Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

download Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

of 16

Transcript of Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    1/16

    This article appeared in a journal published by Elsevier. The attached

    copy is furnished to the author for internal non-commercial research

    and education use, including for instruction at the authors institution

    and sharing with colleagues.

    Other uses, including reproduction and distribution, or selling or

    licensing copies, or posting to personal, institutional or third partywebsites are prohibited.

    In most cases authors are permitted to post their version of the

    article (e.g. in Word or Tex form) to their personal website or

    institutional repository. Authors requiring further information

    regarding Elseviers archiving and manuscript policies are

    encouraged to visit:

    http://www.elsevier.com/copyright

    http://www.elsevier.com/copyrighthttp://www.elsevier.com/copyright
  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    2/16

    Author's personal copy

    Petroleum geochemistry of the Potwar Basin, Pakistan: 1. Oiloil correlation

    using biomarkers, d13C and dD

    Muhammad Asifa,b,, Tahira Fazeelat c, Kliti Grice b

    a Department of Basic Sciences and Humanities, University of Engineering and Technology, KSK Campus, Lahore, Pakistanb WA Organic and Isotope Geochemistry Centre, The Institute for Geoscience Research, Department of Chemistry, Curtin University, GPO Box U1987, Perth, WA 6845, Australiac Chemistry Department, University of Engineering and Technology, GT Road, Lahore, Pakistan

    a r t i c l e i n f o

    Article history:

    Received 16 January 2011

    Received in revised form 21 April 2011

    Accepted 5 August 2011

    Available online 16 August 2011

    a b s t r a c t

    Geochemical characterisation of 18 crude oils from the Potwar Basin (Upper Indus), Pakistan is carried

    out in this study. Their relative thermal maturities, environment of deposition, source of organic matter

    (OM) and the extent of biodegradation based on the hydrocarbon (HC) distributions are investigated. A

    detailed oiloil correlation of the area is established. Gas chromatographymass spectrometry (GC

    MS) analyses and bulk stable carbon and hydrogen isotopic compositions of saturated and aromatic

    HC fractions reveals three compositional groups of oils. Most of the oils from the basin are typically gen-

    erated from shallow marine source rocks. However, group A contains terrigenous OM deposited under

    highly oxic/fluvio-deltaic conditions reflected by high pristane/phytane (Pr/Ph), C30 diahopane/C29Ts,

    diahopane/hopane and diasterane/sterane ratios and low dibenzothiophene (DBT)/phenanthrene (P)

    ratios. The abundance of C19-tricyclic and C24-tetracyclic terpanes are consistent with a predominant ter-

    rigenous OM source for group A. Saturated HC biomarker parameters from the rest of the oils show a pre-

    dominant marine origin, however groups B and C are clearly separated by bulk d13C and dD and the

    distributions of the saturated HC fractions supporting variations in source and environment of depositionof their respective source rocks. Moreover, various saturated HC biomarker ratios such as steranes/

    hopanes, diasteranes/steranes, C23-tricyclic/C30 hopane, C28-tricyclic/C30 hopane, total tricyclic ter-

    panes/hopanes and C31(R + S)/C30 hopane show that two different groups are present. These biomarker

    ratios show that group B oils are generated from clastic-rich source rocks deposited under more suboxic

    depositional environments compared to group C oils. Group C oils show a relatively higher input of algal

    mixed with terrigenous OM, supported by the abundance of extended tricyclic terpanes (up to C 41+) and

    steranes.

    Biomarker thermal maturity parameters mostly reached to their equilibrium values indicating that the

    source rocks for Potwar Basin oils must have reached the early to peak oil generation window, while aro-

    matic HC parameters suggest up to late oil window thermal maturity. Theextent of biodegradation of the

    Potwar Basin oils is determined using various saturated HC parameters and variations in bulk properties

    such as API gravity. Groups A and C oils are not biodegraded and show mature HC profiles, while some of

    the oils from group B show minor levels of biodegradation consistent with high Pr/n-C17, Ph/n-C18 and

    low API gravities.

    2011 Elsevier Ltd. All rights reserved.

    1. Introduction

    The Potwar Basin is the main source of petroleum hydrocarbons

    in northern Pakistan. A number of small and medium sized oil and

    gas fields have been discovered from both terrigenous and marine

    source rocks in the basin. These oil fields are in Precambrian to Ter-

    tiary aged reservoir units. The Paleocene Patala Formation type II

    and III kerogens have been assumed to be the primary source of

    hydrocarbons in the area, but other potential source rocks may

    have also contributed to different parts of the petroleum systems

    within the basin (Wandrey et al., 2004; Fazeelat et al., 2010). The

    oldest potential source rocks are from the Precambrian Salt Range

    Formation with a mixture of clastic, carbonate and evaporite dom-

    inant sections. Similarly, Permian Sardhai and Chhidru formations

    have significantly higher total organic carbon (TOC) contents and

    are possible source rocks (Quadri and Quadri, 1997). In the past,

    a limited number of studies have been carried out mainly based

    on RockEval pyrolysis data (Ahmed and Alam, 1990; Fazeelat

    et al., 2010) and studies of the saturated hydrocarbon (HC) distri-

    butions (Fazeelat, 1994; Ahmed and Alam, 2007) fromthe area. Re-

    cently, a biodegraded oil seep and a crude oil from the Potwar

    0146-6380/$ - see front matter 2011 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.orggeochem.2011.08.003

    Corresponding author at: Department of Basic Sciences and Humanities,

    University of Engineering and Technology, KSK Campus, Lahore, Pakistan.

    Tel.: +92 3218850163.E-mail address: [email protected] (M. Asif).

    Organic Geochemistry 42 (2011) 12261240

    Contents lists available at SciVerse ScienceDirect

    Organic Geochemistry

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / o r g g e o c h e m

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    3/16

    Author's personal copy

    Table

    1

    Geologicalinformation,

    n-a

    lkanes,isoprenoidratiosandbulkisotopedata.

    Sample

    name

    Depth

    (m)

    Reservoir

    API

    gravity

    Pr/Pha

    Pr/n-C17

    b

    Ph/n-C18c

    CP

    Id

    OEPe

    d13Csats

    f

    ()

    d13Caros

    ()

    dDsats

    g

    ()

    dDaros

    ()

    d13Caver

    h

    ()

    dDaver

    ()

    DBT/Pi

    Group

    BLj

    Formation

    Age

    P1

    2680

    Khewra

    Cambrian

    48.0

    3.2

    0.4

    0.2

    1.0

    1.0

    26.4

    24.5

    117

    111

    25.4

    114

    0.16

    A

    0

    P2

    2187

    Chorgali

    Eocene

    36.2

    1.5

    1.0

    0.7

    0.9

    1.0

    23.1

    20.8

    155

    130

    21.9

    142

    0.17

    B

    0

    P3

    2063

    Jutana

    Cambrian

    33.0

    2.0

    1.0

    0.5

    1.0

    1.0

    0.22

    B

    0

    P4

    3640

    Jutana

    Cambrian

    19.3

    1.2

    1.3

    0.9

    1.0

    1.0

    0.52

    B

    2

    P5

    3645

    Jutana

    Cambrian

    22.7

    1.3

    1.3

    0.9

    1.0

    1.0

    22.4

    21.0

    132

    125

    21.7

    128

    0.45

    B

    2

    P6

    2773

    Khewra

    Cambrian

    26.6

    1.3

    1.4

    0.9

    1.0

    1.0

    0.43

    B

    3

    P7

    2694

    Khewra

    Cambrian

    25.0

    1.5

    1.2

    0.8

    1.0

    0.9

    23.0

    21.1

    149

    135

    22.0

    142

    0.37

    B

    2

    P8

    3063

    Khewra/Tobra

    Cambrian

    23.2

    1.3

    1.1

    0.8

    1.0

    0.9

    0.39

    B

    1

    P9

    3318

    Khewra/Tobra

    Cambrian

    25.0

    1.4

    0.8

    0.6

    1.0

    1.0

    22.9

    22.2

    126

    132

    22.5

    129

    0.33

    B

    1

    P10

    2687

    Khewra

    Cambrian

    18.4

    1.2

    1.3

    0.9

    0.9

    0.9

    22.6

    21.9

    132

    141

    22.2

    137

    0.66

    B

    3

    P11

    2179

    Chorgali

    Eocene

    16.0

    1.0

    1.3

    1.0

    0.9

    0.9

    23.0

    21.1

    136

    136

    21.7

    136

    0.84

    B

    3

    P12

    2103

    Chorgali/

    Sakaser

    Eocene

    16.1

    1.0

    1.3

    1.0

    0.9

    0.8

    22.3

    21.1

    130

    134

    21.6

    132

    0.93

    B

    3

    P13

    3612

    Chorgali/

    Sakaser

    Eocene

    16.0

    1.2

    1.1

    0.8

    1.0

    0.9

    22.3

    21.0

    145

    129

    21.8

    137

    0.31

    B

    1

    P14

    Chorgali/

    Sakaser

    Eocene

    40.0

    1.5

    0.8

    0.5

    1.0

    1.0

    23.1

    20.5

    145

    139

    23.5

    142

    0.27

    B

    0

    P15

    4096

    Chorgali/

    Sakaser

    Eocene

    45.0

    1.4

    0.9

    0.7

    1.0

    1.0

    25.0

    22.0

    148

    139

    23.6

    143

    0.26

    C

    0

    P16

    4174

    Chorgali/

    Sakaser

    Eocene

    46.0

    1.4

    0.9

    0.7

    1.0

    1.0

    25.1

    22.1

    23.8

    0.26

    C

    0

    P17

    4485

    Datta

    Jurassic

    38.4

    1.6

    0.6

    0.4

    1.0

    1.0

    26.1

    21.5

    129

    122

    23.7

    126

    0.25

    C

    0

    P18

    4450

    Datta

    Jurassic

    41.1

    1.6

    0.8

    0.6

    1.0

    1.0

    26.1

    21.4

    25.4

    0.16

    C

    0

    :notdeter

    mined.

    aPr/Ph,pristane/phytane.

    bPr/n-C17

    ,pristane/n-C17alkanes.

    cPh/n-C18,phytane/n-C18alkanes.

    dCPI,carbonpreferenceindex.

    eOEP,oddevenpredominance.

    fd13C()withrespecttoVPDBreportedwithinstandard

    deviationof0.2.

    gdD()withrespectofVSMOWwithinstandarddeviati

    onof3.

    h

    d13Caver:age(d13Csats+

    d13Caros)/2;dDaver:(dDsats+dDaros)/2

    .

    iDBT/P,d

    ibenzothiophene/phenanthrene.

    jBL,biod

    egradationlevel(Wengeretal.,2001).

    M. Asif et al. / Organic Geochemistry 42 (2011) 12261240 1227

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    4/16

    Author's personal copy

    Basin have been correlated based on their biomarker distributions

    (Fazeelat et al., 2011). Organic geochemical data, particularly bio-

    markers from potential source rocks in the Potwar Basin, have

    never been reported, nor has a detailed oiloil correlation ever

    been undertaken.

    The application of biomarkers and stable isotope analyses hasbeen recognised as a powerful tool in exploration petroleum geo-

    chemistry (e.g. Dawson et al., 2005, 2007; Asif et al., 2009; Maslen

    et al., 2011). Biomarkers (based on structural grounds) in bitumi-

    nous organic matter (OM) can provide valuable information on:

    (i) the source of their natural product precursors (i.e. Eukaryotes,

    Prokaryotes and Archaea), (ii) paleoenvironmental depositional

    conditions (marine, lacustrine, hypersaline or fluvio-deltaic), (iii)

    lithology of potential petroleum source rocks (carbonate versus

    shale), (iv) relative thermal maturity of potential source rocks

    and (v) extent of biodegradation of petroleum HCs. However, many

    of the above factors are often interrelated and have been consid-

    ered collectively for correlation studies (e.g. Murray and Boreham,

    1992). Variation in biomarker abundance has been used success-

    fully for oil correlation between source rocks and/or other oils(e.g. Jiang and Li, 2002; Obermajer et al., 2002; Pasadakis et al.,

    2004; Zhang and Huang, 2005). Bulk isotope analysis (carbon and

    hydrogen) of crude oils, bitumens and kerogen is also useful. The

    bulk isotope composition of saturated and aromatic fractions of

    crude oils has been applied to determine the source of OM (terrig-

    enous versus marine) for many oils worldwide (Sofer, 1984; Chung

    et al., 1992; Andrusevich et al., 1998). However, a recent study has

    showed limitations of bulk d13C data and reported a d13C values of

    individual aromatic HCs to determine the source OM of Western

    Australian crude oils (Maslen et al., 2011).

    In the current study, organic geochemical parameters based on

    biomarker distributions and stable carbon and hydrogen isotopes

    of saturated and aromatic HC fractions have been used to investi-

    gate the source and thermal maturity of OM, depositional paleoen-

    vironment, lithology and extent of biodegradation of HCs from the

    Potwar Basin (Upper Indus Basin, Pakistan). Heavy to light crude

    oils have been identified in a small region of the Potwar Basin.

    The causes for these differences in physical properties are pres-

    ently unknown. The possible reasons for these different crude oil

    types are related to the complex geology of the area and/or differ-

    ences in the source of OM and its environment of deposition.

    2. Experimental

    2.1. Samples and geological setting

    Eighteen crude oils were selected from the Potwar Basin

    (Table 1). The oils cover half of the Potwar plateau, which is the

    largest HC producing zone in northern Pakistan (Upper Indus Ba-

    sin). The source of these crude oils has not been fully correlated

    with any specific source rocks of the Upper Indus Basin. A few

    studies using organic geochemical properties from cuttings, out-

    crops and core samples of different geological formations were

    undertaken andcorrelated partially with Potwar crude oils (Ahmed

    and Alam, 1990, 2007). Both heavy and light oils have been discov-

    ered in the basin (API gravity, Table 1). The heavy oils are geneti-

    cally related to light oils, and bear a close spatial relationship

    (Asif et al., 2008). The locations of crude oils used in this studyare shown in Fig. 1 and marked on a stratigraphic chart in corre-

    sponding geological formations in Fig. 2.

    The geological structure of the Potwar Basin is very complex

    due to the result of the Tertiary Himalayan collision between the

    Eurasian and Indian plates (Farah et al., 1984). This intense tectonic

    activity has significantly over thrust the Himalayan in the north

    and northwest and a series of faulted and unfaulted anticlines

    developed as the result of the multiple detachments of deep sur-

    faces from Cambrian. The Potwar Basin contains mostly carbonate

    reservoir rocks of Precambrian to Tertiary ages. The basin infill

    started with thick Precambrian evaporite deposits overlain by rel-

    atively thin Cambrian to Eocene age deposits followed by thick

    MiocenePliocene deposits. The Precambrian salt deposits pro-

    vided an easy detachment of Eocene-to-Cambrian (EC) sequencesas a result of intense tectonic activity during Himalayan Orogeny

    during the Pliocene to middle Pleistocene. This EC sequence in

    the Potwar Basin affected by compressional forces has generated

    a large number of folds and faults in the area ( Aamir and Siddiqui,

    2006). These folds and faults have formed many small reservoirs

    Fig. 1. Map of the Potwar Basin indicating locations of oils used in this study (Wandrey et al., 2004; modified from USGS Bulletin 2208B and references therein).

    1228 M. Asif et al. / Organic Geochemistry 42 (2011) 12261240

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    5/16

    Author's personal copy

    Fig. 2. Stratigraphy of the Potwar Basin. The reservoir formations corresponding to oil samples used in this study are also shown in the right column ( Wandrey et al., 2004;modified from USGS Bulletin 2208B and references therein).

    M. Asif et al. / Organic Geochemistry 42 (2011) 12261240 1229

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    6/16

    Author's personal copy

    and thus crude oils discovered in the Potwar Basin have shown to

    be derived from a range of geological formations (Fig. 2).

    The generalised stratigraphy of the Potwar Basin is shown in

    Fig. 2. The oldest producing reservoir is a Precambrian Salt Range

    formation. It consists of thick carbonates overlain by evaporites.

    Marine shales and massive sandstones have been reported in theKhewra Formation of the Lower Cambrian Jhelum Group. The Khe-

    wra Formation has yielded the P1 and P6P10 oils analysed in the

    present study. The overlying Jutana Formation consists primarily of

    sandy carbonates and nearshore sandstones and reservoirs that

    have led to the P3P5 oils. The Permian Tobra Formation consists

    of glacial tillites, siltstone and shales, and samples P8 and P9 are

    derived from the Tobra and the Khewra formations. The Jurassic

    strata include the Datta and Shinawari formations, which are near-

    shore consisting of siliciclastics that contain some non-marine

    sandstone intervals (Khan et al., 1986). The Datta Formation has

    led to the production of the P17 and P18 oils. Shallow marine car-

    bonate strata of the Eocene Chorgali and Sakaser formations form

    an important HC producing horizon in the Potwar Basin. Chorgali

    and Sakaser formations consist of medium-bedded limestonesand fine crystalline dolomites. Both formations are oil and gas pro-

    ducing reservoirs and the P2 and P11P16 oils are derived from the

    Chorgali and Sakesar formations.

    2.2. Sample preparation, liquid chromatography and 5A molecular

    sieving

    The sample preparation and liquid chromatography procedures

    used were similar to those described in Asif et al. (2009). Briefly,

    crude oils were separated using liquid chromatography on silica

    gel. Saturated HC fractions were eluted with n-hexane (30 ml), aro-

    matic fractions with a mixture ofn-hexane:CH2Cl2 (7:3, 30 ml) and

    polars with a mixture of CH2Cl2:CH3OH (1:1, 30 ml). Saturated and

    aromatic fractions were used for bulk d13C analyses and a portion

    of the saturated HC fraction was subjected to 5A molecular sieving.

    Straight chain HCs were separated from branched and cyclic HCs

    by treating the saturated fractions with 5A molecular sieve

    (Murphy, 1969) using the procedure described elsewhere (Grice

    et al., 2008; Maslen et al., 2009).

    2.3. GCMS

    GCMS was performed using similar conditions as in Asif et al.

    (2009). Briefly, we used a HewlettPackard (HP) 5973 Mass Selec-

    tive Detector (MSD) interfaced to a HP6890 gas chromatograph

    equipped with a column (60 m, 0.25 mm ID) with a 0.25 lm 5%phenyl 95% methyl polysiloxane stationary phase (DB-5MS, J&W

    scientific). The injector was operated at 280 C and the He carrier

    gas was maintained at a constant flow of 1.1 ml/min. The GC oven

    was programmed from 40310 C at 3 C/min with initial and final

    hold times of 1 and 30 min, respectively. The transfer line between

    the gas chromatograph and the MSD was held at 310 C. The MS

    source and quadrupole temperatures were 230 C and 106 C,

    respectively. The saturated and aromatic HCs were identified using

    relative retention times, mass spectral data and comparison with

    literature data (Trolio et al., 1999; Grice et al., 2001, 2005; Grimalt

    et al., 2002; Peters et al., 2005 and references therein).

    2.4. Elemental analysis-isotope ratio mass spectrometry (bulk isotope

    analysis)

    Bulk stable isotope compositions of both the saturated and aro-

    matic HC fractions of Potwar Basin oils were analysed. Bulk stableisotope analyses were performed on a Micromass IsoPrime isotope

    ratio mass spectrometer interfaced to a EuroVector EuroEA3000

    elemental analyser. For bulk d13C analyses, each sample was

    weighed accurately (0.050.15 mg) into a small tin capsule which

    was then folded and compressed carefully to remove any tracers of

    atmospheric gases. The tin capsule containing the sample was

    dropped into a combustion reactor at 1025 C with the aid of an

    autosampler. The sample and capsule melted in an atmosphere

    temporarily enriched with oxygen, where the tin promoted flashcombustion. The combustion products, in a constant flow of he-

    lium, passed through an oxidation catalyst (chromium oxide).

    The oxidation products then passed through a reduction reactor

    at 650 C containing copper granules, where any oxides of nitrogen

    (NO, N2O and N2O2) are reduced to N2 and SO2 were separated on a

    3 m chromatographic column (Poropak Q) at ambient temperature.

    After removing the oxides of nitrogen, oxidation products are then

    passed through a thermal conductivity detector (TCD) and into the

    irMS. Average values of at least two analyses and standard devia-

    tions are reported. Isotopic compositions are given in delta nota-

    tion relative to Vienna Peedee belemnite (VPDB).

    For bulk dD analysis, the sample was weighed accurately (0.05

    0.15 mg) into a small silver capsule which was then folded and

    dropped into a pyrolysis reactor containing glassy carbon chipsheld at 1260 C. The sample was pyrolysed to form H2 and CO,

    along with N2 if applicable. The pyrolysis products were separated

    on a 1 m 5A molecular sieve, packed chromatographic column held

    in an oven at 80 C (isothermal), before passing through a TCD,

    then into the irMS. dD values were calculated and reported in delta

    notation relative to Vienna Standard Mean Ocean Water (VSMOW).

    3. Results and discussion

    3.1. Normal alkanes and regular isoprenoid distributions

    The crude oils analysed in this study are listed in Table 1. The

    total ion chromatogram (TIC) of a representative sample is shown

    in Fig. 3. n-Alkanes range from C1037 along with isoprenoids fromC1420 (except i-C17) are observed while n-alkanes less than C10 are

    absent, probably because of evaporative loss during sample pro-

    cessing. The peak areas from the TIC are used to calculate n-alkane

    and isoprenoid parameters listed in Table 1. All the samples show

    Pr/Ph ratios from 1.02.0, representing marine oxic/dysoxic depo-

    sitional environment of OM, with the exception of the P1 oil which

    shows a value >3.2, suggesting more oxic depositional conditions.

    Carbon preference index (CPI) and odd/even predominance

    (OEP) are good indicators for OM type in immature samples where

    higher abundance of C1618 n-alkanes indicate an aquatic source

    while a C2733 odd abundance of n-alkanes reflects terrigenous

    OM (Hunt, 1995). Maturation of OM significantly alters n-alkane

    abundance, and CPI and OEP tend towards 1 in typical mature

    crude oils. The CPI and OEP values for most of the Potwar Basin oilsare close to 1 (Table 1) indicating mid-oil window thermal maturi-

    ties of the source at time of expulsion.

    3.2. Stable carbon and hydrogen isotopic compositions

    The stable isotopic composition of crude oil is mainly depen-

    dent on the d13C and dD value of the kerogen which, in turn, de-

    pends on the biological OM and the depositional environment

    (Schoell, 1984; Chung et al., 1992; Collister and Wavrek, 1996). A

    series of oils ranging from low API condensate (API, 48) to high

    API oils (API, 16) from the Potwar Basin were examined for bulk

    d13C and dD. These results are reported in Table 1. d13C and dD of

    saturated and aromatic HC fractions were analysed to delineate

    different groups of petroleum in the Potwar Basin. The crude oilsfrom eastern Potwar are characterised by less negative values of

    d13C (isotopically heavy) and cluster together based on d13C of sat-

    urated and aromatic HC fractions (Fig. 4a; group B). More negative

    1230 M. Asif et al. / Organic Geochemistry 42 (2011) 12261240

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    7/16

    Author's personal copy

    (isotopically lighter) d13C values (25 to 26.1) are observed

    for P15P18 crude oil samples (Fig. 4a; group C). Among the sam-

    ple suite of oils analysed, P1 oil is the lightest (most negative) ind13C of saturated and aromatic HC fractions and is classified as

    group A (Fig. 4a). The group designation assigned to each crude

    oil from the Potwar Basin is shown in Table 1. Group A comprises

    a single oil, group B contains 13 crude oils and group C contains

    four crude oils. The stable isotopic composition of group A oil is

    most probably controlled by both source and depositional settings

    as indicated by n-alkanes and isoprenoid distributions and bio-

    marker parameters (see below). Group B oils show enrichment in13C for the saturated HC fraction having values up to 34 heavier

    compared to the saturated HC fraction d13C values of group C

    (Fig. 4a; Table 1). The differences observed between d13C of satu-

    rated HC fractions of groups B and C indicate a difference in source

    organisms for the saturated HCs. Another plot represents the dif-ferences between d13C and dD average values of both saturated

    and aromatic HC fractions of crude oils (Table 1) and is shown in

    Fig. 4b. This plot separates the crude oils into three similar groups

    providing additional evidence for the existence of at least three oil

    groups in the Potwar Basin. The difference in d13C and dD of the

    Potwar Basin oils most likely suggests source variations.

    The biomarker parameters listed in Tables 13 are used to ob-

    tain information regarding source OM, thermal maturity of crude

    oils, depositional environment and lithology of OM and the extent

    of biodegradation in the Potwar Basin crude oils. The following sec-

    tions explain these geochemical characteristics one by one, again

    differentiating the Potwar Basin crude oils into three groups.

    3.3. Thermal maturity of Potwar Basin crude oils

    A combination of saturated and aromatic HC biomarkers was

    used to determine the thermal maturity of the Potwar Basin oils.

    The data in Table 2 were obtained from GCMS analysis of the

    branched/cyclic fractions. The hopane based parameters were cal-

    culated from peak areas of 191 Dalton mass chromatograms. The

    C32 homologue ratio 22S/(22S+ 22R) varies between 0.57 and

    0.64, indicating equilibrium values for this ratio, suggesting that

    the maturity is at least equal to early oil generation window for

    all the oil samples of the Potwar Basin (Table 2). However this ratio

    reaches equilibrium in the oil window so has limited application

    for studying the relative maturities of crude oils and condensates

    (Peters et al., 2005). Another hopane based maturity parameter is

    the ratio of 17a(H),21b(H)-hopane relative to corresponding17b(H),21a(H)-moretanes [ab/(ab + ba)] for C29- and C30-com-pounds, which equilibrate at somewhat higher thermal maturities

    (Seifert and Moldowan, 1980; Peters et al., 2005). The values for

    20 40 60 80 100 120

    d

    c

    b e

    a

    Pr

    C25

    Ph

    C17

    Retention Times

    Re

    lativea

    bun

    dance

    Fig. 3. Representative total ion chromatogram of saturated hydrocarbon fractions of the crude oils, showing distribution of n-alkanes (n-C1037) and isoprenoids (a: 2,6-

    dimethylundecane; b: 2,6,10-trimethylundecane (nor-farnesane); c: 2,6, 10-trimethyldodecane (farnesane); d: 2,6,10-trimethyltridecane; e: 2,6,10,-trimethylpentadecane

    (nor-pristane); Pr: pristane and Ph: phytane).

    -25.0

    -24.0

    -23.0

    -22.0

    -21.0

    -20.0

    -27.0 -26.0 -25.0 -24.0 -23.0 -22.0

    13CSats ()

    13CAros

    ()

    A

    B

    C

    (a)

    -150

    -140

    -130

    -120

    -110

    -100

    -26.0 -25.0 -24.0 -23.0 -22.0 -21.0

    13Caver()

    D

    aver

    ()

    B

    A

    C

    (b)

    Fig. 4. Plots to delineate the three oil groups from the Potwar Basin (a) d13Csatsversus d13Caros and (b) d

    13Caver versus dDaver from average values of d13C and dD of

    saturated and aromatic hydrocarbon fractions.

    M. Asif et al. / Organic Geochemistry 42 (2011) 12261240 1231

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    8/16

    Author's personal copy

    Table 2

    Saturated and aromatic hydrocarbon biomarker thermal maturity parameters.

    Sample name Ts/(Ts + Tm)a ab/(ab + ba),C29-Hop

    b

    ab/(ab + ba),C30-Hop

    c

    S/(S+ R)

    C32-Hopd

    bb/(aa + bb)C29-Ster

    e

    S/(S+ R)

    C29-sterf

    DNR-1g TNR-1h TNR-2i Rcbj (%) MPI-1k Rc

    l (%)

    P1 0.53 0.83 0.81 0.62 0.59 0.41 6.8 1.04 0.94 1.02 0.75 0.85

    P2 0.40 0.93 0.86 0.61 0.66 0.43 7.6 1.43 1.04 1.02 1.07 1.04P3 0.36 0.92 0.86 0.62 0.65 0.41 7.2 1.38 1.02 0.99 1.02 1.01

    P4 0.36 0.94 0.87 0.60 0.64 0.44 6.3 1.61 1.03 1.03 0.80 0.88

    P5 0.37 0.93 0.88 0.58 0.63 0.45 5.7 1.44 0.98 1.00 0.90 0.94

    P6 0.38 0.93 0.90 0.56 0.61 0.45 6.2 1.64 1.05 0.96 0.85 0.91

    P7 0.41 0.94 0.87 0.57 0.63 0.45 7.8 1.50 1.00 1.08 0.92 0.95

    P8 0.40 0.93 0.88 0.59 0.61 0.47 8.1 1.85 1.12 1.15 1.04 1.03

    P9 0.45 0.92 0.88 0.61 0.64 0.47 8.0 1.61 1.12 1.11 0.89 0.93

    P10 0.35 0.96 0.89 0.57 0.59 0.48 8.4 1.96 1.14 0.97 0.91 0.94

    P11 0.31 0.94 0.90 0.59 0.59 0.46 5.7 2.31 1.24 1.07 0.85 0.91

    P12 0.38 0.93 0.85 0.60 0.62 0.45 6.9 2.17 1.18 1.07 0.90 0.94

    P13 0.45 0.97 0.92 0.60 0.63 0.45 5.5 1.39 0.95 0.95 1.26 1.16

    P14 0.73 1.00 1.00 0.64 0.63 0.47 6.8 1.25 0.95 0.97 1.14 1.08

    P15 0.67 1.00 1.00 0.63 0.61 0.45 7.7 1.29 0.91 1.01 1.16 1.10

    P16 0.66 1.00 1.00 0.63 0.61 0.43 7.1 1.23 0.96 0.98 1.14 1.08

    P17 0.70 1.00 0.92 0.61 0.63 0.44 7.4 1.39 1.00 1.00 1.07 1.04

    P18 0.70 0.90 0.92 0.61 0.62 0.47 8.5 1.23 0.94 0.96 1.09 1.05

    : not determined.a Ts/(Ts + Tm), 18a(H)-22,29,30-trisnorneohopane/(18a(H)-22,29,30-trisnorneohopane + 17a(H)-22,29,30-trisnorhopane).b ab/(ab + ba) C29 Hop: 17a(H),21b(H)-30-norhopane/(17a(H),21b(H)-30-norhopane + 17b(H),21a(H)-30-norhopane).c ab/(ab + ba), C30 Hop: 17a(H),21b(H)-hopane/(17a(H),21b(H)-hopane + 17b(H),21a(H)-hopane).d S/(S + R), C32 Hop, 22S/(22S + 22R), 17a(H)-bishomohopane.e (bb/aa + bb) C29-Ster: 14b(H),21b(H)/[14a(H),21a(H) + 14b(H),21b(H)] 20R-ethylcholestane.f S/(S+ R) C29 ster: 20S/(20S + 20R) 14a(H),21a(H)-ethylcholestane.

    g DNR-1: dimethylnaphthalene ratio 1 (2,6- + 2,7-DMN/1,5-DMN), Radke, 1987.h TNR-1: trimethylnaphthalene ratio 1 (2,3,6-TMN/1,4,6- + 1,3,5-TMN), Alexander et al., 1985.i TNR-2: trimethylnaphthalene ratio 2 (2,3,6- + 1,3,7-TMN)/1,4,6- + 1,3,5- + 1,3,6-TMN).

    j Rcb: 0.40 + 0.6 (TNR-2) Radke et al., 1986.k MPI-1: methylphenanthrenes index {1.5 [3-MP + 2-MP]/[P + 1-MP + 9-MP]}, Radke et al., 1982.l Rc: calculated vitrinite reflectance (0.6 MPI-1 + 0.4), Radke and Welte, 1983.

    Table 3

    Source OM parameters for Potwar Basin oils.

    Sample

    name

    C19/

    (C19 + C23)

    TTa

    C24TeT/

    (C24TeT + C23TT)b

    C23TT/C30Hopc

    C24-

    TeT/

    C30Hopd

    C28TT/

    C30-

    Hope

    C29/

    C30 abhopf

    Dia/

    hop

    C30g

    C30Dia/

    C29Tsh

    C31(R+ S)/

    C30 hopi

    Steranes/

    hopanesjDia/

    ster

    C27k

    Dia/

    Ster

    C29l

    Total

    Dia/

    Sterm

    C27abb

    C28abb

    C29abb

    P1 0.88 0.77 0.07 0.24 0.00 0.65 0.38 2.23 0.77 0.29 0.53 0.97 1.15 43 18 39

    P2 0.49 0.47 0.40 0.35 0.00 0.58 0.26 1.44 0.75 0.23 0.76 0.60 0.91 37 13 50

    P3 0.46 0.49 0.31 0.30 0.00 0.55 0.22 1.78 0.75 0.24 0.53 0.63 0.80 39 21 40

    P4 0.40 0.56 0.32 0.42 0.11 0.74 0.15 0.83 0.94 0.23 0.55 0.57 0.76 40 17 43

    P5 0.40 0.55 0.37 0.45 0.12 0.78 0.15 0.75 0.97 0.28 0.50 0.51 0.73 41 16 43

    P6 0.39 0.54 0.37 0.44 0.10 0.76 0.17 0.86 0.89 0.27 0.59 0.54 0.76 40 17 43

    P7 0.38 0.52 0.46 0.49 0.17 0.75 0.22 1.13 0.90 0.34 0.50 0.60 0.76 40 20 40

    P8 0.41 0.57 0.33 0.44 0.10 0.65 0.15 0.80 0.82 0.27 0.50 0.49 0.77 48 8 45

    P9 0.45 0.52 0.28 0.30 0.09 0.69 0.18 0.64 0.86 0.35 0.61 0.68 0.82 38 22 40

    P10 0.33 0.61 0.24 0.38 0.08 0.79 0.11 0.72 0.88 0.23 0.44 0.40 0.65 42 12 46

    P11 0.30 0.63 0.23 0.39 0.07 0.87 0.07 0.56 0.96 0.20 0.45 0.38 0.63 42 12 47

    P12 0.38 0.55 0.43 0.52 0.13 0.75 0.20 1.24 0.98 0.33 0.55 0.59 0.83 39 19 42P13 0.44 0.55 0.38 0.47 0.11 0.75 0.19 0.83 0.83 0.35 0.55 0.58 0.74 41 19 40

    P14 0.49 0.51 1.93 2.03 0.64 0.64 0.86 1.34 0.89 1.24 0.65 0.78 0.97 45 18 37

    P15 0.34 0.44 1.54 1.22 0.71 0.61 0.32 1.44 0.44 1.50 0.38 0.06 0.31 24 34 41

    P16 0.33 0.45 1.16 0.93 0.55 0.51 0.29 1.59 0.44 1.45 0.40 0.19 0.40 22 34 44

    P17 0.36 0.40 0.86 0.58 0.41 0.48 0.26 1.29 0.47 0.74 0.43 0.26 0.35 31 32 37

    P18 0.32 0.44 0.50 0.40 0.27 0.57 0.23 0.91 0.50 0.58 0.38 0.27 0.34 30 31 39

    a C19/(C19 + C23) TT, C19-tricyclic terpane/(C19-tericyclic terpane + C23 tricyclic terpane).b C24TeT/(C24TeT + C 23TT), C24-tetracyclic terpane/(C24-tetracyclic terpane + C23 tricyclic terpane).c C23 TT/C30-hop: C23 tricyclic terpane/C30-ab hopane.d C24 TeT/C30-hop: C24 tetracyclic terpane/C30-ab hopane.e C28TT/C30-Hop, C28 tricyclic terpane/C30 ab hopane.f C29/C30 ab hop, C29 30-norhopane/C30 ab-hopane.

    g Dia/Hop C30, C30 ba diahopane/C30 ab-hopane.h C30 Dia/C29 Ts: C30, ba diahopane/18a(H)-30-norneohopane.i C31 (R+ S)/C30 hop, C31 ab-homohopane (22S + 22R)/C30 ab-hopane.

    j Steranes/hopanes: total steranes/total hopanes.

    k Dia/ster C27: ba/(aa + bb) cholestane.l Dia/ster C29: ba/(aa + bb) ethylcholestane.

    m Total dia/ster, total ba steranes/(ab + aa) steranes; C27, C28, C29 abb steranes.

    1232 M. Asif et al. / Organic Geochemistry 42 (2011) 12261240

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    9/16

    Author's personal copy

    both C29- and C30-moretanes, [ab/(ab + ba)] ratios are in the rangeof 0.81 to 1.0 (mostly >0.9, Table 2) typical for oils generated from

    mature source rocks (cf. George et al., 2004). The plot of hopane

    based maturity parameters between C29- and C30-ab/(ab + ba) isshown in Fig. 5a (cf. George et al., 2001), where most of the oil sam-

    ples fall close to equilibrium and higher than equilibrium level

    indicating higher thermal maturity except P1, which shows rela-

    tively low thermal maturity. The slight difference in ab/(ab + ba)ratios for the Potwar Basin oils shows the affects of source and

    depositional environment variations on these values (Rullktter

    and Marzi, 1988; Isaksen and Bohacs, 1995). The Ts/(Ts + Tm) ratio

    continuously varies from the immature to postmature (Peters

    et al., 2005). The values for Ts/(Ts + Tm) ratio range from 0.31

    0.73 for the oils indicating an immature to mature range of thermal

    maturity, however a narrow range of this ratio is observed for indi-

    vidual groups. For example, group B show a Ts/(Ts + Tm) ratio in

    the range of 0.310.45 while group C shows values in the range

    of 0.670.70. The single Group A oil (P1) has an intermediate value

    (0.53, Table 2). One oil sample (P14, Table 2) from group B shows a

    higher value for Ts/(Ts + Tm), 0.73, indicating significant effects of

    source and depositional environments (Moldowan et al., 1986).

    Different Ts/(Ts + Tm) values for each group of oils indicate that

    this ratio is controlled by factors other than thermal maturity, most

    probably source and depositional environments (Moldowan et al.,

    1986) that are shown to affect the source OM of Potwar Basin oils

    (see below).The sterane based thermal maturity parameters such as 20S/

    (20S+ 20R) ethylcholestane and bb/(aa + bb) ethylcholestanerange tightly at 0.410.48 and 0.590.66 respectively, supporting

    a similar thermal maturity for these samples while the equilibrium

    occurs between 0.520.55 and 0.670.7,1 respectively (Seifert and

    Moldowan, 1986). These values suggest that none of the oils have

    reached full maturity with respect to equilibrium values suggested

    by Seifert and Moldowan (1986). However these lower than equi-

    librium values for both sterane parameters are consistent with the

    peak oil generation window for the Potwar Basin oils (Peters et al.,

    2005). Despite the fact that 20S/(20S+ 20R) ratio is very useful to

    indicate thermal maturity, factors other than thermal maturity

    can affect this ratio. For example, reversal of this ratio within

    highly mature intervals could be responsible for the lower values

    (cf. Bishop and Abbott, 1993; Edwards et al., 1997).

    It has been shown that many biomarker maturity parameters

    reach equilibrium at the onset of the oil window and therefore

    may not be useful for highly mature oils and condensates (van

    Graas, 1990). In this scenario, parameters based on aromatic HCs

    may be more effective for evaluation of thermal maturity. The

    methyl phenanthrene index (MPI-1; Radke et al., 1982) appears

    to be useful to estimate vitrinite reflectance (Radke et al., 1982;

    Radke, 1988). The MPI-1 and calculated vitrinite reflectance (Rc)values from Potwar Basin oils are reported in Table 2. The MPI-1

    is in the range of 0.751.26 and Rc in the range of 0.851.15%.

    The Rc for P1 (0.85%, Table 2) suggests a maturity equivalent to

    0.75

    0.80

    0.85

    0.90

    0.95

    1.00

    0.75 0.80 0.85 0.90 0.95 1.00

    /(+), C29

    -Hop

    /(

    +

    ),C30

    -Hop

    Early oil generation

    Equilibrium

    (a)

    0.6

    0.7

    0.8

    0.9

    1.0

    1.1

    1.2

    0.6 0.7 0.8 0.9 1.0 1.1 1.2

    Rcb(%)

    Rc

    (%)

    Late

    Peak

    Early

    (b)

    Fig. 5. (a) Hopane maturity parameter plot between C29 versus C30 ofab/(ab + ba)(cf. George et al., 2001) (b) calculated vitrinite reflectance diagram from Rcb (TNR-2;

    Radke et al., 1986) and Rc (MPI-1; Radke and Welte, 1983) showing different

    thermal maturation stages of oil generation window.

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    10 20 30 40 50

    API gravity ()

    Pr/

    Ph

    B

    A

    C

    (b)

    0

    1

    2

    3

    4

    0.0 1.0 2.0 3.0 4.0

    Pr/Ph

    D

    BT/P

    1A: marine Carbonate

    1B: marine carbonate and marl

    2: Lacustrine hypersaline

    3: marine shale and lacustrine

    4:fluvio-deltaic shale

    Hughes et al, 1995

    1A

    1B

    2 3 B & C 4A

    (a)

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    0.0 0.2 0.4 0.6 0.8 1.0

    Diahopane/hopane, C30

    Dias

    terane

    /steranes,

    C29(c) A

    B

    C

    P14

    Fig. 6. (a) Pr/Ph versus DBT/P indicating lithology and depositional environment

    (Hughes et al., 1995) (b) a cross plot of API gravity and Pr/Ph separating the Potwar

    Basin oils into three groups, (c)C30 diahopane/hopaneversus C29 diasterane/sterane

    ratios.

    M. Asif et al. / Organic Geochemistry 42 (2011) 12261240 1233

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    10/16

    Author's personal copy

    peak oil generation while all other Potwar Basin oils indicate a

    maturity of late oil generation window (>0.9) (cf. Radke, 1987).

    Methylnaphthalene thermal maturity parameters are also listed

    in Table 2. The dimethylnaphthalene ratio (DNR-1, Table 2) >5.5

    (Table 2, mostly $78) clearly indicates that the thermal maturity

    of the Potwar Basin oils has reached the late oil generation win-

    dow. The trimethylnaphthalene ratio 1 (TNR-1, Table 2) has been

    calibrated with the sterane isomerisation ratio (20S/20R), showing

    that the sterane isomerisation ratio of oils reaches equilibrium

    when TNR-1 ratio is >1 (Alexander et al., 1985). TNR-1 values for

    Potwar Basin oils are >1 (mostly >1.2) for all samples, indicating

    that the maturity of source rocks generating these oils reached to

    higher than the peak oil generation window (cf. Alexander et al.,

    1985). Similarly, the trimethylnaphthalene ratio 2 (TNR-2, Table

    2) is another useful aromatic HC thermal maturity parameter

    which has been calibrated with mean vitrinite reflectance (Ro)

    and shows good agreement with increase in thermal maturity

    (Radke et al., 1986). The TNR-2 value (0.91.2, Table 2) and calcu-

    lated vitrinite reflectance Rcb values (>0.95, Table 2) from TNR-2

    indicate thermal maturity of the oil samples from the Potwar Basin

    reached the late oil generation window (Radke et al., 1986). A cross

    plot (Fig. 5b) of calculated vitrinite reflectance values i.e. Rcb (TNR-2) versus Rc (MPI-1) clearly indicates that the thermal maturity of

    Potwar Basin oils reaches to the late oil generation window. It has

    been shown that the biomarker maturity parameters reveal early

    to peak oil generation window for Potwar Basin oils, however these

    parameters mostly reached equilibrium values so showed limited

    application for maturity assessment. Finally, it is concluded that

    equivalent vitrinite reflectance calculated from aromatic HC matu-

    rity parameters reveal source rocks of Potwar Basin oils reached

    maturities of late oil generation window.

    A few anomalies are observed in the alkylnaphthalene maturity

    parameters. For example, TNR-1 shows a wide range of values from

    1.042.31 although most of the values lie between 1 and 2. High

    values (TNR-1 >2.0, Table 2) for some of the oils are probably

    due to secondary effects such as biodegradation. Affects of biodeg-

    radation on alkylnaphthalenes have been shown to affect different

    isomers and thus different susceptibilities towards biodegradation

    (Asif et al., 2009) and thermal maturity parameters are altered

    when using certain isomers in thermal maturity ratio calculations

    (van Aarssen et al., 1999; Obermajer et al., 2004).

    3.4. Lithology and depositional environments

    The crude oils listed in Table 1 were examined for lithology and

    depositional environment using aliphatic and aromatic biomarker

    parameters. The Pr/Ph ratio shows a range of values from 12 (ex-cept sample P1, Pr/Ph = 3.2) for the Potwar Basin oils (Table 1).

    DBT/P ratio is a good indicator of lithology and the values for the

    ratio are

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    11/16

    Author's personal copy

    Potwar Basin oils (Hughes et al., 1995). The Pr/Ph and DBT/P data

    from Potwar Basin oils are plotted based on the Hughes diagram

    (Hughes et al., 1995) and these results are shown in Fig. 6a. The

    largest set of oils from the Potwar Basin (group B and C) are shown

    to originate from marine shale and lacustrine source rocks (Fig. 6a)

    while the single group A oil (P1) indicates a highly oxic fluvio-del-

    taic depositional environment. However no molecular evidence is

    observed for a lacustrine depositional environment of source rocks

    for Potwar Basin oils. These results indicate that most of the oils in

    the Potwar Basin were generated from marine shale source rocks

    and the single oil of group A was generated from fluvio-deltaic

    source rocks.

    Depositional environments based on Pr/Ph ratio and HC compo-

    sitional variations based on API gravity are used to differentiate the

    Potwar Basin oils. A cross plot of API gravity and Pr/Ph is shown in

    Fig. 6b. The single oil positioned in the right top corner of the plot

    indicates higher Pr/Ph and API values consistent with an oxic depo-

    sitional environment and light oil typically generated from

    terrigenous OM. An interesting feature of the plot (Fig. 6b) is theseparation of the group B and C oils that were shown to be similar

    in the Hughes diagram (Fig. 6a). The group B oils show lower Pr/Ph

    (mostly

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    12/16

    Author's personal copy

    (Peters et al., 2005) and this ratio for group A is very high (2.23, Ta-

    ble 3) indicating an highly oxic depositional environment.

    Although data for both groups B and C oils imply marine source

    rocks based on lithology and environments of deposition (see

    above), relative abundance of hopanes differentiates the groups

    (Fig. 6c): group B is more reduced than group C. The C30 diaho-pane/C29Ts ratio (61.0 for most of oils) and C31 (R + S) hopane/

    C30 hopane ratio (>0.75) for group B oils indicate more dysoxic

    depositional settings for OM than for group C oils (cf. Peters and

    Moldowan, 1993). An anomalously high value relative to the group

    B oils is observed for the P14 sample in the C 30 diahopane/hopane

    versus C29 diasteranes/sterane diagram (Fig. 6c). The reason for this

    anomaly could be related to the depositional environment (high

    Pr/Ph ratio for this sample) and source organic facies (Moldowan

    et al., 1991; Peters et al., 2005 and references therein).

    3.5. Source of OM

    The sources of OM were determined from the distributions of

    tricyclic, tetracyclic and pentacyclic terpanes and steranes. Fig. 7

    shows m/z 191 chromatograms of the representative oil samples

    from all delineated groups from the Potwar Basin. The source bio-

    marker parameters are listed in Table 3. The group A oil shows a

    significantly lower relative abundance of tricyclic and tetracyclic

    terpanes except for C19 tricyclic terpane and C24 tetracyclic terpane

    (Fig. 7), both indicators of terrigenous OM (Philp and Gilbert, 1986;

    Grice et al., 2001; Peters et al., 2005; Volk et al., 2005; Nabbefeld

    et al., 2010a). The oil correlation diagram of C19/(C19 + C23) tricyclic

    versus C24 tetracyclic/(C24 tetracyclic + C23 tricyclic terpane) is

    shown in Fig. 8a where group A indicates clearly a different source

    of OM compared to the other groups (cf. Edwards et al., 1997; Volk

    et al., 2005). The presence of the aromatic plant biomarker retene

    also tends to support a terrigenous source of OM for the group A

    oil, although this biomarker can also be ascribed to algae (Nabbe-feld et al., 2010b and references therein). The lowvalues for the C23tricyclic/C30 hopane (

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    13/16

    Author's personal copy

    compared to regular steranes consistent with a clastic source rock

    common in deltaic/oxic depositional settings (Peters et al., 2005).

    The regular steranes show almost equal abundance to the diaster-

    anes in group B oils while in group C oils the former is significantly

    less abundant than the latter. These distribution differences are

    indicating that the group B oils were generated from more clasticsource rocks than the group C oils, consistent with above results

    (Section 3.4). The C29 and C28 regular steranes are present in signif-

    icantly higher abundance than other steranes and diasteranes in

    the group C oils (Fig. 9) indicating the significant contribution of

    terrigenous and algal OM, respectively. These interpretations of

    terrigenous and algal contribution to group C oils are consistent

    with presence of higher concentrations of tricyclic terpanes than

    hopanes (see below) (cf. Philp et al., 1989) and high API values

    for these oils (Peters et al., 2005). The regular sterane correlation

    diagram between C27/C29 aaa versus C28/C29 abb steranes isshown in Fig. 8b which differentiates the Potwar Basin oils into

    the same three groups where group C shows higher C28 regular

    steranes representing a higher algal input. The distribution of rear-

    ranged steranes also differentiate B from C oils ( Table 3). Group Bshows higher C27 and C29 ba/(bb + aa) ratios (>0.45 and >0.4,respectively) indicating comparatively more marine clastic source

    input than group C oils.

    The representative 191 ion chromatogram for group C oils

    shows significantly higher abundance of tricyclic terpanes than ho-

    panes (Fig. 7c). The C23 tricyclic terpane is the most abundant m/z

    191 compound in the group C oils and the presence of extended tri-

    cyclic terpanes up to C41 and possibly higher is an important fea-

    ture of these oils, which can be used to differentiate group C

    from groups A and B. Various organisms have been suggested as

    the source of tricyclic terpanes in oils and bitumens (Ourisson

    et al., 1982; Volkman et al., 1989; Peters and Moldowan, 1993;

    Simoneit et al., 1993) and the ubiquitous occurrence of tricyclic

    terpanes in sedimentary OM of varying ages has been related to

    numerous source origins (Farrimond et al., 1999). The occurrence

    of tricyclic terpanes in the Potwar Basin C oils is probably related

    to an algal source which is supported by comparative abundance

    of regular C28 steranes in group C oils (Table 3; Fig. 8b). Higher

    abundance of tricyclic terpanes than hopanes has also been relatedto terrigenous input (Philp et al., 1989) that is consistent with

    higher abundance of C29 steranes in group C oils. Similarly, a higher

    total sterane/hopane ratio >0.6 ($1.0 for most of the oils) may re-

    flect a greater eukaryotic input (both algae and terrigenous) to

    group C oils source rocks. The C23 tricyclic/C30 hopane and C24 tet-

    racyclic/C30 hopane ratios (0.51.5 and 0.41.2, respectively, Table

    3) indicate typically marine OM input for group B and C oils where

    higher values for group C oils shows a higher marine input (algal

    input) (cf. Peters et al., 2005). This differentiation in numerous

    source OM parameters shows a different origin for source input

    for each group and it is concluded that petroleum from Potwar Ba-

    sin contained three source oil families.

    The data set presented here indicates a few contradictions with

    respect to the classification of crude oils from the Potwar Basin. Inthe above correlations (Fig. 6b and c), sample P14 (group B) posi-

    tioned with group C oils. Similarly many source parameters for

    the same sample such as C23 tricyclic/C30 hopane, C24 tetracyclic/

    C30 hopane, C23 tricyclic/C28 tricyclic terpane and steranes/hopanes

    place this oil close to the group C oils (Table 3). Moreover, the C27,

    C28 and C29 abb steranes show a different distribution trends foreach group i.e. group A shows C27P C29 ) C28, group B shows

    C29P C27 ) C28 and group C shows C29 > C28 > C27 (Table 3). The

    relative distribution profile ofabb steranes (C27P C29 ) C28) fromgroup A reveals higher lacustrine source input while it has been

    shown that this group has a terrigenous origin. The abb steranesprofile from group B oils shows higher C29 compounds, indicating

    terrigenous input in contrast to results drawn from this study that

    P1

    P17

    P14,P18

    P15P16

    P2

    P3

    P8

    P13P7

    P11,P12

    P10

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    Pr/n-C17

    Ph/n-C18

    Biodegradation(a)

    1.2

    P1

    P17

    P9

    P14,P18

    P15,P16

    P2

    P3

    P8

    P13P4-5

    P11-12

    P6

    0

    10

    20

    30

    40

    50

    60

    0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

    0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

    Pr/n-C17

    APIgravity()

    Biodegradation

    (b)

    P6

    Fig. 11. (a) Pr/n-C17 and Ph/n-C18 (b) API value versus Pr/n-C17 showing a decrease in the API gravity of crude oils with increasing biodegradation.

    M. Asif et al. / Organic Geochemistry 42 (2011) 12261240 1237

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    14/16

    Author's personal copy

    this group is marine. Reasons for these contradictory findings could

    be solved by evaluating the source rocks of Potwar Basin oils, but

    source to oil correlation along with migration affects, reservoir

    connectivity and oil mixing in Potwar Basin are still to be

    determined.

    3.6. Extent of biodegradation

    Biodegradation is a process that alters the molecular composi-

    tion and bulk properties (API gravity) of petroleum and sediments

    (Connan, 1984; Fisher et al., 1998). A number of commonly used

    parameters have been used to assess the extent/level of biodegra-

    dation in the Potwar Basin oils. Representative TICs of the satu-

    rated fractions from each group of oils are shown in Fig. 10. The

    distribution pattern of saturated HC fraction from group A and C

    shows the presence of a full suite of n-alkanes and the absence

    of any unresolved complex mixture (UCM), indicating no biodegra-

    dation. While TIC from the a representative group B oil shows a

    substantial UCM in the saturated fraction and a lack of n-alkanes

    indicating that these oils have been biodegraded andthe remainingfraction has become enriched in high molecular weight unresolved

    components. Similarly, isoprenoids show resistance to biodegrada-

    tion compared to the n-alkanes because the n-alkanes are removed

    faster than isoprenoids during biodegradation (Peters et al., 2005).

    Hence isoprenoid/n-alkane ratios from saturated fractions increase

    with an increase in biodegradation (Winters and Williams, 1968)

    and Pr/n-C17 and Ph/n-C18 ratios > 1 typically indicates the effect

    of biodegradation on crude oils. The plot of Pr/n-C17 versus Ph/n-

    C18 (Fig. 11a) shows a trend consistent with biodegradation; these

    ratios increase with increasing biodegradation. The API gravity is a

    bulk property that directly relates to gross compositions of crude

    oils. The Potwar Basin crude oils show a wide range of API gravities

    (1648; Table 1). A plot of API gravity versus Pr/n-C17 (Fig. 11b)

    shows an inverse relationship, a high Pr/n-C17 and lower API grav-

    ity (Fig. 11b) indicative of the oils affected by biodegradation. The

    results show that extent of biodegradation for some of the crude

    oils in this study reaching up to a level of 3 on the Wenger et al.

    (2001) scale. The extent of biodegradation of each crude oil from

    the Potwar Basin is represented with level of biodegradation in Ta-

    ble 1. It is observed that some of the oils from group B are affected

    by minor biodegradation while group A and C are non-biodegraded

    (Fig. 11a and b).

    The representative group B chromatogram shows a high UCM

    but also the presence of n-alkanes (Fig. 10). This type of saturated

    HC profile indicates the possibility of mixing of biodegraded and

    non-biodegraded crude oils in the reservoir. Assessment of biodeg-

    radation and in-reservoir mixing in the Upper Indus oils (Potwar

    Basin) has been reported using biomarker parameters (Asif et al.,

    2009).

    4. Conclusions

    Geochemical characterisation and classification of the Potwar

    Basin crude oils were performed using biomarker and stable iso-

    tope distributions. Saturated HC biomarkers indicate at least early

    to peak oil generation window of thermal maturity while aromatic

    HC parameters and calculated vitrinite reflectance from these

    parameters reveal late oil generation window thermal maturity

    for Potwar Basin oils. Stable carbon and hydrogen isotopes of sat-

    urated and aromatic HC fractions delineated three groups in the

    Potwar Basin oils. These three groups of crude oils are differenti-

    ated based on source OM, depositional environment and lithology.

    Group A oil suggests terrigenous source OM generated from

    fluvio-deltaic source rocks deposited in an oxic depositional

    environment. Group A oil shows more negative (isotopically

    lighter) d13C of both saturated and aromatic HC fractions com-

    pared to all other oils. The abundance of C19 tricyclic and C24tetracyclic terpanes along with a higher abundance of a diag-

    nostic aromatic HC biomarker, retene, suggests a terrigenous

    source OM for group A oil. The other oils from the Potwar Basin analysed in this study are

    marine in origin. However d13C and dD of bulk HC fractions and

    based on tricyclic, tetracyclic and pentacyclic terpane and ster-

    ane distributions separate these oils into groups B and C. Group

    B oils show the heaviest d13C for both saturated and aromatic

    HC fractions. Some of the group B crude oils are biodegraded

    (level 23) and the OM of this group was deposited in a subox-

    ic/dysoxic depositional environment.

    Group C oils are typically non-biodegraded, mature crude oils

    generated from source OM rich in algae with terrigenous input

    deposited under marine oxic environments, which is supported

    by the presence of extended tricyclic terpanes and regular ster-

    anes. This group shows light d13C in the saturated HC fraction

    relative to group B oils; however d

    13

    C of the aromatic fractionof group B and C are not very different from one another.

    Acknowledgements

    The authors thank Mr. G. Chidlow for assistance with GCMS

    and S. Wang for bulk isotope analysis and maintenance. The Higher

    Education Commission, Islamabad, Pakistan is thanked for an IRSIP

    fellowship and a travel award Grant (IRSIP-5-Ps-20) to MA. KG

    acknowledges the ARC for a QEII fellowship (DP0211875,

    DP0877167). The authors thank the following exploration compa-

    nies for providing oil samples: Oil and Gas Development Coopera-

    tion Ltd. (OGDCL), Islamabad, Pakistan Petroleum Ltd. (PPL) and

    Pakistan Oilfields Ltd. (POL). J. Curiale and H. Huang are acknowl-

    edged for constructive reviews of the initial version of this paper.

    Associate EditorMaowen Li

    References

    Aamir, M., Siddiqui, M.M., 2006. Interpretation and visualization of thrust sheets in

    a triangle zone in eastern Potwar, Pakistan. The Leading Edge January 2006, pp.

    2437.

    Ahmed, W., Alam, S., 1990. Organic geochemistry of crude oil from Potwar-Kohat

    region. Pakistan Journal of Hydrocarbon Research 2, 115.

    Ahmed, W., Alam, S., 2007. Organic geochemistry and source rock characteristics of

    Salt Range Formation, Potwar Basin, Pakistan. Pakistan Journal of Hydrocarbon

    Research 17, 3759.

    Alexander, R., Kagi, R.I., Rowland, S.J., Sheppard, P.N., Chirila, T.V., 1985. The effectsof thermal maturity on distributions of dimethylnaphthalenes and

    trimethylnaphthalenes in some ancient sediments and petroleums.

    Geochimica et Cosmochimica Acta 49, 385395.

    Andrusevich, V.E., Engel, M.H., Zumberge, J.E., Brothers, L.A., 1998. Secular, episodic

    changes in stable carbon isotope compositions of crude oils. Chemical Geology

    152, 5972.

    Asif, M., Grice, K., Fazeelat, T., Dawson, D., 2008. Oiloil correlation in the Upper

    Indus Basin (Pakistan) based on biomarker distributions and compound-specific

    d13C and dD. In: 15th Australian Organic Geochemistry Conference, 812th

    September, National Wine Centre, Adelaide, SA, Australia.

    Asif, M., Grice, K., Fazeelat, T., 2009. Assessment of petroleum biodegradation using

    stable hydrogen isotopes of individual saturated hydrocarbons and polycyclic

    aromatic hydrocarbon distributions in oils from the Upper Indus Basin,

    Pakistan. Organic Geochemistry 40, 301311.

    Bishop, A.N., Abbott, G.D., 1993. The interrelationship of biological marker maturity

    parameters and molecular yields during contact metamorphism. Geochimica et

    Cosmochimica Acta 57, 36613668.

    Chung, H.M., Rooney, M.A., Toon, M.B., Claypool, G.E., 1992. Carbon isotope

    compositions of marine crude oils. American Association of PetroleumGeologist Bulletin 76, 10001007.

    Collister, J.W., Wavrek, D.A., 1996. d13C compositions of saturated and aromatic

    fractions of lacustrine oils and bitumen: evidence for water column

    stratification. Organic Geochemistry 24, 913920.

    1238 M. Asif et al. / Organic Geochemistry 42 (2011) 12261240

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    15/16

    Author's personal copy

    Connan, J., 1984. Biodegradation of crude oils in reservoirs. In: Brooks, J., Welte, D.H.

    (Eds.),Advances in PetroleumGeochemistry, vol. 1. Academic Press, London, pp.

    299335.

    Dawson, D., Grice,K., Alexander, R., 2005. Effect on maturation on theindigenous dD

    signatures of individual hydrocarbons in sediments and crude oils from the

    Perth Basin (Western Australia). Organic Geochemistry 36, 95104.

    Dawson, D., Grice, K., Edwards, D.S., Alexander, R., 2007. The effect of source and

    maturity on the stable isotopic compositions of individual hydrocarbons insediments and crude oils from the Vulcan Sub-basin, Timor Sea, Northern

    Australia. Organic Geochemistry 38, 10151038.

    Edwards, D.S., Summons, R.E., Kennard, J.M., Nicoll, R.S., Bradshaw, M., Foster, C.B.,

    OBrien, G.W., Zumberge, J.E., 1997. Geochemical characteristics of Palaeozoic

    petroleum systems in Northwestern Australia. Australian Petroleum Production

    & Exploration Association Journal 37, 351379.

    Farah, A., Abbas, G., DeJong, K.A., Kees, A., Lawrence, R.D., 1984. Evolution of the

    lithosphere in Pakistan. Tectonophysics 105, 207227.

    Farrimond, P., Bevan, J.C., Bishop, A.N., 1999. Tricyclic terpane maturity parameters:

    response to heating by an igneous intrusion. Organic Geochemistry 30, 1011

    1019.

    Fazeelat, T., 1994. 8,14-secohopanes and related compounds: their occurrence in

    sediments and applications to oil-source correlation. PhD thesis, Curtin

    University of Technology, Perth, Australia.

    Fazeelat, T., Jalees, M.I., Bianchi, T.S., 2010. Source rock potential of Eocene,

    Paleocene and Jurassic deposits in the subsurface of the Potwar basin, northern

    Pakistan. Journal of Petroleum Geology 33, 8796.

    Fazeelat, T., Asif, M., Jalees, M.I., Saleem, A., Saleem, H., Nazir, A., Nasir, S., Nadeem,

    S., 2011. Source correlation between biodegraded oil seeps and a commercial

    crude oil from Punjab Basin, Pakistan. Journal of Petroleum Science and

    Engineering 77, 19.

    Fisher, S.J., Alexander, R., Kagi, R.I., Oliver, G.A., 1998. Aromatic hydrocarbons as

    indicator of biodegradation in north western Australian reservoirs. In: Purcell,

    P.G., Purcell, R.R. (Eds.), The Sedimentary Basins of Western Australia 2:

    Proceedings of Petroleum Exploration Society of Australia Symposium. Perth,

    WA, pp. 185194.

    George, S.C., Ruble, T.E., Dutkiewicz, A., Eadington, P.J., 2001. Assessing the maturity

    of oil trapped in fluid inclusions using molecular geochemistry data and

    visually-determined fluorescence colours. Applied Geochemistry 16, 451473.

    George, S.C., Lisk, M., Eadington, P.J., 2004. Fluid inclusion evidence for an early,

    marine-sourced oil charge prior to gas-condensate migration, Bayu-1, Timor

    Sea, Australia. Marine and Petroleum Geology 21, 11071128.

    Grice, K., Audino, M., Alexander, R., Boreham, C.J., Kagi, R.I., 2001. Distributions and

    stable carbon isotopic compositions of biomarkers in torbanites from different

    palaeogeographical locations. Organic Geochemistry 32, 1952110.

    Grice, K., Backhouse, J., Alexander, R., Marshall, N., Logan, G.A., 2005. Correlating

    terrestrial signatures from biomarker distributions, d13

    C, and palynology influvio-deltaic deposits from NW Australia (TriassicJurassic). Organic

    Geochemistry 36, 13471358.

    Grice, K., De Mesmay, R., Wang, S., de Glucina, A., 2008. New 5A molecular sieving

    method for CSIA. Organic Geochemistry 39, 284288.

    Grimalt, J.O., Campos, P.G., Berdie, L., Lopez-Quintero, J.O., Navarrete-Reyes, L.E.,

    2002. Organic geochemistry of the oils from the southern geological province of

    Cuba. Applied Geochemistry 17, 110.

    Hughes, W.B., Holba, A.G., Dzou, L.I., 1995. The ratio of dibenzothiophene to

    phenanthrene and pristane to phytane as indicators of depositional

    environment and lithology of petroleum source rocks. Geochimica et

    Cosmochimica Acta 59, 35813598.

    Hunt, J.M., 1995. Petroleum Geochemistry and Geology. W.H. Freeman and

    Company, New York, pp. 743.

    Isaksen, G.H., Bohacs, K.M., 1995. Geological controls of source rock geochemistry

    through relative sea level; Triassic, Barents Sea. In: Katz, B.J. (Ed.), Petroleum

    Source Rocks. Springer, Berlin, pp. 2550.

    Jiang, C., Li, M., 2002. Bakken/Madison petroleum systems in the Canadian Williston

    Basin, Part 3: geochemical evidence for significant Bakken-derived oils in

    Madison group reservoirs. Organic Geochemistry 3, 761787.Khan, M.A., Ahmed, R., Raza, H.A., Kemal, A., 1986. Geology of petroleum in Kohat-

    Potwar Depression, Pakistan. American Association of Petroleum Geologists

    Bulletin 70, 396414.

    Maslen, E., Grice, K., Gale, J., Hallmann, C., Horsfield, B., 2009. Crocetane: a potential

    marker of photic zone euxinia in thermally mature sediments and crude oils of

    Devonian age. Organic Geochemistry 49, 117.

    Maslen, E., Grice, K., Dawson, D., Le Metayer, P., Edwards, D., 2011. Stable carbon

    isotopic compositions of individual aromatic hydrocarbons as source and age

    indicators in oil from Western Australian Basins. Organic Geochemistry 42,

    387398.

    Moldowan, J.M., Sundararaman, P., Schoell, M., 1986. Sensitivity of biomarker

    properties to depositional environment and/or source input in the Lower

    Toarcian of SW, Germany. Organic Geochemistry 10, 915926.

    Moldowan, J.M., Fago, F.J., Carlson, R.M.K., Young, D.C., Duvne, G.A., Clardy, J.,

    Schoell, M., Pillinger, C.T., Watt, D.S., 1991. Rearranged hopanes in sediments

    and petroleum. Geochimica et Cosmochimica Acta 55, 33333353.

    Murphy, M.T.J., 1969. Analytical methods. In: Eglinton, G., Murphy, M.T.J. (Eds.),

    Organic Geochemistry. Springer, Berlin, pp. 7488.Murray, A.P., Boreham, C.J., 1992. Organic Geochemistry in Petroleum Exploration.

    Australian Geological Survey Organization, Canberra, 230pp.

    Nabbefeld, B., Grice, K., Twitchett, R.J., Summons, R.E., Hays, L., Bttcher, M.E., Asif,

    M., 2010a. An integrated biomarker, isotopic and palaeoenvironmental study

    through the Late Permian event at Lusitaniadalen, Spitsbergen. Earth and

    Planetary Science Letters 291, 8496.

    Nabbefeld, B., Grice, K., Summons, R.E., Hays, L., 2010b. Significance of polycyclic

    aromatic hydrocarbons (PAHs) in Permian/Triassic boundary sections. Applied

    Geochemistry 25, 13741382.

    Obermajer, M., Osadetz, K.G., Fowler, M.G., Silliman, J., Hansen, W.B., Clark, M.,

    2002. Delineating compositional variabilities among crude oils from central

    Montana, USA, using light hydrocarbon and biomarker characteristics. OrganicGeochemistry 33, 13431359.

    Obermajer, M., Osadetz, K.G., Fowler, M.G., Li, M., Snowdon, L.R., 2004. Variable

    alteration in heavy crude oils of west-central Saskatchewan, Canada. Organic

    Geochemistry 35, 469491.

    Ourisson, G., Albrecht, P., Rohmer, M., 1982. Predictive microbial biochemistry

    from molecular fossils to prokaryotic membranes. Trends in Biochemical

    Science 7, 236239.

    Pasadakis, N., Obermajer, M., Osadetz, K.G., 2004. Definition and characterization of

    petroleum compositional families in Williston Basin, North America using

    principal component analysis. Organic Geochemistry 35, 453468.

    Peters, K.E., Moldowan, J.M., 1993. The Biomarker Guide: Interpreting Molecular

    Fossils in Petroleum and Ancient Sediments. Prentice-Hall, Englewood Cliffs,

    New Jersey.

    Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide, second ed.

    Cambridge University Press.

    Philp, R.P., Gilbert, T.D., 1986. Biomarker distributions in oils predominantly derived

    from terrigenous source material. In: Leythaeuser, D., Rullktter, J. (Eds.),

    Advances in Organic Geochemistry 1985, Organic Geochemistry 10. Pergamon,

    Oxford, pp. 7384.

    Philp,R.P., Li, J., Lewis, C.A., 1989. An organic geochemical investigation of crude oils

    from shanganning, Jianghan, Chaidamu and Zhungeer basins, China. Organic

    Geochemistry 14, 447460.

    Quadri, V.N., Quadri, S.M.J.G., 1997. Indus Basin of Pakistan contains few wells. Oil

    and Gas Journal Archives 95, i. 24.

    Radke, M., 1987. Organic geochemistry of aromatic hydrocarbons. In: Brooks, J.,

    Welte, D. (Eds.), Advances in Petroleum Geochemistry. Academic Press,

    London.

    Radke, M., 1988. Application of aromatic compounds as maturity indicators in

    source rocks and crude oils. Marine and Petroleum Geology 5, 224236.

    Radke, M., Welte, D.H., 1983. The methylphenanthrene index (MPI): a maturity

    parameter based on aromatic hydrocarbons. In: Bjory, M. et al. (Eds.),

    Advances in Organic Geochemistry 1981. Wiley, Chichester, pp. 504512.

    Radke, M., Welte, D.H., Willsch, H., 1982. Geochemical study on a well in the

    Western CanadaBasin: relation of thearomatic distribution pattern to maturity

    of organic matter. Geochimica et Cosmochimica Acta 46, 110.

    Radke, M., Welte, D.H., Willsch, H., 1986. Maturity parameters based on aromatic

    hydrocarbons: influence of the organic matter type. In: Leythaeuser, D.,Rullktter, J. (Eds.), Advances in Organic Geochemistry 1985. Organic

    Geochemistry, vol. 10. Pergamon, Oxford, pp. 5163.

    Rullktter, J., Marzi, R., 1988. Natural and artificial maturation of biological markers

    in a Toarcian shale from northern Germany. Organic Geochemistry 13, 639

    645.

    Schoell, M., 1984. Stable isotopes in petroleum research. In: Brooks, J., Welte, D.H.

    (Eds.), Advances in Petroleum Geochemistry, vol. 1. Academic Press, London, pp.

    215245.

    Seifert, W.K., Moldowan, J.M., 1980. The effect of thermal stress on source rock

    quality as measured by hopane stereochemistry. In: Douglas, A.G., Maxwell, J.R.

    (Eds.), Advances in Organic Geochemistry 1979. Pergamon, Oxford, pp. 229

    237.

    Seifert, W.K., Moldowan, J.M., 1986. Use of biological markers in petroleum

    exploration. In: Johns, R.B. (Ed.), Methods in Geochemistry and Geophysics

    24, 261290.

    Simoneit, B.R.T., Schoell, M., Dias, R.F., Aquino Neto, F.R., 1993. Unusual carbon

    isotope compositions of biomarker hydrocarbons in a Permian tasmanite.

    Geochimica et Cosmochimica Acta 57, 42054211.

    Sofer, Z., 1984. Stable carbon isotope compositions of crude oils: applications tosource depositional environments and petroleum alteration. American

    Association of Petroleum Geologist Bulletin 68, 3149.

    Trolio, R., Grice, K., Fisher, S.J., Alexander, R., Kagi, R.I., 1999. Alkylbiphenyls and

    alkyldiphenylmethanes as indicators of petroleum biodegradation. Organic

    Geochemistry 30, 12411253.

    van Aarssen, B.G.K., Bastow, T.P., Alexander, R., Kagi, R.I., 1999. Distributions of

    methylated naphthalenes in crude oils: indicators of maturity, biodegradation

    and mixing. Organic Geochemistry 30, 12131227.

    van Graas, G.W., 1990. Biomarker maturity parameters for high maturities:

    calibration of the working range up to the oil/condensate threshold. Organic

    Geochemistry 16, 10251032.

    Volk, H., George, C.S., Middleton, H., Schofield, S., 2005. Geochemical comparison of

    fluid inclusion and present-day oil accumulations in the Papuan Foreland

    evidence for previously unrecognised petroleum source rocks. Organic

    Geochemistry 36, 2951.

    Volkman, J.K., Banks, M.R., Denwer, K., Aquino Neto, F.R., 1989. Biomarker

    composition and depositional setting of tasmanite oil shale from northern

    Tasmania, Australia [C]. In: 14th International Conference on OrganicGeochemistry, Paris, September 1822, Abstract, p. 168.

    Wandrey,C.J., Law, B.E., Shah, H.A., 2004. Patala-Nammal compositetotalpetroleum

    system, Kohat-Potwar geologic province, Pakistan. US Geological Survey

    Bulletin 2208-B. .

    M. Asif et al. / Organic Geochemistry 42 (2011) 12261240 1239

  • 7/27/2019 Petroleum Goechemistry of Potwar, Asif et al., 2011 OG.pdf

    16/16

    Author's personal copy

    Wenger, L.M., Davis, C.L., Isaksen, G.H., 2001. Multiple controls on petroleum

    biodegradation and impact on oil quality. SPE Reservoir Evaluation Engineering

    5, 375383.

    Winters, J.C., Williams, J.A., 1968. Microbiological alteration of crude oil. I. The

    reservoir. Preprints. ACS Division of Fuel Chemistry, Paper PETR 86; E22E31.

    Zhang, S., Huang, H., 2005. Geochemistry of Palaeozoic marine petroleum from the

    Tarim Basin, NW China: Part 1. Oil family classification. Organic Geochemistry

    36, 12041214.

    1240 M. Asif et al. / Organic Geochemistry 42 (2011) 12261240