sachse et al.2011-petroleum source rocks of the tarfaya basin and adjacent areas, morocco

Upload: bruno-pina

Post on 08-Jul-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    1/19

    Petroleum source rocks of the Tarfaya Basin and adjacent areas, Morocco

    Victoria F. Sachse a, Ralf Littke a,⇑, Sabine Heim a, Oliver Kluth b, Jürgen Schober b, Lahcen Boutib c,Haddou Jabour c, Ferdinand Perssen d, Sven Sindern e

    a Institute of Geology and Geochemistry of Petroleum and Coal, RWTH Aachen University, D-52056 Aachen, Germanyb RWE Dea AG, D-22297 Hamburg, Germanyc National Office of Hydrocarbons and Mining, Rabat 10050, Moroccod GeoForschungs-Zentrum Potsdam, D-14473 Potsdam, Germanye Institute of Mineralogy and Economic Geology, RWTH Aachen University, D-52056 Aachen, Germany

    a r t i c l e i n f o

     Article history:

    Received 25 August 2010

    Received in revised form 9 December 2010

    Accepted 11 December 2010

    Available online 22 December 2010

    a b s t r a c t

    This work characterizes the source rock potential of the Tarfaya Basin and enabled us to reconstruct its

    geochemical history. Outcrop samples covering different stratigraphic intervals, plus the northwestern

    part of the Zag/Tindouf Basin (Bas Draa area), were analyzed for total organic carbon (Corg) and total inor-

    ganic carbon contents and total sulfur content. Rock–Eval analysis and vitrinite reflectance measure-

    ments were performed on 56 samples chosen on the basis Corg   content. A set of 45 samples were

    extracted and non-aromatic hydrocarbons were analyzed by way of gas chromatography-flame ioniza-

    tion detection (GC-FID) and GC-mass spectrometry (GC-MS). Non-isothermal open system pyrolysis at

    different heating rates was applied to obtain kinetic parameters for modelling petroleum generation from

    four different source rocks.

    High quality petroleum source rocks with high Corg content and hydrogen index (HI) values were found

    for samples of Eocene, Coniacian, Turonian and Cenomanian age. Most samples were carbonate rich and

    organic/sulfur values were high to moderate. Various maturity parameters indicated immature or possi-

    bly early mature organic matter. Based on organic geochemical and petrological data, the organic matter

    is of marine/aquatic origin (Cenomanian) or a mixture of aquatic and terrigenous material (Eocene). The

    Early Cretaceous interval did not contain high quality source rocks, but indications of petroleum impreg-

    nation were found.

     2010 Elsevier Ltd. All rights reserved.

    1. Introduction

    For a number of years, exploration interest has focussed on-

    shore in western Morocco and offshore on the Atlantic margin

    ( Jarvis et al., 1999). Several companies drilled wells in the on-

    and offshore parts of the Tarfaya Basin, such as the offshore wells

    MO-1 to MO-8, Cap Juby-1, Puerto Cansado-1 and El Hamra-1

    (Leine, 1986). Oil shows in Mauritania and in the northern Moroccan

    basins emphasize the possible economic relevance of activepetroleum systems along the north-west African continental

    margin. Numerous studies focussed on Cenomanian/Turonian

    sediments and their depositional environment (e.g.  Lüning et al.,

    2004; Kolonic et al., 2002; Kuhnt et al., 2001, 2005).

    The main objective of this study was to determine source rock

    parameters for different sedimentary sequences in the Tarfaya

    Basin and adjacent areas, with a focus on Upper Cretaceous and

    Eocene units. We aimed to assess the depositional environment of 

    these source rocks and provide implications for future petroleum

    exploration and resource assessment in Moroccan coastal regions.

    A detailed overview of organic matter (OM) type, quantity, and

    quality is presented to improve the overall organic geochemical

    information available for the Tarfaya Basin. For this purpose, out-

    crop samples covering various stratigraphic units of the basin were

    collected. Because of the absence of Jurassic outcrops within the

    basin, samples of this age were sampled north of Agadir, the out-

    crop location nearest to the basin. In addition, we analyzed some

    Palaeozoic sediments, which occur in the deep subsurface. BecausePaleozoic sequences are not exposed or are not present in the

    Tarfaya Basin, a few sediments of this age were sampled along

    the northern flank of the adjacent Tindouf Basin.

    2. Tarfaya Basin – geological background

    Tarfaya Basin is an Atlantic margin basin in the southernmost

    part of Morocco, stretching over 1000 km along the western mar-

    gin of the Sahara (Fig. 1). It is bounded by the Mauritanide thrust

    belt (Adrar Souttouf, Dhoul, Zemmour) and the Precambrian

    Reguibat Arch to the southeast and by the Paleozoic outcrops of 

    the Anti-Atlas and the Tindouf Basin to the northeast (Fig. 1). The

    0146-6380/$ - see front matter   2010 Elsevier Ltd. All rights reserved.doi:10.1016/j.orggeochem.2010.12.004

    ⇑ Corresponding author.

    E-mail address: [email protected] (R. Littke).

    Organic Geochemistry 42 (2011) 209–227

    Contents lists available at   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

    http://dx.doi.org/10.1016/j.orggeochem.2010.12.004mailto:[email protected]://dx.doi.org/10.1016/j.orggeochem.2010.12.004http://www.sciencedirect.com/science/journal/01466380http://www.elsevier.com/locate/orggeochemhttp://www.elsevier.com/locate/orggeochemhttp://www.sciencedirect.com/science/journal/01466380http://dx.doi.org/10.1016/j.orggeochem.2010.12.004mailto:[email protected]://dx.doi.org/10.1016/j.orggeochem.2010.12.004

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    2/19

    basement consists of the Precambrian North-West African Craton

    and Paleozoic rocks discordantly overlain by Mesozoic sedimentsthat reach a local maximum thickness of 12 km (Kolonic et al.,

    2002). The geological and stratigraphic structure of the basin have

    been investigated in detail using well and seismic data. The more

    recent onshore part of the basin is characterized by syn-rift

    structures in terms of grabens and half grabens overlain by

    almost undeformed Jurassic to Cretaceous formations that form a

    westward dipping monocline.

    As a result of early rifting of the central and southern North

    Atlantic during the Permian and a marine incursion into this rift

    system during the Triassic, a major phase of evaporite deposition

    occurred in the region (Heyman, 1989). The resulting salt province

    extends along the northwestern coast of Morocco and salt diapirs

    are important structural elements in the offshore area. Major

    transgression occurred in the Jurassic, coupled with high subsi-dence rates that were compensated by thick terrigenous clastic se-

    quences as well as carbonate platform buildups in the shallower

    water shelf area, with a transition into a thin, basinal sequence.

    The 700 m of Jurassic sands, silts and shales are preserved on

    Fuerteventura, indicating transport over the shelf into the deep

    water environment ( Jarvis et al., 1999). The end of the Jurassic is

    marked by a sea level fall and sub-aerial exposure of the shelf 

     Jurassic sequence, as well as karstification and presence of red

    beds, which were drilled in wells Cap Juby-1, MO-2 and MO-8.

    Early Cretaceous (Aptian–Albian) sedimentation was mainly

    deltaic, most likely with sediment input from the Tindouf Basin

    in the northeast and the African Craton in the southeast. Major

    transgressions then occurred during the Cretaceous, mainly Mid

    to Late Cretaceous (Albian/Cenomanian, Cenomanian/Turonian,Santonian/Campanian), which initiated deposition of >800 m of 

    laminated biogenic sediments of Cenomanian to Santonian age in

    the Tarfaya Basin (Kolonic et al., 2002; Kuhnt et al., 2001, 2005).The sediments are rich in calcareous nanoplankton, dispersed

    biogenic silica, planktonic foraminifera, and organic matter (OM).

    Sedimentation rates exceeded 0.1 cm/ky for deep areas (DSDP

    367; Nzoussi-Mbassani et al., 2003) to ca. 1.1–1.7 cm/ky at Tarfaya

    and around 10–22 cm/ky for continental shelves (Kolonic et al.,

    2002; Nzoussi-Mbassani et al., 2003). The depositional environ-

    ment was nutrient-rich as a result of an open shelf upwelling sys-

    tem. During the Cenomanian/Turonian anoxic event, organic

    carbon rich black shales were deposited (Lüning et al., 2004;

    Kolonic et al., 2002; Kuhnt et al., 2001, 2005).

    An erosional unconformity truncates all of the Paleocene, Upper

    Cretaceous and part of the Lower Cretaceous at the shelf edge. The

    erosion probably took place in Santonian to Paleocene times.

    Eocene and Oligocene units are overlain by a thicker Miocenesequence (max. thickness 1 km;  Davison, 2005). More detailed

    geological descriptions are given by e.g.   Choubert et al. (1966,

    1972), Wiedmann et al. (1978), von Rad and Arthur (1979), von

    Rad and Einsele (1980), Ranke et al. (1982), Heyman (1989),

    Davison (2005) and Michard et al. (2008).

    3. Methods

    Samples were collected from outcrops of Albian, Cenomanian,

    Turonian, Coniacian, Santonian, Campanian and Eocene formations

    within the Tarfaya Basin, as well as Palaeozoic units (Ordovician,

    Devonian, and Carboniferous) in the northern flank of the Tindouf 

    Basin.

    Total inorganic carbon (Cinorg) and organic carbon (Corg) weremeasured with a LECO RC-412 carbon analyzer via IR absorption

    Fig. 1.   Location of study area.

    210   V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227 

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    3/19

    in a two stage measurement process (Corg  between 350 C and

    520 C; Cinorg between 520 C and 1050 C). With this method Corgand Cinorg   could be determined in a single run without prior

    removal of carbonate via acid treatment. Total carbon (Ctotal) con-

    centration was determined using Ctotal = Cinorg + Corg. The CaCO3proportion (%) was calculated using CaCO3 = Cinorg 8.333. This

    leads to a slight overestimation of carbonate content if dolomite

    is present instead of calcite. Microscopically, however, no dolomiterhomboeders were detected. Additionally, total sulfur concentra-

    tion (TS) was measured using a Leco S 200 sulfur analyser (preci-

    sion is

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    4/19

    Inc.). The generated bulk petroleum formation curves served as in-

    put for the bulk kinetic model consisting of an activation energy

    distribution and a single pre-exponential factor. The pyrolysis

    products were transferred via a constant He flow (50 ml/min)

    and detected via FID. Low heating rates were used to avoid heat

    transfer problems that might influence the product evolution

    curves and, consequently, the geological predictions (Schenk and

    Dieckman, 2004). The discrete activation energy distribution witha single frequency factor was determined using the Kinetics 2000

    software (Burnham et al., 1987).

    4. Results

    4.1. Elemental analysis

    Highest Corg  values were found for the Eocene, Coniacian and

    Cenomanian/Turonian samples, with values up to 17% (Appendix).

    For the Eocene and Cenomanian/Turonian samples, there was a

    tendency for the highest Corg values to occur in outcrops close tothe coast [Outcrop 1 (Eocene) and Outcrop 2 for Cenomanian sam-

    ples]. The Cinorg  content and thus the CaCO3 content was generally

     Table 1

    Overview of Rock–Eval pyrolysis data of selected samples from the Tarfaya Basin.

    Sample Age/outcrop Corg (%)   S 1  (mg/g rock)   S 2 (mg/g rock)   S 3  (mg/g rock)   T max (C) HI (mg/g Corg) OI (mg CO2/g Corg) PI (S 1/S 1 + S 2)

    681 Eocene 0.6 0.08 0.53 1.86 412 87 311 0.13

    682 Eocene 1.0 0.14 1.65 3.86 409 164 383 0.08

    683 Eocene 3.29 1.2 11.52 12.68 423 350 385 0.09

    684 Eocene 1.96 0.81 5.17 5.57 418 263 284 0.14

    685 Eocene 2.29 0.95 10.64 21.58 417 465 944 0.08

    686 Eocene (Outcrop 1) 7.2 2.34 29.15 7.38 425 405 103 0.07

    687 Eocene 2.75 1.69 9.35 24.49 408 340 891 0.15

    688 Eocene 1.68 0.65 8.31 2.32 427 494 138 0.07

    690 Eocene 0.5 0.04 0.06 3.86 – 12 766 0.4

    705 Eocene 3.54 0.48 13.51 2.91 423 382 82 0.03

    706 Eocene 3.34 0.51 16.42 3.6 420 492 108 0.03

    707 Eocene 4.47 2.45 31.04 1.92 411 694 43 0.07

    674 Campanian 0.8 0.09 0.47 2.28 429 59 285 0.16

    675 Campanian 1.66 0.22 5.56 6.99 422 335 421 0.04

    676 Campanian (Outcrop 4) 2.05 0.39 7.82 9.76 428 382 477 0.05

    692 Campanian 0.37 0.02 – 0.2 – – 54 –

    723 Santonian 0.36 0.09 0.01 0.89 – 3 246 0.9

    726 Santonian 0.51 0.08 0.1 1.46 – 19 286 0.44

    728 Santonain 0.31 0.02 0.04 0.96 – 13 311 0.33

    693 Coniacian 6.1 1.78 37.24 5.24 408 610 86 0.05

    694 Coniacian 2.31 1.15 12.08 10.64 414 522 460 0.09

    695 Coniacian 5.21 1.72 38.49 21.84 409 738 419 0.04

    696 Coniacian (Outcrop 3) 6.99 2.04 43.98 3.07 417 629 44 0.04

    698 Coniacian 1.09 0.14 4.95 1.61 413 454 148 0.03

    699 Coniacian 7.02 1.72 44.07 2.93 413 628 42 0.04700 Coniacian 4.33 0.85 24.03 3.77 412 554 87 0.03

    701 Coniacian 5.63 2.11 35.1 5.15 409 624 92 0.06

    702 Coniacian 4.58 1.57 24.01 25.13 410 524 549 0.06

    703 Coniacian 0.61 0.06 0.05 1.97 – 8 323 0.55

    704 Coniacian 0.75 0.07 0.12 – – 16 – 0.37

    705 Coniacian 1.56 0.8 6.63 1.05 417 426 67 0.11

    770 Turonian 8.54 2.54 50.68 2.68 413 594 31 0.05

    771 Turonian 5.86 1.73 31.88 3.27 410 544 56 0.05

    772 Turonian 7.79 2.2 49.01 4.6 417 629 59 0.04

    773 Turonian 4.8 1.2 25.49 5.67 417 530 118 0.04

    774 Turonian 5.06 1.41 30.57 2.72 415 604 54 0.04

    775 Turonian 7.35 1.84 40.28 5.82 410 548 79 0.04

    776 Turonian 3.9 0.97 18.81 8.2 419 482 210 0.05

    668 Cenomanian 14.48 4.12 98.15 2.53 412 678 17 0.04

    669 Cenomanian 11.32 2.99 78.23 4.84 418 691 43 0.04

    670 Cenomanian 15.57 3.06 104.92 4.43 417 674 28 0.03

    671 Cenomanian (Outcrop 2) 16.79 4.46 122.6 4.9 422 730 29 0.04672 Cenomanian 10.05 2.12 71.79 5.62 412 714 33 0.03

    673 Cenomanian 8.29 2.57 64.47 5.62 411 778 68 0.04

    711 Cenomanian 0.34 0.06 0.01 1.51 – 3 447 0.86

    722 Cenomanian 0.32 0.02 – 2.17 – – 682 –

    733 Cenomanian 0.41 0.11 0.17 0.71 – 42 173 0.39

    779 Cenomanian 0.38 0.07 0.04 1.4 – 11 370 0.64

    794 Albian 0.38 0.16 0.39 0.51 415 103 135 0.29

    750 Albian 0.5 0.04 0.06 – – 12 – 0.4

    751 Albian 0.31 0.03 0.15 0.27 – 49 88 0.17

    752 Albian 0.38 0.09 1.27 0.72 – 335 190 0.07

    753 Albian 0.51 0.09 0.39 1.48 419 76 289 0.19

    754 Albian 0.57 0.06 0.4 0.94 – 70 166 0.13

    755 Albian 0.65 0.04 0.09 0.84 – 14 130 0.31

    759 Albian 0.41 0.06 0.04 2.36 – 10 570 0.6

    718 Lower Cretaceous 0.32 0.03 – 0.88 – – 272 –

    719 Lower Cretaceous 0.41 0.64 0.62 1.8 – 150 435 0.51

    658 Carboniferous 0.5 0.03 0.08 0.28 – 16 56 0.27

    649 Upper Devonian 0.38 0.16 0.39 0.51 415 103 135 0.29

    212   V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227 

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    5/19

    high for the Cenomanian/Turonian samples and variable or low for

    the Eocene and Aptian samples, respectively.

    Coniacian samples were selected from only one outcrop area

    (Outcrop 3) and had Corg  up to 7% and CaCO3  content >50%. Corgcontents of

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    6/19

    Cinorg/Corg patterns defined clear trend lines for the Carbonifer-

    ous, Devonian, and Ordovician samples: with Corg increasing with

    increasing, CaCO3   content. A different pattern was obvious for

    the Lower Cretaceous and Jurassic samples: despite low Corg values,

    the CaCO3  values were high for some samples.

    Sulfur content was highly variable. Highest values – up to 2.8%

    – occurred in the Cenomanian/Turonian. Nevertheless, because of 

    high Corg values, the respective TS/Corg values were only moderateto low, ranging between 0.1 and 0.24. Moderate to high sulfur con-

    tents also occur in the Eocene and the Coniacian (up to 2%). High

    TS/Corg values seemed to be characteristic of the Eocene, with val-

    ues up to 0.6 (Fig. 2a).

    In the Santonian and Campanian samples, the sulfur content

    was quite low, with values

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    7/19

    land-derived eroded material in the Tarfaya Basin during Turonian

    and Cenomanian times. A higher amount of vitrinite/vitrinite-like

    particles could be observed in samples of Eocene and Campanian

    samples. The Santonian samples revealed a high amount of 

    opaque black particles, which represent strongly oxidized and

    re-sedimented terrestrial OM. Coniacian samples revealed a highamount of liptinite, but also vitrinite/vitrinite-like particles. The

    Albian samples contained predominantly vitrinite.

    Characteristic microscopic observations for selected samples

    are shown in Fig. 5 and listed in Table 2.

    4.4. Molecular composition

    Biomarker ratios were used to characterize the depositional

    environment and the maturation range of potential source rocks.

    The biomarkers evaluated were tri- and pentacyclic triterpanes,

    as well as steranes (cf.  Peters and Moldowan, 1991), with a focus

    on evaluation of peak area ratios from tri- and pentacyclic terpanes

    measured from the m/ z  191 trace and steranes measured from the

    m/ z  217 trace (Table 3).The n-and iso-alkane patterns (Fig. 6), except for the Albian and

    some of the Eocene samples, showed a clear dominance of short

    chain (C15–C19) vs. long chain  n-alkanes (C27–C31). In the Eocene

    samples (e.g., 695), long and short chain   n-alkanes occurred in

    similar concentration. In contrast, the Albian sediments showed a

    different pattern, characterized by a clear predominance of long

    chain n-alkanes.

    The abundance of pristane (Pr) and phytane (Ph) relative to the

    n-alkanes was moderate to high, especially in Coniacian, Turonian

    and Cenomanian samples. Samples of the Eocene, Santonian and

    Albian samples showed a dominance of C17 over Pr and of C18 over

    Ph (Fig. 7). In particular, the Cenomanian samples had a dominance

    of Pr vs. n-C17. Pr/Ph for the Eocene samples revealed values up to

    1.4; similar values were also given by Albian samples with valuesup to 1.2. Lowest values were given by Campanian samples, which

    were 1 for nearly all samples(Table 4). Eocene, Campanian, Coniacian, and Santonian samples

    had a wide range of values. A narrow range was given for the

    Turonian (1.03–1.44) and Cenomanian (1.03–1.16) samples,

    whereas highest values were for the Albian (1.85–2.85). The

    OEP showed a similar pattern to the CPI: Eocene and Campanian

    samples showed a wide variation. Samples of the Coniacian

    (0.58–6.30), Turonian (1.94–4.7) and Cenomanian (1.13–2.01), as

    well as the Albian ones (2.17–3.13) showed a narrower range

    for the OEP values.

    Tricyclic terpanes in the range C20–C28 were nearly absent from

    all samples. Pentacyclic terpanes of the hopane series from C27 to

    C35 were dominated by C30  hopane as well as C31 to C33  hopanes.

    The 17a(H)21b(H) and 17b(H)21b(H) isomers dominated the  m/ z 

    191 fragmentograms. In particular, C31–C33 hopanes were presentin the Coniacian, Cenomanian and Turonian samples, but were

    not present or only at low concentration in the Eocene and Albian

    samples. The norhopane (C29)/hopane (C30) ratio was

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    8/19

    5. Discussion

    5.1. General sediment composition and depositional environment 

    TS was measured to provide insight into the depositional envi-

    ronment and in particular the intensity of bacterial sulfate reduc-

    tion (Berner, 1970, 1984). Under anoxic conditions dissolved

    sulfate is reduced to H2S, which reacts with iron minerals to form

    iron sulfides. TS vs. Corg ratio reflects the intensity of microbial sul-

    fate reduction in OM decomposition, giving a qualitative indicationof the redox status of the environment of deposition. Berner (1970,

    1984) found an empirical relationship between S content and C orgcontent, which is typical of most marine sediments deposited un-

    der aerobic bottom waters (Fig. 2a).

    Moderate or high TS content and TS/Corg   values (Fig. 2a) are

    characteristic of nearly all of our Eocene and Upper Cretaceous

    samples. These moderate to high ratio values indicate, in general,

    strong bacterial sulfate reduction. Very high TS/Corg values indicate

    that more OM is consumed via sulfate reduction than under nor-

    mal marine conditions (Berner, 1984). Visual analysis indicates sig-

    nificant amounts of pyrite in the Eocene, Coniacian, Turonian and

    Cenomanian samples, derived from bacterial sulfate reduction

    and OM oxidation. This indicates that sufficient iron was available

    to fix most of the sulfur in the form of iron sulfide.   Littke et al.(1991a) and Lückge et al. (1996) showed that the consumption of 

    part of the metabolized OM during early diagenesis greatly influ-

    ences the OM quality in shallow marine sediments. The bulk of 

    the Cretaceous and Eocene samples shows high Corg  and TS con-

    tents, but only moderate TS/Corg  values. This might indicate that

    these samples were deposited under high productivity conditions,

    with bottom waters that were not anoxic, but only partly reduced

    in molecular oxygen. Thus, sulfur uptake into the sediments oc-

    curred only in the freshly deposited sediment, probably at a fewdecimetres below the sediment/sea water interface. In completely

    anoxic waters, sulfide precipitation and sulfur uptake into OM

    already starts within the water column, leading to high TS/Corgvalues (Sinninghe Damsté and Köster, 1998).

    Interestingly, very high values also are typical of the Albian, but

    at low Corg and low HI values. The latter indicate only poor preser-

    vation of the primary OM, i.e. no anoxic bottom water. Another

    explanation for the high TS/Corg  values in these sediments would

    be petroleum impregnation followed by (bacterial) sulfate reduc-

    tion and petroleum oxidation. However, allochtonous solid bitu-

    men was not observed in the samples in great quantity. Further

    studies would be necessary to test this hypothesis of petroleum

    impregnation, which is supported by high PI values for the Albian.

    Further insight is provided by sulfur/iron values (Fig. 2b). Most

    Eocene and Campanian samples plot close to the pyrite line, which

    implies that essentially all iron is fixed in pyrite. Nevertheless, for

    these samples, iron availability was not limiting with respect to

    pyrite formation. Some Eocene and Campanian samples even have

    excess iron, which could be fixed in minerals such as smectite,

    chlorite and other clay minerals, or magnetite. In contrast, the

    Cenomanian and Turonian samples clearly have excess sulfur

    (Fig. 2b). This observation implies that not all the sulfur is fixed

    in pyrite, as a result of limited availability of reactive iron. Most

    of the excess sulfur was probably incorporated into OM.

    V and Ni are common in marine carbonate source rocks

    (Barwise, 1990). In our samples, the highest V/(V + Ni) values occured

    in the Eocene, Cenomanian and Turonian samples, up to 0.89

    (Fig. 3a, b), indicating a more marine than terrestrial milieu

    (Barwise, 1990). An increase in V/(V + Ni) is often interpreted asan indication of anoxic bottom water conditions (Barwise, 1990;

    Sundararaman et al., 1991). Most of our Eocene samples have

    Ni/V values

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    9/19

    Turonian define trends, showing that, with an increase in silicate,

    the content of OM also increases. This tendency is explained as a

    consequence of a higher nutrient supply when the supply of clastic

    material increased in otherwise carbonate-dominated environ-

    ments (Stein and Littke, 1990; Thurow et al., 1992).

    The Albian samples represent another system, with a high

    proportion of silicate and a low amount of OM. Here, silicate-rich,

    carbonate-lean sediments may represent times in which

    bioproduction was significantly reduced; therefore, both OM and

    carbonate were depleted. In the case of the Eocene, there is no clear

    trend because of the occurrence of several siliciclastic and several

    carbonate dominated samples. In these sediments, OM productiv-

    ity increased with increasing nutrient supply, usually in combina-tion with a higher input of siliciclastics (Fig. 9). In contrast, most of 

    the siliciclastic Eocene sediments showed an increase in OM con-

    tent with increasing carbonate content, probably reflecting the in-

    put of algal/phytoplankton biomass, which was almost absent from

    the pure siliciclastic sediments. Deposition may have occurred

    when bioproduction was much reduced, so carbonate and OM

    were depleted. Oxidation within these sediments further contrib-

    uted to organic carbon loss.

    Short chain  n-alkanes (

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    10/19

    sedimentary OM. Most of the samples had Pr/Ph values 1 for

    Eocene samples indicate a higher oxygen content in bottom

    waters during deposition. The same holds true for Albian samples,

    which showed a wider variation in Pr/Ph. Coniacian, Turonian and

    Cenomanian samples exhibited a predominance of Ph. Santonian

    samples showed a wider spread of values, possibly indicating a

    variation in the depositional environment (Table 4).

    Eocene OM in our samples comprises a mixture of marine and

    terrestrially derived material. The dominance of the short chain

    n-alkanes supports the assumption of mainly marine, aquatic OM

    in the case of the Turonian and Cenomanian samples. In contrast,

    the n-alkane pattern of Albian sediments indicates a predominance

    of terrestrial OM, with little input of algae/phytoplankton material

    (Fig. 6).

    Trycyclic terpanes in the range C20–C28 are thought to represent

    evidence for a contribution from higher plant material (Tissot and

    Welte, 1984) but were only minor compounds close to the detec-

    tion limit. The occurrence of hopanes reflects the contribution of 

    prokaryotic membranes in bacteria and blue-green algae (Tissot

    and Welte, 1984; Moldowan et al., 1985). The high contribution

    of C31–C33   hopanes in Coniacian, Turonian and Cenomanian

    (Fig. 8) samples is an indicator for deposition under marine (shelf)

    conditions because the C31–C33   homohopanes are derived from

    bacteriohopanetetrol. Homohopanes were only detected in small

    amounts in Eocene and Albian samples. Low homohopane index

    values suggest that anoxic conditions did not occur during deposi-

    tion, but suboxic to dysoxic bottom water may have existed (Peters

    and Moldowan, 1991; Tyson and Pearson, 1991).High concentrations of C29 norhopane, as a result of preferential

    preservation in the presence of sulfur, are frequently observed in

    sediments with limited iron availability, such as carbonates (Blanc

    and Connan, 1992). Neohop-13(18)-ene was detected in all sam-

    ples, except from the Eocene (Fig. 8; Table 3), which is typical of 

    OM rich sediments (Greiner et al., 1977). It is assumed that such

    components are derived from bacteria dwelling at or below the

    chemocline, and that they may be used as indicators of water

    bodies which were stratified in the past. Gammacerane, indicative

    of a stratified water column, often accompanying hypersalinity,

    was not detected.

    The dominance of C29 steranes, especially in Eocene sample 705,

    leads to the conclusion of a mainly terrestrial OM contribution.

    Samples from the Cenomanian and Turonian mainly showed an in-

    put of the C28  steranes associated with marine OM (Fig. 11). C27steranes were dominant in samples from the Coniacian, Turonian,

    and Cenomanian, but not from the Eocene: This biomarker is most

    probably derived from various types of marine algae.

    5.2. Petroleum potential and maturity

    The generative potential of sedimentary rocks is defined by Corgcontent and hydrogen richness, which is a function of kerogen type

    and preservation. Sedimentary rocks with Rock–Eval  S 2  values >5

    and 10 mg HC/g rock are considered as having a good and very

    good source potential, respectively (Peters, 1986). Given the rela-

    tively low maturity of the Morrocan outcrop samples, the mea-

    sured Corg   and pyrolysate yields provide a close estimate of original generative potential.

    The Eocene samples exhibited a wide variation in HI and OI val-

    ues, with the lowest HI values correlating with samples lower in

    Corg. The data fluctuation may be caused by the presence of re-

    sedimented organic particles (in samples with S 2 < 5 mgHC/g rock)

    and/or the moderate quality of amorphous OM even within the

    Fig. 7.   Pr/nC17 vs. Ph/nC18 for selected samples from various stratigraphic units in

    the Tarfaya Basin (after Shanmungam, 1985).

     Table 4

    Hopane and sterane biomarker parameters.

    Sample Age Hopanes Steranes   n-and iso-

    alkanes

    22S/

    (22S + 22R)

    Ts/

    (Ts + Tm)

    Ts/

    C3017aHopaneC29Hopane/

    (C29Hopane + C29Ts)

    17b,21aHopane/17a,21bHopane

    Steranes/

    17aHopanesC29aaa(S)/C29aaa(S + R)

    20S/

    (20S+ 20R)

    Pr/

    Ph

    CPI OEP

    683 Eocene – 0.41 – – – – – – 1.17 1.97 0.21688 Eocene – 0.16 – – – – – – – – –

    690 Eocene – 0.11 – – – – –– – 0.78 1.18 1.83

    705 Eocene 0.45 0.48 0.33 – 0.40 3.01 0.20 0.20 0.77 1.07 1.73

    707 Eocene 0.55 0.44 – – – 3.86 0.23 0.24 1.33 1.19 2.09

    728 Santonian – 0.74 – – – – – – 0.05 0.92 1.13

    693 Coniacian 0.64 0.52 0.17 0.37 0.41 3.09 0.29 0.29 0.50 1.00 2.14

    695 Coniacian 0.49 0.71 0.1 0.27 0.48 2.86 0.33 0.33 0.52 1.15 –

    696 Coniacian 0.38 0.56 – – – 9.00 0.34 0.33 0.97 0.64 0.58

    698 Coniacian – 0.86 – – – – 0.34 0.34 0.50 1.11 1.54

    700 Coniacian 0.48 0.47 0.16 0.65 0.57 1.39 0.32 0.31 0.40 1.07 2.38

    770 Turonian – 0.47 0.24 – 0.52 16.76 0.38 0.37 0.46 1.03 1.94

    773 Turonian – 0.56 0.18 – 0.44 38.32 0.46 0.46 0.46 1.02 1.87

    774 Turonian 0.51 0.53 0.17 – 0.50 1.49 0.37 0.37 0.39 1.11 4.10

    775 Turonian 0.45 0.45 0.16 – 0.46 1.9 0.40 0.39 0.42 1.44 4.70

    668 Cenomanian 0.47 0.78 0.21 0.45 0.49 3.92 0.41 0.41 0.44 – –

    669 Cenomanian 0.57 0.58 0.14 0.41 0.54 5.07 0.47 0.47 0.58 – –

    671 Cenomanian 0.55 0.40 0.05 0.37 0.53 3.19 0.50 0.50 0.50 1.03 1.45

    755 Albian – 0.39 0.30 0.30 – – – – 1.22 2.17 3.13

    218   V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227 

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    11/19

    intervals with moderate Corg   values. Similarly, higher HI values

    (>300 mg HC/g rock) for the majority of the samples may reflect

    better preservation of marine derived amorphous OM or a smaller

    supply of terrestrial or inert OM (Farrimond et al., 1990). The ker-

    ogen in Eocene samples is mainly Type II. Incident light microscopy

    results indicate a high abundance of algal material (Fig. 5) as well

    as some inertinite and re-sedimented vitrinite, consistent with a

    Type II classification. The Rock–Eval maturity indicators,  T max  and

    PI, showed an overall immature or early mature stage for the Eo-cene sediments on the coastal area of the Tarfaya Basin.

    The Campanian samples were characterized by Type III kerogen,

    based on the HI and OI values (Fig. 4a). This is supported by micro-

    scopic observations, which revealed a high proportion of vitrinite

    or vitrinite-like particles, T max and PI identifying this unit as imma-

    ture with respect to oil generation.

    According to the method of  Langford and Blanc-Valleron (1990)

    using a plotof Corg vs. S 2 (Fig. 4b), the OM of the Santonian samples

    was allocated to Type IV kerogen, which implies no potential for oil

    generation. The  T max  and PI values identify this unit as immaturewith respect to oil generation.

    Fig. 8.   Hopane (A–E) andsterane (F–J) distributions of selected samples from theTarfaya Basin (A andF, Eocene; B andG, Coniacian; C andH, Turonian; D andI, Cenomanian;

    E and J, Albian). For peak identifications see Table 3.

    V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227    219

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    12/19

    HI values for the Coniacian samples were moderate to high,

    accompanied by low OI values and high Corg  contents. Based on

    these data and the petrological observations, the OM was allocated

    to oil-prone Type II kerogen, with some outlier samples that

    plotted between the trend lines for oil prone Type I and Type II

    kerogens (Fig. 4a).   T max   values and vitrinite reflectance values

    reveal an immature source rock.

    Type I/II kerogen with a high potential for oil generation is pres-

    ent in Cenomanian/Turonian samples. Plots of HI vs. OI and HI vs.

    T max supported the latter observation (Fig. 4a and b; Table 1). Per-

    sistently low T max values and a vitrinite reflectance of ca. 0.3, con-

    firmed the immature stage (Table 2). Fluorescence microscopy

    supports these data, because strong fluorescence was observed

    (see Leine, 1986).The Albian samples contain Type III kerogen, having only a low

    potential for oil generation, with low HI and  S 2 values (Fig. 4b and

    c) supporting the latter observation. The PI values for the Albian

    samples were up to 0.6, indicating active petroleum migration

    and sediment impregnation.

    All the Lower Cretaceous, Jurassic, Carboniferous, and Devonian

    samples had very low potential for oil or gas generation, based on

    OM quantity and quality (Table 1;  Appendix).

    CPI and the OEP values support the above conclusion about

    immature to early mature OM, except for some of the Eocene sam-

    ples (Table 4). This is consistent with the Pr/n-C17 and Ph/n-C18 val-

    ues, which are generally dominated by Pr and Ph ( Fig. 7). Only in

    the cases of the Albian and Eocene samples did  n-C17   and  n-C18show a dominance, potentially indicating a more mature level.

    Maturity as expressed by hopane isomerisation ratios [22S/

    (22S + 22R)] of nearly all samples showed that the OM is immature

    to marginally mature, as expressed in Cenomanian samples (Fig. 10a,

    Table 4). The Ts/Tm ratio [18a(H)-22,29,30-trisnorneohopane(Ts)/17a(H)-22,29,30-trisnorhopane (Tm), Table 4] is dependent onboth thermal maturation and source lithology. During catagenesis,

    Tm is less stable than Ts, so the ratio of Ts to Tm should increasewith maturity. This is also the case for the C29Ts/(C29   hopane +

    C29Ts) equivalent. In our samples, both ratios show similar values

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    13/19

    Maturity information derived from typical sterane ratios shows

    clear immaturity for all samples, i.e. the 20S/(20S + 20R) values

    are all

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    14/19

    Fig. 13.   Calculated generation rate (normalized in %; black lines;) vs. temperature for a geological heating rate of 3 C/Ma and calculatedvitrinite reflectance (thin black lines)

    of selected samples from the Tarfaya Basin.

    Fig. 14.  Calculated predictions of kerogen transformation ratio (black lines) vs. temperature for a geological heating rate of 3 C/Ma and calculated vitrinite reflectance (thinblack lines) of selected samples from the Tarfaya Basin.

    222   V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227 

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    15/19

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    16/19

     Appendix A  (continued)

    Sample Age Outcrop Lat. Long. Corg (%) Ctotal (%) S (%) CaCO3 (%) Fe (%) V/(Ni+V)

    694 Coniacian 2759.8640N 1207.2300W 2.31 11.30 0.93 74.86

    695 Coniacian 2759.8650N 1207.2030W 5.21 10.81 46.61

    696 Coniacian 3 2759.8640N 1207.1890W 6.99 14.00 1.53 58.39

    697 Coniacian 2759.864‘N 1207.1890W 0.00

    698 Coniacian 2759.865‘N 12

    07.180

    0

    W 1.09 10.96 0.35 82.21699 Coniacian 2759.8620N 1207.1720W 7.02 11.89 1.25 40.59

    700 Coniacian 2759.8590N 1207.1670W 4.33 10.08 47.86

    701 Coniacian 2759.8580N 1207.1650W 5.63 12.24 55.08

    702 Coniacian 2759.8620N 1207.1580W 4.58 12.48 65.80

    703 Coniacian 2759.858‘N 1207.1520W 0.61 10.01 0.12 78.30

    704 Coniacian 2759.8560N 1207.1470W 0.75 6.87 51.05

    679 Coniacian 2650.5830N 1303.9260W 1.56 2.26 2.14 5.84

    770 Turonian 2812.1610N 1146.8920W 8.54 12.14 2.01 30.03 0.89 0.75

    771 Turonian 2812.1600N 1146.8900W 5.86 12.15 1.05 52.43

    772 Turonian 2812.1670N 1146.8930W 7.79 15.10 1.14 60.89 0.28 0.62

    773 Turonian 2812.1700N 1146.8930W 4.81 12.03 60.11

    774 Turonian 2812.1710N 1146.9000W 5.06 12.88 0.85 65.11 0.19 0.54

    775 Turonian 2812.1700N 1146.9020W 7.35 13.02 47.20

    776 Turonian 2812.1720N 1146.9010W 3.90 12.38 0.49 70.63 0.26 0.57

    709 Cenomanian/Turonian 2637.9920N 1158.7210W 0.05 11.27 93.44

    710 Cenomanian/Turonian 2637.9960N 1158.7310W 0.08 12.04 99.62

    711 Cenomanian/Turonian 2638.0010N 1158.7370W 0.34 11.04 89.15 0.27 0.17

    720 Cenomanian/Turonian 2634.6620N 1214.7920W 0.12 11.97 98.69

    721 Cenomanian/Turonian 2634.6810N 1214.7900W 0.07 9.49 78.45

    722 Cenomanian/Turonian 2634.6810N 1214.6970W 0.32 10.28 82.98

    730 Cenomanian/Turonian 2655.1750N 1212.0430W 0.06 11.08 91.84

    731 Cenomanian/Turonian 2655.1780N 1212.0170W 0.18 12.06 98.95

    732 Cenomanian/Turonian 2655.1440N 1212.0110W 0.24 10.85 88.38

    733 Cenomanian/Turonian 2655.1440N 1212.0120W 0.41 2.19 0.04 14.86 0.75 0.26

    734 Cenomanian/Turonian 2655.4930N 1212.5670W 0.26 9.19 74.39

    735 Cenomanian/Turonian 2655.5100N 1212.5830W 0.29 9.93 80.30 0.38 0.26

    736 Cenomanian/Turonian 2655.5470N 1212.5950W 0.27 11.18 90.88 0.14 0.28

    777 Cenomanian/Turonian 2809.7520N 1138.2480W 0.07 0.09 0.14

    778 Cenomanian/Turonian 2809.6180N 1138.2950W 0.12 9.61 0.05 79.13779 Cenomanian/Turonian 2809.5910N 1138.3150W 0.38 0.85 3.93

    654 Cenomanian/Turonian 2802.4290N 916.1760W 0.04 0.05 0.11

    655 Cenomanian/Turonian 2802.4100N 916.4340W 0.03 0.29 2.17

    668 Cenomanian 2800.2750N 1228.6810W 14.48 17.21 22.74

    669 Cenomanian 2800.2750N 1228.6760W 11.32 16.04 1.7 39.32 0.16 0.48

    670 Cenomanian 2800.2740N 1228.6540W 15.57 18.79 2.33 26.82

    671 Cenomanian 2 2800.2790N 1228.6450W 16.79 19.89 2.84 25.82

    672 Cenomanian 2800.2850N 1228.6330W 10.05 17.50 2.16 62.06 0.11 0.53

    673 Cenomanian 2800.2820N 1228.6200W 8.29 15.50 1.34 60.07 0.38 0.72

    749 Albian 2817.1670N 1132.2910W 0.06 0.09 0.18

    750 Albian 2817.4910N 1132.2880W 0.50 3.04 0.1 21.15

    751 Albian 2817.6340N 1132.5920W 0.31 0.47 1.38

    752 Albian 2817.6130N 1132.6330W 0.38 0.49 0.92

    753 Albian 2817.6110N 1132.6250W 0.51 0.76 0.77 2.02754 Albian 2817.6110N 1132.6260W 0.57 0.67 1.34 0.82

    755 Albian 2817.6120N 1132.6230W 0.65 0.87 0.73 1.87

    756 Albian 2817.6090N 1132.6230W 0.28 0.30 0.16

    757 Albian 2817.6070N 1132.6270W 0.33 2.81 0.52 20.63

    758 Albian 2817.6130N 1132.6190W 0.27 3.68 0.08 28.45

    759 Albian 2817.6090N 1132.6170W 0.41 2.53 17.59

    667 Cretaceous 2806.3120N 918.5800W 0.04 0.11 0.64

    708 Cretaceous 2637.7040N 1152.7870W 0.04 0.08 0.40

    714 Lower Cretaceous 2701.8390N 1200.2750W 0.13 11.94 98.41

    715 Lower Cretaceous 2701.8610N 1200.2790W 0.22 10.04 81.83

    716 Lower Cretaceous 2701.8700N 1200.2870W 0.13 11.83 97.47

    717 Lower Cretaceous 2707.2640N 1152.7540W 0.33 9.96 80.25

    718 Lower Cretaceous 2707.2970N 1152.7350W 0.11 0.19 0.67

    224   V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227 

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    17/19

     Appendix A (continued)

    Sample Age Outcrop Lat. Long. Corg (%) Ctotal (%) S (%) CaCO3 (%) Fe (%) V/(Ni+V)

    719 Lower Cretaceous 2707.3130N 1152.7520W 0.41 1.40 8.17

    737 Lower Cretaceous 2707.1310N 1134.5710W 0.02 0.03 0.11

    738 Lower Cretaceous 2707.1150N 1134.5950W 0.03 0.04 0.09

    760 Lower Cretaceous 2816.3020N 1130.2650W 0.26 0.47 1.75

    761 Lower Cretaceous 28

    16.2690

    N 11

    30.2490

    W 0.03 0.04 0.06762 Lower Cretaceous 2816.2710N 1130.2490W 0.05 0.05 0.02

    763 Lower Cretaceous 2816.3190N 1130.2520W 0.38 3.33 24.56

    764 Lower Cretaceous 2840.5690N 1107.4570W 0.26 6.37 0.44 50.85

    765 Lower Cretaceous 2840.6350N 1107.3410W 0.24 4.85 38.44

    766 Lower Cretaceous 2840.6220N 1107.3410W 0.16 3.41 27.10

    767 Lower Cretaceous 2840.6220N 1107.3410W 0.42 5.16 0.11 39.53

    768 Lower Cretaceous 2840.6160N 1107.3410W 0.39 5.66 43.88

    769 Lower Cretaceous 2840.6000N 1107.3490W 0.42 1.86 12.02

    780 Jurassic 3033.7630N 932.670W 0.07 9.31 76.95

    781 Jurassic 3033.9620N 932.2530W 0.06 11.51 0.05 95.37

    782 Jurassic 3035.3250N 931.8750W 0.09 7.70 63.39

    783 Jurassic 3035.6070N 931.6660W 0.13 11.39 93.81

    784 Jurassic 3035.8760N 930.5590W 0.06 7.53 0.10 62.24

    785 Jurassic 3036.0070N 929.5350W 0.06 7.49 61.93

    786 Jurassic 3036.1730N 929.6630W 0.12 4.28 34.69

    787 Jurassic 3037.3530N 930.3200W

    788 Jurassic 3037.7690N 930.6670W 0.04 6.16 0.07 50.91

    789 Jurassic 3038.8860N 931.1760W 0.02 3.74 30.99

    790 Jurassic 3039.4420N 928.9880W 0.08 8.62 71.13

    791 Jurassic 3040.3430N 928.9400W 0.04 11.72 0.16 97.31

    650 Namurian 2821.9250N 921.6310W 0.04 0.12 0.68

    651 Late Namurian 2809.6740N 918.4700W 0.05 0.37 2.71

    652 Late Namurian 2809.6740N 918.4700W 0.03 0.05 0.10

    653 Late Namurian 2809.6740N 918.4700W 0.01 0.04 0.19

    658 Vise 2822.9130N 929.3890W 0.50 0.61 0.94

    644 Upper Carboniferous. Vise 2824.1060N 923.9400W 0.04 7.82 0.08 64.84

    645 Upper Carboniferous. Vise 2824.1060N 923.9400W 0.05 11.84 0.14 98.25

    646 Upper Carboniferous. Vise 2824.1060N 923.9400W 0.03 0.06 0.27

    647 Lower Vise 2825.0450N 924.5980W 0.06 0.61 4.55648 Lower Vise 2831.4440N 924.0440W 0.13 0.17 0.34

    641 Late Famennian/Tournai 2827.5290N 921.6890W 0.07 0.21 1.10

    642 Late Famennian/Tournai 2827.5290N 921.6890W 0.02 0.57 4.59

    643 Late Famennian/Tournai 2827.5290N 921.6890W 0.01 2.00 16.52

    712 Devonian 2649.3140N 1146.9010W 0.04 11.88 98.59

    739 Devonian 2701.4770N 1133.5360W 0.11 0.16 0.37

    649 Upper Devonian 2831.4300N 924.0050W 0.38 0.46 1.55 0.69

    659 Upper Devonian 2837.5090N 924.0200W 0.06 10.42 86.33

    660 Upper Devonian 2837.5090N 924.0200W 0.06 0.12 0.54

    661 Upper Devonian 2652.3900N 952.1390W 0.04 0.11 0.56

    662 Upper Devonian 2652.3900N 952.1390W 0.02 0.07 0.44

    663 Upper Devonian 2652.3840N 952.1480W 0.02 0.05 0.25

    664 Upper Devonian 2652.4730N 952.2490W 0.05 0.11 0.48

    665 Upper Devonian 2652.5060N 952.2340W 0.03 0.14 0.97666 Upper Devonian 2652.5060N 952.2340W 0.05 0.25 1.68

    740 Upper Devonian 2635.0650N 1126.0560W 0.16 0.70 4.47

    713 Middle Devonian 2648.9230N 1148.6780W 0.06 1.63 13.01

    741 Middle Devonian 2634.0920N 1125.4170W 0.05 10.54 87.42

    742 Middle Devonian 2634.1670N 1125.4170W 0.03 11.25 93.43

    743 Upper Ordovician 2634.8880N 1125.6110W 0.03 1.32 10.69

    744 Upper Ordovician 2635.1790N 1126.2800W 0.09 0.92 6.90

    745 Upper Ordovician 2638.2200N 1126.6410W 0.04 1.22 9.79

    746 Upper Ordovician 2640.8430N 1128.6740W 0.16 4.23 33.94

    747 Upper Ordovician 2641.8470N 1132.2640W 0.04 0.06 0.17

    748 Upper Ordovician 2643.3260N 1136.5530W 0.06 0.10 0.35

    636 Upper Ordovician 2843.5980N 927.9040W 0.05 0.07 0.18

    637 Lower Ordovcian 2849.7430N 931.6180W 0.05 0.08 0.21

    (continued on next page)

    V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227    225

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    18/19

     Associate Editor —C.C. Walters

    References

    Amijaya, H., Littke, R., 2006. Properties of thermally metamorphosed coal fromTanjung Enim Area, south Sumatra Basin, Indonesia with special reference tothe coalification path of macerals. International Journal of Coal Geology 66,271–295.

    Barwise, A.J.G., 1990. Role of nickel and vanadium in petroleum classification.Energy and Fuels 4, 647–652.

    Berner, R.A., 1970. Sedimentary pyrite formation. American Journal of Science 268,1–23.

    Berner, R.A., 1984. Sedimentary pyrite formation: an update. Geochimica et

    Cosmochimica Acta 48, 605–615.Blanc, P.L., Connan, J., 1992. Origin and occurrence of 25-norhopanes: a statistical

    study. Organic Geochemistry 18, 813–828.Bray, E.E., Evans, E.D., 1961. Distribution of  n-paraffins as a clue to recognition of 

    source beds. Geochimica et Cosmochimica Acta 22, 2–15.Burnham, A.K., Braun, R.L., Gregg, H.R., Samoun, A.M., 1987. Comparison of methods

    for measuring kerogen pyrolysis rates and fitting kinetic parameters. Energyand Fuels 1, 452–458.

    Choubert, G., Faure Muret, A., Hottinger, L., 1966. Aperçu géologique du basin côtierde Tarfaya. Notes et Mémoire du Service Géologique du Maroc 175, 9–219.

    Choubert, G., Faure-Muret, A., Hottinger, L., 1972. La série stratigraphique deTarfaya Maroc sud occidental, et le problème de la naissance de ĺAtlantique.Notes et Mémoire du Service Géologique du Maroc 31, 29–40.

    Cranwell, P.A., 1977. Organic geochemistry of Cam Loch (Sutherland) sediments.Chemical Geology 20, 205–221.

    Davison, I., 2005. Central Atlantic margin basins of North West Africa: geology andhydrocarbon potential (Morocco to Guinea). Journal of African Earth Sciences43, 254–274.

    Dessort, D., Connan, J., Derenne, S., Largeau, C., 1997. Comparative studies of thekinetic parameters of various algaenans and kerogens via open systempyrolysis. Organic Geochemistry 26, 705–720.

    Didyk, B.M., Simonet, R.T., Brassell, S.C., 1978. Geochemical indicators of paleoenvironmental conditions of sedimentation. Nature 272, 216–222.

    Dow, W.G., 1977. Kerogen studies and geological interpretations. Journal of Geochemical Exploration 7, 77–99.

    Espitalié, J., Deroo, G., Marquis, F., 1985. La pyrolyse Rock–Eval et ses applications.Revue de l‘Institut Français du Petrole 40, 563–579.

    Espitalié, J., Ungerer, P., Irwin, I., Marquis, F., 1988. Primary cracking of kerogens;experimenting and modeling C1, C2-C5, C6-C15 and C15+  classes of hydrocarbonsformed. In: Matavelli, L., Novelli, L. (Eds.), Advances in Organic Geochemistry(1987) Part II: Analytical Geochemistry. Organic Geochemistry, vol. 14,pp. 893–899.

    Farrimond, P., Eglinton, G., Brassell, S.C., 1990. The Cenomanian/Turonian anoxicevent in Europe: an organic geochemical study. Marine and Petroleum Geology7, 75–89.

    Greiner, A.C., Spyckerelle, C., Albrecht, P., Orrison, G., 1977. Aromatic hydrocarbonsfrom geological sources. Part V: mono- and di-aromatic hopane derivates.

     Journal of Chemical Research, 3829–3871.Heyman, M.A.W., 1989. Tectonic and depositional history of the Morocconcontinental margin. In: Tankard, A.J., Blankwill, H.R. (Eds.), ExtensionalTectonics and Stratigraphy of the North Atlantic Margins, vol. 46. TheAmerican Association of Petroleum Geologists Memoir, pp. 23–340.

    Hunt, J.M., 1996. Petroleum Geochemistry and Geology. W.H. Freeman andCompany, New York. 743 pp.

     Jarvis, J., Fish, P., Garwood, T., 1999. Morocco‘s Tarfaya deepwater prospects areencouraging. Oil and Gas Journal 16, 90–94.

    Kolonic, S., Sinninghe Damsté, J.S., Böttcher, M.E., Kuypers, M.M.M., Kuhnt, W.,Beckmann, B., Scheeder, G., Wagner, T., 2002. Geochemical characterization of Cenomanian/Turonian black shales from the Tarfaya Basin (SW Morocco).

     Journal of Petroleum Geology 25, 325–350.Kuhnt, W.,Chellai, E.H., Holbourn,A., Luderer, F., Thurow, J., Wagner, T., El Albani, A.,

    Beckmann, B., Herbin, J.P., Kawamura, H., Kolonic, S., Nederbragt, S., Street, S.,Ravilious, K., 2001. Morocco basins sedimentary record may providecorrelations for Cretaceous paleoceanographic events worldwide. EOSTransactions 82, 361–364.

    Kuhnt, W., Luderer, F., Nederbragt, S., Thurow, J., Wagner, Th., 2005. Orbital-scale

    record of the Late Cenomanian–Turonian oceanic anoxic event (OAE-2) in theTarfaya Basin (Morocco). International Journal of Earth Science 94, 147–159.

    Lallier-Vergès, E., Bertrand, P., Huc, A.Y., Tremblay, P., Bückel, D., 1993. Control of the preservation of organic matter by productivity and sulphate reduction inmarine anoxic sediments: Kimmeridgian shales from Dorset (UK). Marine andPetroleum Geology 10, 600–605.

    Langford, F.F., Blanc-Valleron, M.M., 1990. Interpreting Rock–Eval pyrolysis datausing graphsof pyrolizablehydrocarbons vs. total organic carbon. TheAmericanAssociation of Petroleum Geologists Bulletin 74, 799–804.

    Leine, L., 1986. Geology of the Tarfaya oil shale deposit, Morocco. Geologie enMijnbouw 65, 57–74.

    Littke, R., 1993. Deposition, Diagenesis and Weathering of Organic Matter-RichSediments. Lecture Notes in Earth Sciences. Springer Verlag, Berlin. 216 pp.

    Littke, R., Sachsenhofer, R.F., 1994. Organic petrology of deep sea sediments: acompilation of results from the Ocean Drilling Program and the Deep SeaDrilling Project. Energy and Fuels 8, 1498–1512.

    Littke, R., Baker, D.R., Rullkötter, J., 1991a. Keys to the depositional history of the

    Posidonia Shale (Toarcian) in the Hils Syncline, Northern Germany. In: Pearson,R.V., Tyson, T.H. (Eds.), Modern and Ancient Shelf Anoxia. Geological Society,London, pp. 311–334.

    Littke, R., Klussmann, U., Krooß, B., Leythaeuser, D., 1991b. Quantification of calcite-,pyrite-, and organic matter-loss due to weathering of Toarcian black shales andeffects on kerogen and bitumen characteristics. Geochimica et CosmochimicaActa 55, 3369–3378.

    Lückge, A., Boussafir, M., Lallier-Vergés, E., Littke, R., 1996. Comparative study of organic matter preservation in immature sediments along the continentalmargins of Peru and Oman. Part I: results of petrographical and bulkgeochemical data. Organic Geochemistry 24, 437–451.

    Lüning, S., Kolonic, S., Belhadj, E.M., Belhadj, Z., Cota, L., Baric, G., Wagner, T., 2004.Integrated depositional model for the Cenomanian–Turonian organic-richstrata in North Africa. Earth-Science Reviews 64, 51–117.

    Michard, A., Saddiqi, O., Chalouan, A., Frizon de Lamotte, D. (Eds.), 2008. ContinentalEvolution: TheGeology of Morocco. Structure, Stratigraphy, andTectonics of theAfrica–Atlantic–Mediterranean Triple Junction. Lecture Notes in Earth Sciences.Springer-Verlag, Berlin, Heidelberg. 424 pp.

    Moldowan, J.M., Sundaraman, P., Schoell, M., 1985. Sensitivity of biomarker

    properties of depositional environment and/or source input in the LowerToarcian of SW Germany. Organic Geochemistry 10, 915–926.

    Moretti, I., Deacon, K., 1995. Subsidence, maturation and migration history of theTampen Spur area. Marine and Petroleum Geology 12, 345–375.

    Nzoussi-Mbassani, P., Disnar, J.R., Laggoun-Défarge, F., 2003. Organic mattercharacteristics of Cenomanian–Turonian source rocks: implications forpetroleum and gas exploration onshore Senegal. Marine and PetroleumGeology 20, 411–427.

    Peters, K.E., 1986. Guidelines for evaluating petroleum source rocks usingprogrammed pyrolysis. The American Association of Petroleum GeologistsBulletin 70, 318–329.

    Peters, K.E., Moldowan, J.M., 1991. Effect of source, thermal maturity andbiodegradation on the distribution and isomerisation of homohopanes inpetroleum. Organic Geochemistry 17, 47–61.

    Ranke, U., von Rad, U., Wissmann, G., 1982. Stratigraphy, facies and tectonicdevelopment of the on- and offshore Aaiun-Tarfaya Basin – a review. In: vonRad, U., Hinz, K., Sarntheim, M., Seibold, E. (Eds.), Geology of the NorthwestAfrican Continental Margin. Springer Verlag, Berlin, Heidelberg, pp. 86–105.

    Reynolds, J.G., Burnham, A.K., 1995. Comparison of kinetic analysis of source rocksand kerogen concentrates. Organic Geochemistry 23, 11–19.Scalan, R.S., Smith, J.E., 1970. An improved measure of the odd-to-even

    predominance in the normal alkanes of sediment extracts and petroleum.Geochimica et Cosmochimica Acta 34, 611–620.

    Schenk, H.J., Dieckman, V., 2004. Prediction of petroleumformation: theinfluenceof laboratory heating rates on kinetic parameters and geological extrapolations.Marine and Petroleum Geology 21, 79–95.

    Schenk, H.J., Di Primio, R., Horsfield, B., 1997. The conversion of oil into gas inpetroleum reservoirs. Part I: comparative kinetic investigation of gas generationfrom crude oils of lacustrine, marine and fluviodeltaic origin by programmed-temperature closed-system pyrolysis. Organic Geochemistry 26, 467–481.

    Shanmungam, G., 1985. Significance of coniferous rain forests and related organicmatter in generating commercial quantities of oil, Gippsland Basin, Australia.The American Association of Petroleum Geologists Bulletin 69, 1241–1254.

    Sinninghe Damsté, J.S., Köster, J., 1998. An euxinic southern North Atlantic Oceanduring the Cenomanian/Turonian oceanic anoxic event. Earth Planetary ScienceLetters 158, 165–173.

    Stein, R., Littke, R., 1990. Organic- carbon-rich sediments and paleoenvironment:

    results from Baffin Bay (ODP-Leg 105) and the upwelling area off NorthwestAfrica (ODP-Leg 108). In: Huc, A.Y. (Ed.), Deposition of Organic Facies, vol. 30.

     Appendix A  (continued)

    Sample Age Outcrop Lat. Long. Corg (%) Ctotal (%) S (%) CaCO3 (%) Fe (%) V/(Ni+V)

    638 Lower Ordovcian 2849.7430N 931.6180W 0.05 0.12 0.56

    639 Lower Ordovcian 2849.7430N 931.6180W 0.15 0.25 0.82

    640 Lower Ordovcian 2843.6680N 927.9640W 0.01 0.02 0.04

    656 Lower Ordovcian 2901.7270N 854.0640W 0.11 1.19 9.02

    657 Lower Ordovcian 2901.727

    0

    N 854.064

    0

    W 0.12 0.99 7.25

    226   V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227 

  • 8/19/2019 Sachse Et Al.2011-Petroleum Source Rocks of the Tarfaya Basin and Adjacent Areas, Morocco

    19/19

    The American Association of Petroleum Geologists Studies in Geology, pp. 41–56.

    Sundararaman, P., Schoell, M., Littke, R., Baker, D.R., Leythaeuser, D., Rullkötter, J.,1991. Depositional environment of Toarcian shales from NW-Germanyas monitored by porphyrins and carbon isotopes. In: Manning, D.A.C. (Ed.),Organic Geochemistry. Advances and Applications in Energy and theNatural Environment. Manchester University Press, Manchester, pp. 389–390.

    Taylor, G.H., Teichmüller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998.Organic Petrology. Borntraeger, Stuttgart. 704 pp.

    Thurow, J., Brumsack, H.J., Littke, R., Meyers, P., Rullkötter, J., 1992. TheCenomanian/Turonian boundary event in the Indian Ocean – a key tounderstand the global picture. Geophysical Monograph 70, 253–273.

    Tissot, B., Welte, D.H., 1984. Petroleum Formation and Occurrence. Springer Verlag,New York. 699 pp.

    Tyson, R.V., Pearson, T.H., 1991. Modern and Ancient Continental Shelf Anoxia, vol.58. Geological Society Special Publication. 470 pp.

    von Rad, U., Arthur, M.A., 1979. Geodynamic, sedimentary and volcanic evolution of the Cap Boujdour continental margin (NW Africa). In: Talwani, M. (Ed.), Deep

    Drilling Results in the Atlantic Ocean: Continental Margin andPaleoenvironment, vol. 3. American Geophysical Union, pp. 187–203.

    von Rad, U., Einsele, G., 1980. Mesozoic–Cenozoic subsidence history andpaleobathymethry of the northwest African continental margin (Aaiun Basinto DSDP 397). The evolution of passive continental margins. PhilosophicalTransactions of the Royal Society Series A 294, 37–50.

    Waples, D.W., Kamata, H., Suizu, M., 1992a. The art of maturity modeling. Part 1:finding a satisfactory geological model. The American Association of PetroleumGeologists Bulletin 76, 31–46.

    Waples, D.W., Kamata, H., Suizu, M., 1992b. The art of maturity modeling. Part 2:

    alternative models and sensitivity alalyses. The American Association of Petroleum Geologists Bulletin 76, 47–66.

    Wei, H., Bjorøy, M., Roaldset, E., 1994. Multicomponent generation kinetics of source rocks and coals from Norwegian North Sea. In: 6th Conf. of Europ.Association of Petroleum Geosciences and Engineers, Extended Abstract P532.

    Wiedmann, J., Butt, A., Einsele, G., 1978. Vergleich von marokkanischen Kreide-Küstenaufschlüssen und Tiefseebohrungen (DSDP): Stratigraphie,Paläoenvironment und Subsidenz an einem passiven Kontinentalrand.International Journal of Earth Sciences 67, 454–508.

    V.F. Sachse et al. / Organic Geochemistry 42 (2011) 209–227    227