AAPG, 2016) and deposit into sediment that has experiences
less thermal alteration (Sokolov et al., 1969; Shibaoka et al., 1973; Tissot et al., 1974;
Vassoevich et al., 1974; Dow, 1977). During this migration period the kerogen and other
hydrocarbon contents break down and defunctionalise to form smaller chain n-alkanes
(Largeau et al., 1986; Tegelaar et al., 1989; De Leeuw & Largeau, 1993). These n-alkanes
form with no odd or even chain length predominance and become incorporated into
sediments, therefore creating immature sediment with mature n-alkane contents.

39
In addition, the Salas, Pardina and the Montllobar chromatograms show a large area of
unresolved complex mixtures beneath the various peaks. As stated by Gough & Rowland
1991 hydrocarbons often appear as unresolved complex mixtures (UCMs) when examined
by techniques which are routinely successfully used to characterize other hydrocarbon
mixtures such as crude oils (e.g., gas chromatography (GC) or GC-mass spectroscopy (GC-
MS)). Therefore indicating these samples are likely to contain large amounts of unresolved
hydrocarbon mixtures, such as kerogen breakdown products. This high unresolved complex
mixture further indicates mature n-alkane migration into the samples.

The 4 samples are generally comparable in term of biomarker sources, showing


predominantly a terrestrial source showing an even-over-odd predominance of fatty acid
(methylesters), n-alcohols as well as triterpenoids from higher plants (Eglington and
Hamilton, 1963, 1967) and a limited amount of marine derived biomarkers. 2 samples, the
Albian and the Salas show very high total lipid extract (TLE) values (7.3.0) in comparison to
the Pardina and Montllobar. Firstly, this may be indicative of a higher biological productivity,
or, the Albian and Salas lithology may have more pore spacing in which the migrating n-
alkanes can become stored into. Alternatively, the Albian and Salas may have been
deposited stratigraphically more proximal to n-alkane source rocks, so the there was a
higher abundance available for absorption into the lithology. Another variation in bulk
results show the Albian and Pardina samples to show high n-alkane values and lower fatty
acid (methylester) and n-alcohol values in comparison to the Montllobar and Salas (Table
7.3). This is likely due to an age v degradation relationship. The Albian and the Pardina
samples are older that the Montllobar and Salas, therefore they have undergone more
deformation and thermal heating resulting with degradation and defunctionalisation. This
results with less fatty acids (methylesters) and n-alcohols and more abundance of the
migrated n-alkanes. This difference is reflected in figure 7.3.3 for reduced abundance and
distribution of n-alcohols for the Pardina and reduced fatty acid (methylesters) in figure
7.3.4 for the Albian.

There are also key differences in calculations based on homologous distribution of n-


alkanes, fatty acid (methylesters) and n-alcohols. The n-alkane distributions show 2 key
differences in samples between the Pardina and the Montllobar. The Pardina shows a lower
ALC and CPI than the other 3 samples (Table 7.3); this is likely related to the Pardina
showing a stronger oil GC-MS chromatogram signature (Figure 7.3.1) and more dominant in
the migrated n-alkane. This causes a more mature signature with the organic products being
more broken down through migration into shorter chained n-alkanes (Gonzales-Vila F, 1996)
and dominating the low TLE compound. The fatty acid (methyl ester) value ACL show Albian
having a significantly lower value of 15.6 in comparison to the Montllobar, Salas and Pardina
showing a range of 20-21.2. This is likely to be a because of the thermal degradation of fatty
acids to form n-alkanes, work from Shimoyama and Johns, 1972 shows that the breakdown
of fatty acids to n-alkanes are catalysed in the presence of calcite, such as shown in the
Albian.

40
8.1.3 Loss on Ignition
The Montllobar, Salas and Albian samples show high values for organic carbon % and water
content % (Table 7.4.0). These high organic values are indicative of high organic migration
and potential high biological productivity. The Reguard 1, Reguard 4 and Pardina samples
show low organic carbon % and water content% (Table 7.4.0) indicating either low biological
productivity or high organic migration.

8.2 Source rock discussion

Typically, source rocks are produced by settling of organic rich sediments in environments
which support biologic activity and production of large volumes of organic matter. These
depositional settings concentrate and preserve this organic matter over time (Jacobsen,
1991). As the sediment is buried it experiences greater geothermal temperatures causing
organic matter to transform into insoluble organic matter Kerogen. As heat increases the
kerogen evolves into petroleum and bitumen compounds. As the Kerogen continues to
evolve the initially complex compound undergoes structural simplification, loss of hydrogen
to produce oil, then wet gas and finally dry gas (McCarthy et al., 2011). Typically, average
ideal shale petroleum source rocks consist of 2.2% total organic carbon, calcareous shale
source rocks consist of 1.8% total organic carbon and carbonate source rocks consist of 0.7%
total organic carbon (Chinn, 1991).

Isotopic data of the Montllobar and Salas samples have the greatest 13C disequilibrium. The
Montllobar sample has the greatest negative equilibriu, indicating that this sample has the
greatest organic content (Graph 7.1.0). Therefore, based on these initial results these would
be the most potential to be source rocks, while the Reguard 1, Reguard 4 and Pardina
display little/no disequilibrium relationship and this low organic environment would be
undesirable for an organic source rock.
In addition, the chemostratigraphy of the bulk isotopic signatures of the 13C and 18O for
the Reguard 1 and Reguard 4 samples are very similar. Therefore this shows that this
method can be used as a viable lateral correlation tool in addition to sequence stratigraphy.

Results from the loss of ignition analysis would indicate based on work from Chinn, 1991
that the Salas, Albian and Montllobar would be the most ideal source rocks (table 6.4.0).
Salas values of 1.885% total organic carbon show it as a good calcareous shale source
rock, Albian values of 1.774% total organic carbon indicate it as a good calcareous shale
source rock and 1.540% total organic carbon indicate it as a good calcareous shale source
rock (AAPG, 2015). Low total organic carbon values for the Pardina, Reguard 1 and samples
indicate it to be a poor option for a petroleum source rock, while Reguard 4 poses a
marginally fair option (AAPG, 2015).

Based on calculations in Table 7.4.1 the Salas, Albian and Montllobar source rocks display
Kerogen rich, lipid poor organic contents. This is likely to relate to type III Kerogen formed

41
from terrigenous plant deposited in shallow to deep marine or non-marine environment
(Tissot et al., 1974).

However, for these source rocks to be economical, they require ideal structural orientation
to allow organic content from these samples to become trapped and ideal lithology for the
organic content of the lithologys to be stored in a reservoir rock and trapped by an
overlying impermeable non-porous lithology.

42
9. Conclusion
Due to the migration of the petroleum sourced n-alkanes the immature sediment has
become altered and contaminated, therefore palaeoenvironment deductions and
conclusions for the lithology cannot be made based on n-alkane composition and
distribution. Palaeoenvironment deductions will be based on fatty acid (methylester), n-
alcohol and biomarker data for each individual sample and organic activity determined by
isotopic relationships and loss on ignition data.

The Montllobar sample shows a mudstone sample deposited in intermitted symmetrical


chanel deposits (Meulen, 1986). Although Loss on Ignition data and isotopic (13C)
disequilibrium would suggest a high organic content, only a small fraction of it is deposited
as lipid n-alkane, fatty acid (methylester), n-alcohol and biomarker content. This organic
content suggests that this mudstone has had a high immature terrestrial input based on the
strong even over odd predominance of the fatty acid (methylesters) and low ACL for n-
alcohols. Biomarker content of Betulin and Oleane show triterpenoids indicative of higher
plants, indicating terrestrial environment formation (Bu'Lock, 1989). Low total lipid extract
(TLE) indicates low biological productivity. Based on this, the Montllobar shows a lacustrine
depositional system with strong terrestrial inputs from higher plants.

The Salas sample shows a mud grade, carbonate rich marl. This marine deposits contains
abundant benthonic foraminifera fauna indicating a bathyal environment (Nagtegaal et al.,
1983; Puigde-fabregas & Qouquet, 1986). The high 13C disequilibrium expresses a high
organic input which has become altered. Loss on ignition data also shows a high proportion
organic input with GC-MS data showing that a large proportion of this organic matter is
derived from lipid content. This organic content suggests that this mudstone has had a high
immature terrestrial input based on the strong even over odd predominance of the fatty
acid (methylesters) and low ACL for n-alcohols. Triterpene lupanone and oleanol biomarkers
were found indicating higher plant terrestrial input, while phenanthrene was also found,
indicating organic derivation from a high sulfur component carbonate (Hughes et al., 1995).
Biomarker norhopane, formed from marine algal bacteria was observed as well as
bisnorhopane, a marine biomarker formed from algae indicating a reducing environment
(Schoell et al., 1992; Moldowan et al., 1984). Therefore this sample shows a carbon formed
in marine bathyal conditions under reducing conditions with strong terrestrial higher plant
input. High total lipid extraction (TLE) also indicates high biological productivity.

The Reguard 1 and Reguard 4 limestones show a platform limestone with a packstone
texture dominated in skeletal fragments shelf (Booler & Tucker 2002). Loss on Ignition data
and isotopic (13C, 18O) equilibrium would suggest a low organic content of this sample.
The X-Ray Diffraction data indicates an immature sample with abundant mineralogical
compositions. This data suggests a reefal limestone with minimal n-alkane migration
potentially due to minimal pore spacing due to a muddy matrix. Isotopic analysis and X-Ray

43
Diffraction analysis and analysis indicate minimal lateral variation, whereas differences in
loss on ignition values indicate minor local influences on organic content.

The Pardina sample shows a laminated limestone platform with a grainstone texture
abundant sponge spicules and nodular cherts have been observed (Sanchez de la Torre,
2014). Loss on Ignition data and isotopic (13C, 18O) equilibrium would suggest a low organic
input with GC-MS confirming a low total lipid extract. This organic content suggests that this
mudstone has had a high immature terrestrial input based on the strong even over odd
predominance of the fatty acid (methylesters) and low ACL for n-alcohols. The Pardina
shows abundant aromatic hydrocarbons Fluorenone, Phenanthrene and Anthracene,
indicative of a mature hydrocarbon source. Phenanthrene is also indicative of a sulfur rich
carbonate source (Hughes et al., 1995). This sample therefore shows a carbonate formed in
marine conditions with a strong oil signature formed under high sulphur conditions. Low
total lipid extract (TLE) indicates low biological productivity.

The Albian sample shows a calcite dominated marl deposited stratigraphically abelow the
Congost and Santa Fe formation (Bond & McClay, 1995). Loss on Ignition data and isotopic
(13C, 18O) disequilibrium would suggest a high organic content. Loss on ignition data also
shows a high proportion organic input with GC-MS data showing that a large proportion of
this organic matter is derived from lipid content. This organic content suggests that this marl
has had a high immature terrestrial input based on the strong even over odd predominance
of the n-alcohols and low ACL for fatty acid (methylesters). Bisnorhopane was observed as a
bacterially derrived biomarker within the sample indicative of marine environments with a
reducing environment (Schoell et al., 1992; Moldowan et al., 1984). Based on this, the
Albian shows formation in a marine environment under reducing conditions with a strong
terrestrial input.

Based on data published by Chinn, 1991; the Montllobar, Salas and Albian samples display
the greatest potential for source rock analysis based on organic composition, while Reguard
1, Reguard 4 and Pardina samples do not appear promising source rocks for the area.

For further scientific investigation to understand more about the depositional environment I
would analyse 3 key aspects of available samples. Firstly, I would undertake inorganic
analysis for samples abundant in pyrite (Montllobar, Reguard 1 and Reguard 4) I would
investigate to see abundant of framboidal shaped pyrite, and if there I would analyse 34S
isotopic values to understand oceanic conditions when deposited (Krouse and Mayer 2000).
Secondly, I would undertake C/N analysis for organic samples. It can lead to understanding
of sources of sedimentary organic matter. Nitrogen is commonly formed in cellulose of
algae, therefore low C/N ratios can be used to determine marine or terrestrial sources of
the sample. Marine ratios are commonly 4-10:1 while terrestrial shows >20:1 (Shiwatari, R.,
and M. Uzaki, 1987). Finally, I would apply pyrolysis GC-MS with Fourier Transform Infra-red
Radiation (FTIR) to characterize the Kerogen content of individual samples. This method has
been used successfully by Mon Han et al. 2013 for weathered and unweathered coal

44
samples and by van Dongen et al. 2003 for identifying Sulfurization of carbohydrates results
in a sulphur-rich, unresolved complex mixture in kerogen pyrolysates.

45
10. Acknowledgements
This project was funded by the University of Manchester through the grant scheme for
Undergraduate Students to allow travel and accommodation in the area mapped. In
addition funding was also made available to fund analytical techniques used in this project.
Thanks to Stefan Schroeder for supervising the dissertation and travelling out to the field
locality to supervise and assist with sample collection as well as obtaining isotopic analysis
data from the University of Liverpool. I am grateful for help and assistance provided in the
field by my independent mapping partner Samuel Cook who aided me to understand the
broad scale geology of the area. In addition, thanks go to Bart van Dongen and Daniel
Magnone for laboratory assistance and supervision during organc extraction and processing
using the GC-MS. I thank John Waters for technical assistance in obtaining X-Ray Diffraction
results for all samples and Alastair Bewsher for undertaking the loss of Ignition analytical
protocol. Finally, thanks to Stephen Stockley for milling the samples to be used in further
analysis.

46
11. References
AAPG. 2015. Total Organic Carbon (TOC). Available at:
http://wiki.aapg.org/Total_organic_carbon_%28TOC%29#cite_note-ch06r3-1. Last accessed:
12/04/2016.

AAPG. 2016. Hydrocarbon Migration. Available at http://wiki.aapg.org/Hydrocarbon_migration. Last accessed:


10/04/2016.

Al Lawati, W.M., Rizoulis, A., Eiche, E., Boothman, C., Polya, D.A., Lloyd, J.R., Berg, J.R.M. Vasquez-Aguilar, P.
and van Dongen, B.E. 2012. Characterisation of organic matter and microbial communities in contrasting
arsenic-rich Holocene and arsenic-poor Pleistocene aquifers, red river delta, Vietnam. Applied Geochemistry,
27, 315325

Berner, R.A. 1970. Sedimentary pyrite formation. American Journal of Science, 268, 123

Biovision. 2015. Betulin. Available at: http://www.biovision.com/betulin-4949.html. Last accessed: 13/04/2016

Bond, R.M.G, & McClay, K.R. 1995. Inversion of a lower Cretaceous extensional basin, south central Pyrenees,
Spain, in J. G Buchanan and P. G Buchanan, eds., Basin inversion: Geological Society (London) Special
Publication 88, p. 415-431.

Booler, J. & Tucker, M.E. 2002. Distribution and geometry of facies and early diagenesis: The key to
accommodation space variation and sequence stratigraphy: Upper Cretaceous Congost Carbonate platform,
Spanish Pyrenees. Sedimentary Geology, 146, 225247

Bray, E. & Evans, E. 1961. Distribution of n-paraffins as a clue to recognition of source beds. Geochimica et
Cosmochimica Acta, 22, 215

Bristow, T.F., Kennedy, M.J., Derkowski, A., Droser, M.L., Jiang, G. and Creaser, R.A. 2009. Mineralogical
constraints on the paleoenvironments of the Ediacaran Doushantuo formation. Proceedings of the National
Academy of Sciences, 106, 1319013195

Brodie, C. R., Leng, M.J., Casford, J.S.L., Kendrick, C.P., Lloyd, J.M., Zong, Y.Q. and Bird, M.J. 2011. Evidence for
bias in C and N concentrations and 13 C composition of terrestrial and aquatic organic materials due to pre-
analysis acid preparation methods. Chemical Geology 282, 67-83

Bu'Lock, J.D. 1989. Biosynthesis. A Review of Chemical Literature, Vol 4. Page 46.

Caus, E., Parente, M., Vicedo, V., Frijia, G. & Martnez, R. 2013. Sp. Nov., a larger foraminiferal index fossil for the
middle Coniacian shallow-water deposits of the Pyrenean basin (NE Spain). Cretaceous Research, 45, 7690

Chemistrysynthesis. 2016. Available at: http://www.chemsynthesis.com/base/chemical-structure-6760.html. Last


accessed: 13/04/2016.

Chemspider. 2015. Norhopane. Available at: http://www.chemspider.com/Chemical-Structure.10197921.html.


Last accessed 11/04/2016

Chikaraishi, Y. & Naraoka, H. 2001. Organic hydrogen-carbon isotope signatures of terrestrial higher plants during
biosynthesis for distinctive photosynthetic pathways. GEOCHEMICAL JOURNAL, 35, 451458, doi:
10.2343/geochemj.35.451.

47
Chinn, E.W. 1991. The role of organic geochemistry in petroleum exploration: Basin Research Institute Bulletin,
Louisiana State University, Baton Rouge, LA, p. 1523

Choukroune, P. & ECORS Team. 1989. The ECORS Pyrenean deep seismic profile reflection data and the overall
structure of an orogenic belt Tectonics 8, 23-39

Craig, H.S. 1957. Isotopic standards for carbon and oxygen and correction factors for mass-spectrometric analysis
of carbon dioxide. Geochemica et Cosmochimica Acta, 12, 133-149.

Currie, B. & Johns, R. 1989. An organic Geochemical analysis of terrestrial biomarkers in a transect of the great
barrier reef lagoon. Marine and Freshwater Research, 40, 275

Deines, P. 1980. "The isotopic composition of reduced organic carbon." In: P. Fritz and J.Ch. Fontes (Eds.),
Handbook of Environmental Isotope Geochemistry, Vol. 1. Elsevier, New York, pp. 329-406

Dinars-Turell, J. & Garcia-Senz, J. 2000. Remagnetization of lower Cretaceous limestones from the southern
Pyrenees and relation to the Iberian plate geodynamic evolution. Journal of Geophysical Research: Solid Earth,
105, 1940519418

Dow, W. G. 1977. Kerogen studies and geological interpretations: Jour. Geochem. Explor., v. 7, p. 79-99.

ECORS-Pyrenees Team. 1988. The ECORS deep reflection seismic survey across the Pyrenees Nature 331, 508-511

Ekweozor, C. M. & O. T. Udo. 1988. The oleananes: Origin, maturation, and limits of occurrence in Southern
Nigeria sedimentary basins, in L. Mattavelli, and L. Novelli, eds., Advances in Organic Geochemistry 1987,
Organic Geochemistry, v. 13, Pergamon Press, p. 131-140.

Eglinton, G., Hamilton, R.J., 1963. The distribution of n-alkanes. In: Swain, T. (Ed.), Chemical Plant Taxonomy.
Academic Press, London/New York, pp. 187217.

Eglinton, G., Hamilton, R.J., 1967. Leaf epicuticular waxes. Science 156, 13221335.

Friedman, I. & ONeil, J.R. 1977. Compilation of stable isotope fractionation factors of geochemical interest. In:
Fleischer, M. (ed.) Data of Geochemistry (6th edition). United States Geological Survey Professional Paper 440,
chapter KK.12pp.

Garcia-Senez, J. 2002. Cuencas extensivas del Cretacico Infererior en los Pirineos Centrales, formacion y
subsecuente inversion: Ph.D. thesis, Department of Geodynamics and Geophyics, University of Baecelona, 310
p.

Garzione, C., Dettman, D. & Horton, B. 2004. Carbonate oxygen isotope paleoaltimetry: Evaluating the effect of
diagenesis on paleoelevation estimates for the Tibetan plateau. Palaeogeography, Palaeoclimatology,
Palaeoecology, 212, 119140

Gonzales-Vila, F. 1996. Alkane biomarkers. Geochemical significance and application in oil shale geochemistry.
Fuel and Energy Abstracts, 37, 10

Gradstein, F., Ogg, J. and Smith, A., 2004. A Geologic Time Scale 2004, Cambridge Univ. Press, Cambridge.

Gough, M. & Rowland, S. 1991. Characterization of unresolved complex mixtures of hydrocarbons from
lubricating oil feedstocks. Energy & Fuels, 5, 869874

Gross, M. G. 1964. Variations in the 18O/16O and 13C/12C ratios of diagenetically altered limestones in the
Bermuda islands. J. Geol. 72, 172193

48
Guidechem. 2016. Bisnorhopane. Available at: http://www.guidechem.com/dictionary/84276-47-1.html. Last
accessed 11/04/2016

Hatch, J. R. & Leventhal, J. S. 1997. Early diagenetic partial oxidation of organicmatter and sulfides in the Middle
Pennsylvanian (Desmoinesian) Excell Shale Member of the Fort Scott Limestone and equivalents, northern
Midcontinent region, USA. Chem. Geol. 134, 215235.

Hedges, J. I., E. Mayorga, E. Tsamakis, M. E. Mcclain, A. K. Aufdenkampe, P. Quay, J. E. Richey, R. Benner, S.


Opsahl, B. Black, T. Pimentel, J. Quintanilla, L. and Maurice. 2000. Organic matter in Bolivian tributaries of the
Amazon River: A comparison to the lower mainstem. Limnol. Oceanogr., 45: 1449-1466.

Hillier, S. 1993. Origin, Diagenesis, and mineralogy of Chlorite minerals in Devonian Lacustrine Mudrocks,
Orcadian basin, Scotland. Clays and Clay Minerals, 41, 240259

Hughes, W. B., Holba, A.G. and Dzou, L.I.P. 1995. The ratios of dibenzothiophene to phenanthrene and pristane
to phytane as indicators of depositional environment and lithology of petroleum source rocks: Geochimica et
Cosmochimica Acta, v. 59, p. 3581-3598.

Jacobson, S.R. 1991. Petroleum Source Rocks and Organic Facies in Meeeill EK9ed): Source and Migration
Processes and Evaluation Techniques, Tulsa: AAPG, 3-11.

Kendall, C., Caldwell, E. & Synder, D. 2004. Resources on Isotopes. The United States Geological Survey. Available
at: http://wwwrcamnl.wr.usgs.gov/isoig/period/c_iig.html. Last accessed: 02/05/2016.

Krouse, H.R. & Mayer, B. 2000. Sulfur and oxygen isotopes in sulfate, in Environmental Tracers in Subsurface
Hydrology, ed. by P. Cook and A.L. Herczeg, Kluwer Academic Publishers, Norwell, MA.

Largeau C., Derenne S., Casadevall E., Kadouri A. and Sellier N. 1986. Pyrolysis of immature Torbanite and of the
resistant biopolymer (PRBA) isolated from extant alga Botryococcus braunii. Mechanism of formation and
structure of Torbanite. In Advances in Organic Geochemistry 1985 (Edited by Leythaeuser D. and Rullktter J.)
Org. Geochem., 10, 1023-1032. Pergamon Press, Oxford.

Leeuw J.W. de & Largeau C. 1993. A review of macromolecular organic compounds that comprise living organisms
and their role in kerogen, coal and petroleum formation. In Organic Geochemistry principles and applications
(Edited by Engel M. H. and Macko S. A.) pp. 23-72, Plenum Publishing Corp., New-York.

Liebau, A., 1973. El Maastrichtiense lagunar (Garumniense") de Isona. XIII Col. Europ. Micropal.. Espafia, pp. 87-
112.

McCarthy, L., Niemann, M., Palmowski, D., Peters, L. and Syankiewicz, A. 2011. Basic Petroleum Geochemistry
for Source Rock Evaluation. Oilfield Review Summer 23, no 2.

Martinius, A.W. 2012. Contrasting Styles of Silliclastic Tidal Deposit in a Developing Thrust-Sheet-Top-Basins The
Lower Eocene of the Central Pyrenees (Spain). Statoil Research and Development. Principles of Tidal
Sedimentology, p. 473 - 506.

McCrea, J.M. 1950. On the isotopic chemistry of carbonates and paleotemperature scale. Journal of Chemical
Physics, 18, 849-853

Mellere, D. 1993. Thrust generated, back-fill stacking of alluvial fan sequences, south-central Pyrenees, Spain (La
Pobla de Segur Conglomerates) International Association of Sedimentologists, Special Publication, 20, p. 259
276

49
Mencos, J., Munoz, J.A., and Hardy, S. 2011. Three-dimensional geometry and forward numerical modelling of
the Sant Corneli anticline (southern Pyrenees, Spain) in K. McClay, J. Shaw and J. Suppe, eds., Thrust fault-
related folding: AAPG Memoir 94, p. 283-300.

Moldowan, J. M., Seifert, W.K., Arnold, E. and Clardy, J. 1984. Structure proof and significance of stereoisomeric
28,30-bisnorhopanes in petroleum and petroleum source rocks: Geochimica et Cosmochimica Acta, v. 48, p.
1651-1661.

Mon Han, E., Sampei, Y. and Roser, B.P. 2013. Characterization of kerogen using combined pyrolysis- GC-MS and
FT-IR in weathered and unweathered coals and coaly shales from the Central Myanmar Basin, Myanmar. Res.
Org. Geochem. 29, 49 59.

Mook, W.G. 1971. Paleotemperatures and chlorinities from stable carbon and oxygen isotopes in shell carbonate.
Palaeogeography, Palaeoclimatology, Palaeoecology, 9, 245263

Munoz, J. A. 1992. Evolution of a continental collision belt: ECORS-Pyrenees crustal balanced cross-section. In:
Thrust Tectonics (Ed. K. McClay), Chapman & Hall, London, 235-246

Munoz, J. A. 2002. Alpine tectonics: I. The Pyrenees, in W. Gibbons and T. Moreno, eds., Geology of Spain: Bath,
United Kingdom, The Geological Society, p. 367-401.

Nagtegaal, P.J.C., Van Vliet, A. and Brouwer, J. 1983. Syntectonic coastal offlap and concurrent turbidite
deposition: The upper Cretaceous Aren sandstone in the south-central Pyrenees, Spain. Sedimentary Geology,
34, 185218

Nagtegaal, P.J.C. 1972. Depositional history and clay minerals of the Upper Cretaceous basin in the South Central
Pyrenees, Spain. LEIDSE GEOLOGISCHE MEDEDELINGEN, Deel 47, Aflevering 2, pp. 251-275.

Nijman, W. 1998. Cyclicity and basin axis shift in a piggyback basin: Towards modelling of the Eocene Tremp-Ager
basin, south Pyrenees, Spain. Geological Society, London, Special Publications, 134, 135162

Pomar, L., Gili, E., Obrador, A. and Ward, W.C. 2005. Facies architecture and high-resolution sequence
stratigraphy of an upper Cretaceous platform margin succession, southern central Pyrenees, Spain.
Sedimentary Geology, 175, 339365

Pons, J.M., 1977. Estudio estratigrafico y paleontolbgico de los yacimientos de rudistidos dcl ('retacico Sup. del
prepirineo de la prov. de Lerida. Univ. Autonoma de Barcelona, Publ. Geol., 3, 1-105.

Prokoph, A., Shields, G.A. and Veizer, J. 2008. Compilation and time-series analysis of a marine carbonate 18O,
13C, 87Sr/86Sr and 34S database through earth history. Earth-Science Reviews, 87, 113133

Puigde-fabregas, C. & Souquet, P. 1986. Tectosedimentary cycles and depositional sequences of the Mesozoic
and Tertiary from the Pyrenees Tectonophysics 129, 173-203

Rosenbaum, J. & Sheppard, S.M.F. 1986. An isotopic study of siderites, dolomites and ankerites at high
temperatures. Geochimica et Cosmochimica Acta, 50, 11471150.

Roure, F., Choukroune, P., Berastegui, X., Mufioz, J. A., Villien, A., Matheron, P., Bareyt, M., Siguret, M.,
Camara, P. and Daramond, J. 1989. ECORS deep seismic data and balanced cross-sections: geometric
constraints to trace the evolution of the Pyrenees Tectonics 8, 41-50

Sanchez de la Torre, M. 2004. Detecting human mobility in the Pyrenees through the analysis of chert tools during
the Upper Palaeolithic. Journal of Lithic Studeis, Vol 1. No 1.

50
Santisteban, J.I., Mediavilla, R., Lopez-Pamo, E., Dabrio, C.J., Zapata, M.B.R., Garcia, M.J.G., Castano, S. and
Martinez-Alfaro, P.E., 2004. Loss on ignition: a qualitative or quantitative method for organic matter and
carbonate mineral content in sediments? Journal of Paleolimnology 32, 287-299

Schefub, E., Ratmeyer, V., Stuut, J.B., Jansen, J.H.F. and Sinninghe Damst, J.S., 2003. Carbon isotope analyses of
n-alkanes in dust from the lower atmosphere over the central eastern Atlantic ,Geochimica et Cosmochimica
Acta, 67, pp. 17571767

Schoell, M., McCaffrey, M. A., Fago, F.J. and Moldowan, J.M. 1992. Carbon isotopic compositions of 28,30-
bisnorhopanes and other biological markers in a Monterey crude oil: Geochimica et Cosmochimica Acta, v. 56,
p. 1391-1399.

Scielo. 2015. Constituents of Corynaea crassa "Peruvian Viagra". Available at:


http://www.scielo.br/scielo.php?script=sci_arttext&pid=S0102-695X2015000200092. Last accessed:
13/03/2016.

Shackleton, J.R., Cooke, M.L., Vergs, J. and Sim, T. 2011. Temporal constraints on fracturing associated with
fault-related folding at Sant Corneli anticline, Spanish Pyrenees. Journal of Structural Geology, 33, 519

Shibaoka, M., Bennet, A. J. R. and Gould, K. W. 1973. Diagenesis of organic matter and occurrence of
hydrocarbons in some Australian sedimentary basins: APEA Jour., v. 13, p. 73-80.

Shimoyama, A. & Johns, W.D. 1972. Catalytic formation of hydrocarbons from fatty acids. Nature Physical
Science, 237, 6464

Shiwatari, R. & Uzaki, M. 1987. Diagenetic Changes of Lignin Compounds in a More Than 0.6 Million-Year-Old
Lacustrine Sediment (Lake Biwa, Japan). Geochimica Et Cosmochimica Acta 51, no. 2: 321-28

Sigma Aldrich. 2016. Available at:


http://www.sigmaaldrich.com/catalog/product/aldrich/f1506?lang=en&region=GB. Last accessed:
13/04/2016.

Sigma-Aldrich. 2016. Phenanthrene. Available at:


http://www.sigmaaldrich.com/catalog/product/aldrich/p11409?lang=en&region=GB. Last accessed
11/04/2016.

Simo, A. 1986. Carbonate platform depositional sequences, Upper Cretaceous, south-central Pyrenees (Spain):
Tectonophysics, v. 129, p. 205-231.

Sinclair, H.D., Gibson, M., Naylor, M., Morris, R.G. 2005. Asymmetric growth of the Pyrenees revealed through
measurement and modeling of orogenic fluxes. American Journal of Science, 305

Sokolov, V.A., Buniat-Zade, Z.A., Geodekian, A.A. and Dadashev, F.D. 1969. The origin of gases of mud volcanoes
and the regularities of their powerful eruptions, in P. A. Schenck and I. Havenaar, eds., Advances in organic
geochemisty, 1968: Oxford, England, Pergamon Press, p. 473-484.

Stable Isotope Ecology Laboratory, 1997. Overview of Stable Isotope Research. Available at:
http://sisbl.uga.edu/stable.html#calib. Last accessed 08/04/2016.

Swart, P.K., Burns, S.J. and Leder J.J. 1991. Fractionation of the stable isotopes of oxygen and carbon in carbon-
dioxide during reaction of calcite with phosphoric acid as a function of temperature and technique. Chemical
Geology, 52, 365-374

51
Tavani, S., Mencos, J., Baus, J. and Muoz, J.A. 2011. The fracture pattern of the Sant Corneli Bixols oblique
inversion anticline (Spanish Pyrenees). Journal of Structural Geology, 33, 16621680

Tegelaar E. W., Matthezing R. M., Jansen J. B. H., Horsfield B. and de Leeuw J. W. 1989. Possible origin of n-
alkanes in high-wax crude oils. Nature 342, 529-531.

The British Geological Survey, 2012. Cretaceous. Available at:


http://www.bgs.ac.uk/discoveringGeology/time/timechart/phanerozoic/cretaceous.html. Last accessed:
06/04/2016.

The British Geological Survey, 2012. Palaeogene to Quaternary. Available at:


http://www.bgs.ac.uk/discoveringGeology/time/timechart/phanerozoic/cenozoic.html. Last accessed:
06/04/2016.

The University of Manchester, 2013. School of earth, atmospheric and environmental sciences. World Wide Web
Address: http://www.seaes.manchester.ac.uk/our-research/research-areas/mes/facilitiesold/x-
raydiffractionmossbauer/. Last Accessed 25/02/2016

Tissot, B., Durand, B., Espitali, J., and Combaz, A. 1974. Influence of Nature and Diagenesis of Organic Matter in
Formation of Petroleum, AAPG Bulletin 58, no. 3: 499506.

Tissot, B., Durand., Espitalie, J. and Combaz, A. 1974. Influence of nature and diagenesis of organic matter in
formation of Petroleum: AAPG Bull., v. 58, p. 499-500.

Tulloch, A.P. 1976. Epicuticular wax of Agropyron smithii leaves. Phytochemistry, 15, 11531156

Van der Meulen, S. 1986. Sedimentary stratigraphy of Eocene sheetflood deposits, southern Pyrenees, Spain.
Geological Magazine, 123, 167

Van Dongen, B.E., Schouten, S. & Sinninghe Damst, J.S. 2003. Sulfurization of Carbohydrates results in a sulfur-
rich, unresolved complex mixture in Kerogen Pyrolysates. Energy & Fuels, 17, 11091118,

Vassoevich, N. B., Akramkhodzhaev, A.M. and Geodekyan, A.A. 1974. Principal zone of oil formation, in B. Tissot
and F. Bienner, eds., Advances in organic geochemistry, 1973: Paris, Editions Technip, p. 309-314.

Yamamoto, M. & Wantanabe, Y. 1994. Biomarker geochemistry and palaeoceanography of Miocene Onnagawa
diatomeous sediment, Northern Honshu, Japan. Geological Survey of Japan pp 53-74.

Zhang, Y.F. 2013. Natural Gas Exploration Using Carbon Isotopic Fractionation Effect: A Case Study of Shanxi
Formation, Upper Palaeozoic Group in the Center of Ordos Basin, China. International Journal of Chemical
Engineering and Applications, Vol. 4, No.1. pp 18-20.

52
12. Appendices
n-alkanes n-alcohols Fatty Acids
Retention Time (m) Area Compound Retention Time (m) Area Compound Retention Time (m) Area Compound
26.274 3007344 22 11.681 1099409 12 15.408 24789922 13
26.911 887137 23 13.914 4227497 13 16.195 21965334 14
30.398 6638693 24 16.195 22772738 14 16.575 902611843 15
32.44 176001549 25 18.517 51062813 15 20.83 19834202 16
34.254 34963814 26 20.825 55501188 16 22.422 10559706 17
36.084 169036540 27 22.44 10559706 17 24.685 87369606 18
37.844 55829285 28 25.299 316634314 18 27.546 87443591 19
39.553 37149868 29 26.785 3007344 19 28.978 24892728 20
41.197 24362714 30 29.426 5160969 20 1179466932
42.795 20600859 31 30.845 4889773 21
44.334 14571588 32 33.289 1068950 22
45.845 10634502 33 34.618 863940 23
47.313 7538290 34 36.879 4402835 24
48.733 4344569 35 481251476
50.115 2496893 36
51.462 3393078 37
Sum: 571456723 Tetracosane: 27407900
Figure 14.1 showing Retention times, area and chain length of n-alkanes, n-alcohols and Fatty Acids of Albian.

n-alkanes n-alcohols Fatty Acids


Retention Time (m) Area Compound Retention Time (m) Area Compound Retention Time
Area(m) Compound
14.783 147629101 17 11.701 11082749 12 10.947 15963101 12
17.112 96251059 18 12.353 80603647 13 13
19.495 40553860 19 16.222 74014327 14 15.374 146300592 14
21.816 30643404 20 18.585 111791796 15 17.736 28680676 15
24.077 25981412 21 20.852 39609400 16 20.133 420135048 16
26.276 49754705 22 22.257 19567990 17 22.454 30631035 17
28.374 23179981 23 25.278 31849109 18 24.694 261986638 18
30.41 8795808 24 368519018 26.894 6607992 19
32.372 230485078 25 28.985 53938985 20
34.273 39079105 26 Tetracosane 40321388 31.015 14543704 21
36.113 82102498 27 32.963 380337669 22
37.864 143866546 28 34.843 20852219 23
39.575 104334803 29 36.67 15162383 24
41.204 68893341 30 38.421 48985042 25
42.813 105155264 31 40.105 20589303 26
44.348 32480849 32 41.754 35426463 27
45.848 44312972 33 42.884 105155264 28
47.314 12045599 34 43.336 13910164 29
48.733 12273365 35 46.384 45022926 30
51.326 13373055 36 47.85 8119146 31
1311191805 49.263 226957732 32
1899306082
Figure 14.2 showing Retention times, area and chain length of n-alkanes, n-alcohols and Fatty Acids of Montllobar.

n-alkanes n-alcohols Fatty Acids


Retention Time (m) Area Compound Retention Time
Area
(m) Compound Retention Time (m) Area Compound
8.375 49274522 14 11.64 48571393 12 10.812 17571805 12
10.269 165951250 15 25.312 24373160 18 n/a n/a 13
12.455 296273134 16 15.353 23988544 14
14.831 504153468 17 Tetracosane 67051522 17.736 23406041 15
17.2 404198838 18 20.126 117169015 16
19.562 225630387 19 22.448 35568047 17
21.891 224681546 20 24.708 112147430 18
24.158 240527387 21 26.887 14696244 19
26.351 273691589 22 28.992 9269066 20
28.455 225616511 23 31.022 5227248 21
30.485 220988844 24 32.99 267826161 22
32.461 442975899 25 34.85 28865882 23
34.341 200287971 26 36.677 21601997 24
36.167 255733019 27 38.435 10389103 25
37.926 206932876 28 40.139 26252967 26
39.623 171019905 29 41.788 21059930 27
41.272 130327191 30 43.357 14308009 28
42.847 82748143 31 44.904 6369533 29
44.382 64674947 32 755717022
45.882 38078412 33
47.341 21874240 34
4445640079
Figure 14.3 showing Retention times, area and chain length of n-alkanes, n-alcohols and Fatty Acids of Pardina.

53
n-alkanes n-alcohols Fatty Acids
Retention Time (m) Area Compound Retention Time
Area
(m) Compound Retention Time
Area(m) Compound
12.441 84105705 16 11.674 19932962 12 10.886 29896645 12
14.81 109483629 17 13.901 57587558 13 13
17.125 34380339 18 16.215 34112020 14 15.36 95407670 14
19.494 86774449 19 25.278 19183765 18 17.743 32502723 15
21.816 31975193 20 130816305 20.126 279142000 16
24.077 37713813 21 22.441 31784843 17
26.276 30818150 22 Tetracosane 49544257 24.688 145676543 18
28.367 29423267 23 26.86 68015100 19
30.397 37080137 24 28.978 18305466 20
32.379 328958804 25 31.001 11774172 21
34.26 63561709 26 32.977 509537504 22
36.106 106747708 27 35.357 567717996 23
37.858 64668431 28 36.88 18014746 24
39.568 72917248 29 38.421 11093675 25
41.211 36251713 30 40.111 5913257 26
42.807 43021094 31 41.748 14555797 27
44.354 20738098 32 43.329 1764553 28
45.848 20192023 33 44.891 20413588 29
47.301 21561017 34 46.371 4665507 30
48.719 4565769 35 47.83 6690890 31
1264938296 49.242 8340697 32
50.627 9058626 33
1890271998

Figure 14.4 showing Retention times, area and chain length of n-alkanes, n-alcohols and Fatty Acids of Salas.

Sample Crucible Mass (g) Crucible Mass with Sample (g) Mass after 105C (g) Mass after 550C (g) Weight after 1000C (g)
Reguard 1 14.7579 16.0271 16.0202 16.015 15.5315
Pardina 15.4226 16.4978 16.4953 16.4942 16.0386
Salas 11.1773 12.1693 12.1458 12.1271 11.879
Albian 11.2334 12.4373 12.4225 12.4015 12.0662
Montollobar 14.7816 15.9441 15.9319 15.914 15.7141
Reguard 4 14.9841 16.0184 16.0117 16.0063 15.6333

Sample Mass of Sediment (g) Mass of Water (g) Mass of Organic Carbon (g) Mass of Inorganic Carbon (g)
Reguard 1 1.2692 0.0069 0.0052 0.4835
Pardina 1.0752 0.0025 0.0011 0.4556
Salas 0.992 0.0235 0.0187 0.2481
Albian 1.2039 0.0148 0.021 0.3353
Montollobar 1.1625 0.0122 0.0179 0.1999
Reguard 4 1.0343 0.0067 0.0054 0.373

Sample Mass of Water (%) Mass of Organic Carbon (%) Mass of Inorganic Carbon (%)
Reguard 1 0.543649543 0.409706902 38.09486291
Pardina 0.232514881 0.102306548 42.3735119
Salas 2.368951613 1.885080645 25.01008065
Albian 1.229337985 1.744330924 27.85115043
Montollobar 1.049462366 1.539784946 17.19569892
Reguard 4 0.647781108 0.522092236 36.0630378

Figure 14.5 showing Loss on Ignition calculations

54
Chemspider, 2015: Norhopane Guidechem, 2016: Bisnorhopane

Biovision, 2015: Betulin Sigma-Aldrich, 2016. Phenenthrene

Lookfordiagnosis, 2016. Oleanoic Acid. Similar to Scielo, 2015. Lupeol, similar Lupanone
Oleane and Oleanol

55
Sigmaaldrich, 2016. Fluorenone. Chemistrysynthesis, 2016. Anthracene.

56

You might also like