01.03.2013 Views

Source rock quality and hydrocarbon migration pathways within the ...

Source rock quality and hydrocarbon migration pathways within the ...

Source rock quality and hydrocarbon migration pathways within the ...

SHOW MORE
SHOW LESS

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

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

<strong>Source</strong> <strong>rock</strong> <strong>quality</strong> <strong>and</strong><br />

<strong>hydrocarbon</strong> <strong>migration</strong><br />

<strong>pathways</strong> <strong>within</strong> <strong>the</strong> greater<br />

Utsira High area, Viking<br />

Graben, Norwegian North Sea<br />

Gary H. Isaksen <strong>and</strong> K. Haakan I. Ledje<br />

ABSTRACT<br />

The greater Utsira High area is located <strong>within</strong> <strong>the</strong> sou<strong>the</strong>rn part of<br />

quadrants 24 <strong>and</strong> 25 <strong>and</strong> <strong>the</strong> nor<strong>the</strong>rn part of quadrants 15 <strong>and</strong> 16<br />

in <strong>the</strong> Norwegian North Sea. In this part of <strong>the</strong> Viking Graben <strong>the</strong><br />

main exploration play is <strong>the</strong> submarine fan s<strong>and</strong>s of Paleocene <strong>and</strong><br />

Eocene age. These s<strong>and</strong>s (Balder, Heimdal, <strong>and</strong> Ty formations)<br />

pinch out to <strong>the</strong> east in blocks 25/8 (Jotun field) <strong>and</strong> 25/11 (Balder<br />

<strong>and</strong> Grane fields) <strong>and</strong> along <strong>the</strong> western margin of <strong>the</strong> Utsira High<br />

to form a combination of stratigraphic <strong>and</strong> structural traps. Marine<br />

s<strong>and</strong>s of Middle <strong>and</strong> Late Jurassic age, typically present in rotated<br />

fault blocks, constitute ano<strong>the</strong>r important play.<br />

Geochemical analyses show that <strong>the</strong> Upper Jurassic Draupne<br />

Formation has a good potential for oil generation along <strong>the</strong> entire<br />

western margin <strong>and</strong> nor<strong>the</strong>rn nose of <strong>the</strong> Utsira High. Both upper<br />

<strong>and</strong> lower Draupne source intervals along <strong>the</strong> western graben margin,<br />

however, contain more terrigenous kerogen than in <strong>the</strong> eastern<br />

part of <strong>the</strong> graben. Such change in organic facies <strong>within</strong> <strong>the</strong><br />

Draupne source interval naturally results in a higher proportion of<br />

gas generation <strong>and</strong> <strong>the</strong> possibility for generating a more waxy crude<br />

than typically encountered in <strong>the</strong> Viking Graben. Detection <strong>and</strong><br />

characterization of oil <strong>and</strong> gas shows <strong>within</strong> <strong>the</strong> Tertiary section<br />

permit mapping of <strong>migration</strong> entry points from <strong>the</strong> Jurassic source<br />

<strong>rock</strong>s <strong>and</strong> help delineate secondary <strong>and</strong> tertiary <strong>migration</strong> <strong>pathways</strong><br />

<strong>within</strong> <strong>the</strong> Paleocene–Eocene play.<br />

INTRODUCTION<br />

In this article, we examine <strong>the</strong> <strong>hydrocarbon</strong> systems <strong>within</strong> <strong>the</strong><br />

sou<strong>the</strong>rn part of quadrants 24 <strong>and</strong> 25 <strong>and</strong> <strong>the</strong> nor<strong>the</strong>rn part of quadrants<br />

15 <strong>and</strong> 16 in <strong>the</strong> Norwegian North Sea (Figure 1). We have<br />

Copyright 2001. The American Association of Petroleum Geologists. All rights reserved.<br />

Manuscript received January 26,1999; revised manuscript received June 19,2000; final acceptance<br />

August 31,2000.<br />

AAPG Bulletin, v. 85, no. 5 (May 2001), pp. 861–883 861<br />

AUTHORS<br />

Gary H. Isaksen ExxonMobil Upstream<br />

Research Company, 3120 Buffalo Speedway,<br />

Houston, Texas, 77252;<br />

ghisaks@upstream.xomcorp.com<br />

Gary H. Isaksen is research supervisor for<br />

petroleum geochemistry <strong>and</strong> source <strong>rock</strong> modeling<br />

with ExxonMobil Upstream Research Company.<br />

Isaksen graduated from <strong>the</strong> University of Bergen,<br />

Norway, with an M.S. degree <strong>and</strong> a Ph.D. in<br />

petroleum geochemistry <strong>and</strong> petroleum geology.<br />

Since joining Exxon in 1985, <strong>the</strong> major <strong>the</strong>mes of<br />

his work have been integration of geology <strong>and</strong><br />

organic geochemistry, molecular geochemistry<br />

research, risking of play elements, <strong>and</strong> seep <strong>and</strong><br />

production geochemistry. During 1993–1995 he<br />

worked established <strong>and</strong> frontier plays <strong>within</strong> United<br />

Kingdom <strong>and</strong> Norwegian territories, <strong>and</strong> during<br />

1997–1999 he worked on regional- <strong>and</strong> prospectscale<br />

oil <strong>and</strong> gas assessments <strong>within</strong> Azerbaijan,<br />

Turkmenistan, Uzbekistan, Kazakhstan, <strong>and</strong> Russia,<br />

as well as production geochemistry <strong>within</strong> fields of<br />

<strong>the</strong> South Caspian basin. His current geoscience<br />

focus is on applied research <strong>and</strong> <strong>the</strong> transfer of<br />

geochemical technologies <strong>within</strong> ExxonMobil’s<br />

exploration, development, <strong>and</strong> production<br />

functions.<br />

K. Haakan I. Ledje Esso Norge AS,<br />

N-4033 Forus, Norway;<br />

haakan.karl.ledje@exxonmobil.com<br />

Haakan Ledje is a senior geologist with Esso Norge<br />

in Stavanger, Norway. He received his B.Sc. degree<br />

in sedimentology in 1983 from <strong>the</strong> University of<br />

Lund, Sweden, <strong>and</strong> M.Sc. degree in sedimentology<br />

<strong>and</strong> tectonics in 1985 from <strong>the</strong> University of<br />

California, Los Angeles. He also received an MBA<br />

degree in 1993 from <strong>the</strong> Norwegian School of<br />

Management, Oslo, Norway. He has broad<br />

experience with <strong>hydrocarbon</strong> systems studies on<br />

<strong>the</strong> Norwegian Continental Shelf. During 1998–<br />

1999 he was <strong>the</strong> project leader for ExxonMobil’s<br />

play <strong>and</strong> prospect evaluation offshore mid-Norway.<br />

Ledje has special expertise in source <strong>rock</strong> <strong>quality</strong><br />

assessment <strong>and</strong> <strong>hydrocarbon</strong> <strong>migration</strong> analyses.<br />

ACKNOWLEDGEMENTS<br />

This article is published by permission of Esso<br />

Norge a.s. <strong>and</strong> Enterprise Oil Norwegian a.s. We<br />

also wish to acknowledge Robertson Research International<br />

Ltd., GeoLab Nor a.s., Saga Petroleum<br />

a.s., <strong>and</strong> Conoco Norway Inc. for permission to use<br />

some of <strong>the</strong>ir geochemical data in this article.


Figure 1. Location map of<br />

study area. Samples were obtained<br />

from Norwegian quadrants<br />

24 <strong>and</strong> 25 <strong>and</strong> <strong>the</strong> nor<strong>the</strong>rn<br />

part of quadrants 15<br />

<strong>and</strong> 16.<br />

focused our attention on <strong>the</strong> prospectivity in <strong>the</strong> Utsira<br />

High area, which is a prolific <strong>hydrocarbon</strong> province,<br />

with discovered, producible reserves estimated at 1.0–<br />

1.5 billion st<strong>and</strong>ard bbl. The main group of fields<br />

<strong>within</strong> this area includes Balder, Grane, Jotun, <strong>and</strong><br />

Heimdal. This area continues to be prospective; Esso<br />

Norge in 1997 alone made a series of successful smaller<br />

discoveries including 25/8–10, 25/8–11, <strong>and</strong> 25/10–8.<br />

Most oil <strong>and</strong> gas discoveries to date have been <strong>within</strong><br />

<strong>the</strong> Eocene <strong>and</strong> Paleocene submarine-fan s<strong>and</strong>s (Figure<br />

2). Recent exploration has tested <strong>the</strong> prospectivity of<br />

<strong>the</strong> Jurassic play along <strong>the</strong> eastern margin of <strong>the</strong> graben.<br />

An underst<strong>and</strong>ing of <strong>the</strong> source <strong>rock</strong><strong>quality</strong><br />

<strong>within</strong> <strong>the</strong> drainage area, <strong>hydrocarbon</strong> <strong>migration</strong> <strong>pathways</strong>,<br />

<strong>and</strong> <strong>the</strong> location of <strong>migration</strong> entry points into<br />

Tertiary reservoirs is critical to future exploration<br />

<strong>within</strong> this play. The primary source <strong>rock</strong>s in <strong>the</strong> area<br />

are organic-rich shales of <strong>the</strong> Kimmeridgian to Vol-<br />

862 Utsira High Area (Norwegian North Sea)<br />

gian–Ryazanian Draupne Formation. The Oxfordian<br />

Hea<strong>the</strong>r Formation forms a secondary source. Regional<br />

studies of Draupne (Kimmeridge Clay) <strong>and</strong> Hea<strong>the</strong>r<br />

formations source <strong>rock</strong>facies <strong>and</strong> <strong>hydrocarbon</strong> distributions<br />

in this part of <strong>the</strong> Viking Graben have been<br />

carried out by several researchers. Early overviews include<br />

those by Skarpnes et al. (1981), Barnard <strong>and</strong><br />

Cooper (1981), <strong>and</strong> Barnard et al. (1981), who reported<br />

on source <strong>rock</strong><strong>quality</strong> variations from bulkkerogen<br />

<strong>and</strong> pyrolysis data throughout platform <strong>and</strong> graben<br />

areas. Goff (1983), Cornford (1984), Field (1985),<br />

Cooper <strong>and</strong> Barnard (1984), Thomas et al. (1985), <strong>and</strong><br />

Harris <strong>and</strong> Fowler (1987) followed with more detailed<br />

studies as more data became available. For example,<br />

Thomas et al. (1985) carried out geochemical characterization<br />

of cores taken at a variety of structural positions<br />

in <strong>the</strong> basin <strong>and</strong> at different <strong>the</strong>rmal maturity<br />

levels. A common observation from <strong>the</strong> regional stud-


Figure 2. Lithostratigraphy of <strong>the</strong> greater Utsira High area.<br />

Significant discoveries have been made in <strong>the</strong> Paleocene Balder,<br />

Heimdal, <strong>and</strong> Ty formations <strong>and</strong> in <strong>the</strong> Jurassic Draupne <strong>and</strong><br />

Hugin formations. The Upper Jurassic Draupne <strong>and</strong> Hea<strong>the</strong>r formations<br />

are <strong>the</strong> main source <strong>rock</strong>s for oil <strong>and</strong> gas.<br />

ies published during <strong>the</strong> early to mid-1980s was that<br />

<strong>the</strong> Draupne Formation indeed showed organic facies<br />

variations but that such intra-Draupne variability (Huc<br />

et al., 1985) was only a secondary influence on <strong>the</strong><br />

distribution of oil-prone <strong>and</strong> gas-prone areas in <strong>the</strong><br />

North Sea. <strong>Source</strong> <strong>rock</strong><strong>the</strong>rmal maturity was viewed<br />

as <strong>the</strong> primary control on oil vs. gas.<br />

Cooper et al. (1995) provided a more detailed picture<br />

of biomarker <strong>and</strong> stable carbon isotopic compositions<br />

for <strong>the</strong> Kimmeridge Clay, pointing out that <strong>the</strong><br />

organic enrichment was mostly a function of dysaerobic<br />

to anaerobic development <strong>within</strong> submarine rifts<br />

<strong>and</strong> associated sills. Controls on organic-facies development<br />

of <strong>the</strong> Draupne Formation were modeled by<br />

Miller (1990), using paleoenvironmental, geochemical,<br />

<strong>and</strong> petrophysical data sets, whereas Ungerer et al.<br />

(1985) modeled oil <strong>and</strong> gas yields <strong>and</strong> <strong>migration</strong> <strong>pathways</strong><br />

in <strong>the</strong> Frigg area.<br />

Coastal plain <strong>and</strong> deltaic units in <strong>the</strong> Middle Jurassic<br />

Brent <strong>and</strong> Sleipner/Hugin formations are prevalent<br />

in <strong>the</strong> North Viking Graben <strong>and</strong> South Viking<br />

Graben, respectively. Such geographically extensive<br />

coaly <strong>rock</strong>s, however, are not thought to be present in<br />

<strong>the</strong> greater Utsira High area.<br />

The organic-rich Permian Kupferschiefer Formation<br />

was also deposited this far north in <strong>the</strong> Viking Graben<br />

(e.g., well 25/10–2), but is generally too thin to<br />

yield significant volumes of oil <strong>and</strong> gas. Petroleum geochemical<br />

studies of <strong>the</strong> Kupferschiefer Formation (aka<br />

Marl Slate) were reported by Dungworth (1972) <strong>and</strong><br />

Gibbons (1978), who both recognized its source <strong>rock</strong><br />

potential.<br />

Study Objectives<br />

The main objectives of this study were to (1) characterize<br />

<strong>and</strong> map <strong>the</strong> regional source <strong>rock</strong><strong>quality</strong>,<br />

(2) measure, <strong>and</strong> predict beyond points of control, oil<br />

vs. gas yields from regionally extensive source <strong>rock</strong>s,<br />

<strong>and</strong> (3) correlate <strong>the</strong> molecular signatures from shows<br />

<strong>and</strong> stains to assess <strong>the</strong>ir likely origin <strong>and</strong> secondary<strong>migration</strong><br />

<strong>and</strong> tertiary-<strong>migration</strong> <strong>pathways</strong>.<br />

GEOLOGICAL SETTING<br />

The Utsira High is a large basement high, flanked by<br />

<strong>the</strong> Viking Graben to <strong>the</strong> west <strong>and</strong> <strong>the</strong> Stord Basin to<br />

<strong>the</strong> east. The present structural configuration was inherited<br />

from extensional tectonism that occurred in<br />

<strong>the</strong> late Paleozoic <strong>and</strong> Mesozoic. At <strong>the</strong> end of <strong>the</strong><br />

Isaksen <strong>and</strong>Ledje 863


Rotliegende, two large intracratonic basins occupied an<br />

area stretching from <strong>the</strong> Republic of Pol<strong>and</strong> <strong>and</strong> nor<strong>the</strong>rn<br />

Germany to <strong>the</strong> middle parts of <strong>the</strong> Viking Graben.<br />

The so-called nor<strong>the</strong>rn Permian Basin is thought to<br />

have been invaded by marine waters from <strong>the</strong><br />

Norwegian–Greenl<strong>and</strong> Sea in <strong>the</strong> north, thus forming<br />

<strong>the</strong> Zechstein Sea. The relatively shallow water depth<br />

of large parts of <strong>the</strong> area (less than 250 m according to<br />

Smith [1980] <strong>and</strong> Gibbons [1987]), combined with an<br />

arid climate, led to elevated salinity levels that could<br />

host only a few adaptable species (Paul, 1986a) <strong>and</strong> a<br />

chemically stratified water column (Hirst <strong>and</strong> Dunham,<br />

1963). These resultant organic-rich shales constitute<br />

<strong>the</strong> Kupferschiefer Formation. During <strong>the</strong> latest<br />

Permian <strong>the</strong> global sea level was at a significant low<br />

due to cooling <strong>and</strong> contraction of <strong>the</strong> oceanic lithosphere<br />

(Ziegler, 1988). Over much of <strong>the</strong> nor<strong>the</strong>rn<br />

Permian Basin, <strong>the</strong> Kupferschiefer Formation is recorded<br />

as a thin (0.5 m) blackshale (Oszczepalski,<br />

1986; Paul, 1986b;). In well 25/10–2 in <strong>the</strong> Norwegian<br />

North Sea it occurs as a 5 m section <strong>and</strong> is readily identified<br />

by <strong>the</strong> gamma-ray logging tool because of its high<br />

content of radioactive elements.<br />

The Triassic was a period of accelerated crustal extension<br />

with <strong>the</strong> formation of large, rotated fault<br />

blocks. Rocks of this age <strong>within</strong> <strong>the</strong> Viking Graben are<br />

typically nonmarine clastics with development of red<br />

beds in an arid to semiarid climate. Subsequent rifting<br />

<strong>and</strong> subsidence during <strong>the</strong> Jurassic Kimmerian tectonic<br />

phases resulted in <strong>the</strong> present structure of <strong>the</strong> graben.<br />

The largest pulse of rifting associated with <strong>the</strong> most<br />

rapid phase of extension occurred during <strong>the</strong> Middle<br />

Jurassic (mid-Kimmerian phase) (Curtin <strong>and</strong> Ballestad,<br />

1986). Erosion of <strong>the</strong> uplifted rift flanks led to deposition<br />

of nonmarine s<strong>and</strong>stones on <strong>the</strong> western flankof<br />

<strong>the</strong> Utsira High. These Middle Jurassic s<strong>and</strong>stones are<br />

important reservoir targets.<br />

Establishment of <strong>the</strong> rift system in <strong>the</strong> Late Jurassic<br />

coupled with eustatic sea level rise led to widespread<br />

transgression <strong>and</strong> deep-water sedimentation<br />

(Ziegler, 1988). Under <strong>the</strong>se conditions <strong>the</strong> Kimmeridgian<br />

to Ryazanian organic-rich shales of <strong>the</strong><br />

Draupne Formation were deposited. These shales are<br />

<strong>the</strong> primary source <strong>rock</strong>s in <strong>the</strong> Viking Graben. Concurrently,<br />

<strong>the</strong> western margin of <strong>the</strong> South Viking Graben<br />

experienced extensive <strong>and</strong> rapid syntectonic sedimentation<br />

from <strong>the</strong> graben scarps resulting in<br />

mass-flow sediments that thin abruptly eastward <strong>and</strong><br />

interfinger with <strong>the</strong> organic-rich shales (Figure 3). During<br />

<strong>the</strong> earliest Cretaceous a new rifting phase (late<br />

Kimmerian phase) affected <strong>the</strong> area. This phase reac-<br />

864 Utsira High Area (Norwegian North Sea)<br />

tivated existing faults <strong>and</strong> created fur<strong>the</strong>r fault-block<br />

relief. This late Kimmerian rift phase was mostly concentrated<br />

along <strong>the</strong> major north-south fault trend to<br />

<strong>the</strong> west, resulting in an asymmetric Upper Jurassic–<br />

earliest Cretaceous basin. More uniform subsidence<br />

occurred in Cretaceous <strong>and</strong> Tertiary times (Curtin <strong>and</strong><br />

Ballestad, 1986)<br />

The dominant Tertiary event was <strong>the</strong> deposition<br />

of Paleocene <strong>and</strong> lower Eocene deep-marine s<strong>and</strong>stones.<br />

These s<strong>and</strong>s were shed from <strong>the</strong> uplifted Shetl<strong>and</strong><br />

Platform to <strong>the</strong> west <strong>and</strong> deposited into <strong>the</strong> basin<br />

by eastward-prograding deltaic complexes <strong>and</strong> turbidity<br />

currents during low st<strong>and</strong>s of sea level. The Paleocene<br />

<strong>and</strong> Eocene s<strong>and</strong>s are <strong>the</strong> primary reservoir targets<br />

in <strong>the</strong> Utsira High area.<br />

SAMPLES ANALYZED<br />

Samples of organically enriched Jurassic shales were<br />

collected from wells in <strong>the</strong> Norwegian quadrants 15,<br />

16, 24, <strong>and</strong> 25 (Table 1). The study is comprehensive<br />

in that 246 samples were analyzed for organic richness<br />

by total organic carbon (TOC), 239 samples were analyzed<br />

for source <strong>quality</strong> <strong>and</strong> <strong>the</strong>rmal maturity by<br />

Rock-Eval pyrolysis, <strong>and</strong> 447 readings were taken of<br />

vitrinite reflectance from 12 well profiles. Selected<br />

samples were also analyzed for <strong>the</strong>ir content of cuttings<br />

gas for five well profiles (Table 2) <strong>and</strong> for fluid inclusions<br />

from 33 samples (Table 3). A special investigation<br />

of shows <strong>and</strong> stains was performed on <strong>the</strong> 18 samples<br />

listed in Table 4.<br />

RESULTS AND DISCUSSION<br />

<strong>Source</strong> Rock Quality<br />

Herein, we consider a shale (without <strong>hydrocarbon</strong><br />

staining or coaly fragments) to be oil prone when its<br />

original TOC <strong>and</strong> hydrogen index (HI) exceeds about<br />

1.5 wt. % <strong>and</strong> 250–300 mg HC/g organic carbon (org<br />

C), respectively, <strong>and</strong> its <strong>the</strong>rmal or solvent extract is<br />

dominated by a well-defined distribution of aliphatic<br />

compounds. Fur<strong>the</strong>r supporting evidence includes<br />

richly fluorescent algal or amorphous organic matter as<br />

seen by visual kerogen analyses. In our study area, <strong>the</strong>se<br />

shales also display organic stringers <strong>within</strong> a laminated<br />

clay matrix (thin-section analyses). Gas-prone shales in<br />

<strong>the</strong> North Sea are characterized by original HI values<br />

below 200 mg HC/g org C, visually identifiable


Figure 3. Schematic cross section of <strong>the</strong> Central Viking Graben from north of <strong>the</strong> Brae fields in <strong>the</strong> west, through <strong>the</strong> Gudrun field (15/3), to <strong>the</strong> Balder field complex in <strong>the</strong><br />

east. The figure depicts structural styles, maturation levels, <strong>and</strong> preferred <strong>hydrocarbon</strong> <strong>migration</strong> <strong>pathways</strong> <strong>and</strong> entry points into tertiary reservoirs. Gamma-ray log signatures<br />

are added for select wells. The location of <strong>the</strong> cross section is shown in Figure 1.<br />

Isaksen <strong>and</strong>Ledje 865


Table 1. Geochemical Data for Wells Studied*<br />

Visual OMT<br />

Number of<br />

Readings<br />

Ro/TOC/HI Rock-Eval Pyrolysis<br />

Thickness Maturity<br />

<strong>Source</strong> Rock<br />

Rating<br />

Chemical<br />

Kerogen<br />

Classification<br />

TOCo (%) HI HIo TOC<br />

(%)<br />

Vitrinite GCData<br />

Gas-Prone Pr/Ph CPI<br />

Intertinite<br />

Coal<br />

Liptinite<br />

Oil-Prone<br />

Ro (%) LOM<br />

Net<br />

(m)<br />

Net<br />

(m)<br />

Base<br />

(m)<br />

Top<br />

(m)<br />

Well Zone<br />

15/3-1 A B 3947 4083 136 96 0.6 8.4 87% 1% 12% 1.3 1.0 5.6 5.6 324 385 II/III-III/II Good oil 20/12/12<br />

C4083 4753 670 410 0.72 9.4 56% 14% 30% 1.4 1.1 5.9 5.9 96 100 III Good gas 40/23/23<br />

D 4753 4986 233 233 1.08 11.1 50% 13% 37% 1.6 1.0 4.2 4.2 24 24 IV/III – 31/4/4<br />

15/3-2 A B 4236 4301 65 65 0.76 9.6 47% 1% 52% 1.3 1.0 4.1 4.1 136 290 III-III/II Fair oil/trans 30/8/8<br />

C4301 4352 51 51 – 9.8 – – – 1.2 1.0 4.9 4.9 80 100 III Fair gas 0/3/3<br />

D 4352 4400 48 48 0.94 10.5 35% 5% 60% 1.4 1.4 2.9 2.9 41 41 IV-III – 3/3/3<br />

15/3-3 A B 4018 4018 162 162 0.61 8.5 0% 0% 100% 0.9 1.0 6.2 6.2 151 220 III-III/II Fair oil/trans 13/6/4<br />

C4180 4450 270 150 0.88 10.2 0% 0% 100% 5.2 5.2 83 100 III Good gas 14/7/4<br />

D 4450 4452 72 72 – 10.6 – – – 1.5 1.0 5.5 5.5 105 150 III Good gas 0/2/1<br />

15/5-1 A B 3492 3543 51 51 0.58 8.2 75% 8% 17% 1.2 1.0 6.0 6.0 309 400 III/II-II/III Good oil 27/6/6<br />

C3543 3553 10 10 – 8.2 – – – – – 6.3 6.3 229 300 III-III/II Good oil 0/1/1<br />

D 3553 3558 5 5 – 8.2 60% 10% 30% 1.0 1.0 5.6 5.6 275 380 II/II Good oil 0/1/1<br />

16/1-2 A B 2242 2298 56 11 – 6.8 70% 20% 10% – – 3.7 3.7 490 510 II Good oil 0/1/1<br />

C2298 2390 92 92 0.48 7.0 74% 14% 12% 1.5 1.1 3.4 3.4 462 490 II-II/III Good oil 19/16/16<br />

D 2390 2424 34 34 0.46 7.0 10% 20% 70% 1.9 1.1 3.9 3.9 253 290 III/II Fair oil 14/4/4<br />

16/1-3 A – B 2707 2707 0 0 – – – – – – – – – – – – – –<br />

C2707 2717 10 10 – 7.2 95% 5% Tr – – 3.9 3.9 275 300 III/II Fair oil 0/1/1<br />

D 2717 2730 13 13 0.50 7.2 – – – 1.8 1.2 4.1 4.1 284 300 III/II Fair oil 40/3/3<br />

24/9-1 A B 4330 4797 467 467 0.94 10.6 41% 3% 56% 1.6 1.0 5.4 7.8 141 460 II/II Excel. oil 52/29/26<br />

C4797 4907 110 110 1.17 11.2 33% 12% 55% 1.6 1.0 3.8 4.8 112 320 III/II-III Good oil 10/9/9<br />

D Not penetrated<br />

24/2-1 A B 4055 4094 39 39 1.00 10.8 55% Tr 45% 1.1 1.1 6.2 6.5 122 200 III-III/II Good gas 4/7/07<br />

C4094 4140 46 46 0.96 10.8 50% Tr 50% 1.3 4.9 5.6 119 280 15.7 III-III/II Fair oil/trans 25/8/8<br />

D 4140 4163 23 23 – 10.8 25% 10% 65% 1.5 1.0 3.4 4.1 150 300 II/II Fair oil 0/2/2<br />

24/12-2 A B 4261 4565 304 304 0.89 10.2 58% 13% 29% 1.4 1.1 4.6 5.1 40 220 III-III/II Fair oil/trans 28/29/29<br />

C4565 4640 75 75 0.92 10.6 50% 40% 10% 1.2 0.9 5.9 5.9 49 49 IV – 4/7/07<br />

D 4640 4955 315 315 1.28 11.6 32% 22% 46% 1.6 1.1 4.5 4.5 15 15 IV/III – 23/21/21<br />

25/4-1 A B 3171 3181 10 10 0.42 6.0 50% 50% Tr – – 8.5 8.5 138 138 III Good gas 4/1/02<br />

C3181 3184 3 3 – – – – – – – – – – – – –<br />

D 3184 3185 1 1 – – – – – – – – – – – – –<br />

25/6-1 A B 2233 2253 20 20 – 6.0 95% 5% 0% 1.0 1.0 7.4 7.4 573 573 II Excel. oil 0/6/4<br />

C2253 2256 3 3 – 6.0 95% 5% 0% 0.8 1.1 3.6 3.6 240 240 III-III/II Fair oil/trans 0/2/2<br />

D 2256 2297 41 41 – 6.0 90% 10% 0% 1.3 1.1 5.5 5.5 374 374 III-III/II Good oil 0/8/8<br />

25/7-2 A – B 4056 4119 63 63 0.88 10.2 95% Tr 5% 1.1 1.0 3.9 8.6 320 740 II-II/III Excel. oil 11/7/7<br />

C4119 4293 174 45 1.00 10.8 – – – 2.1 1.1 1.7 3.2 245 580 II/III-II Good oil 14/2/2<br />

D 4293 4407 114 114 1.14 11.2 75% 10% 15% 2.8 1.1 1.6 5.1 246 830 II Excel. oil 21/7/7<br />

866 Utsira High Area (Norwegian North Sea)<br />

*Zones A B, upper Draupne Fm; Zone C, lower Draupne Fm; Zone D, Hea<strong>the</strong>r Fm; LOM, level of organic maturity; Pr, pristane; Ph, phytane; CPI, carbon preference index (odd/even n-alkanes); TOCo, original total organic carbon; HIo, original hydrogen index.


Table 2. Well-Sections Evaluated for Cuttings-Gas Content<br />

Well<br />

Depth Interval<br />

(m) Age Interval<br />

16/1-1 350–3000 Pleistocene–Upper Cretaceous<br />

16/1-2 300–3000 Pleistocene–Rotliegende (Permian)<br />

16/1-3 2000–3200 Eocene–Zechstein (Permian)<br />

25/8-2 200–2600 Pleistocene–Triassic<br />

25/10-2 2250–3200 Undiff. Tertiary–Rotliegende<br />

Table 3. Samples Analyzed for Possible Content of Fluid<br />

Inclusions*<br />

Well Depth (m) Formation or Age Sample Type<br />

16/1-1 2321 Heimdal Core<br />

16/1-1 2321 Heimdal Core<br />

16/1-1 2330 Heimdal Core<br />

16/1-1 2348 Heimdal Core<br />

16/1-1 2419 Heimdal Core<br />

16/1-1 2427 Heimdal Core<br />

16/1-1 2533 Heimdal Core<br />

16/1-1 2648 Heimdal Core<br />

16/1-2 2102 Heimdal Cuttings<br />

16/1-2 2108 Heimdal Cuttings<br />

16/1-2 2114 Heimdal Cuttings<br />

16/1-2 2240 Upper Jurassic Cuttings<br />

16/1-2 2246 Upper Jurassic Cuttings<br />

16/1-2 2252 Upper Jurassic Cuttings<br />

16/1-2 2258 Upper Jurassic Cuttings<br />

16/1-2 2264 Upper Jurassic Cuttings<br />

16/1-2 2270 Upper Jurassic Cuttings<br />

16/1-2 2273 Late Jurassic Cuttings<br />

16/1-2 2423 Triassic(?) Cuttings<br />

16/1-2 2426 Triassic(?) Cuttings<br />

16/1-2 2429 Triassic(?) Cuttings<br />

25/10-2 1893 Heimdal Core<br />

25/10-2 2013 Heimdal Core<br />

25/10-2 2014 Heimdal Core<br />

25/10-2 2015 Heimdal Core<br />

25/10-2 2015 Heimdal Core<br />

25/10-2 2023 Heimdal Core<br />

25/10-2 2088 Heimdal Core<br />

25/10-2 2100 Heimdal Core<br />

25/10-2 2102 Heimdal Core<br />

25/10-2 2684 Triassic(?) Cuttings<br />

25/10-2 2687 Triassic(?) Cuttings<br />

25/10-2 2690 Triassic(?) Cuttings<br />

*Courtesy of R. J. Pottorf, ExxonMobil Upstream Research Company.<br />

Table 4. Samples Selected for Thermal <strong>and</strong> Solvent<br />

Extraction because of Identified Shows <strong>and</strong> Stains<br />

Well Depth (m) Sample Type<br />

16/1-2 2100 Shale<br />

16/1-2 2130 Shale<br />

16/1-2 2250 Shale/silt/s<strong>and</strong>stone<br />

16/1-2 2102 Silty shale<br />

16/1-2 2108 Silty shale<br />

16/1-2 2114 Silty shale<br />

25/10-2 RE 1983 Shale<br />

25/10-2 RE 2013 Shale<br />

25/10-2 RE 2682 S<strong>and</strong>stone<br />

25/10-2 RE 2713 S<strong>and</strong>stone<br />

25/10-2 RE 3007 Shale (Kupferschiefer)<br />

25/10-2 RE 3009 Conglomerate<br />

25/10-2 RE 3037 Conglomerate<br />

16/1-1 1920 Shale/silt<br />

16/1-1 1936 Shale/silt<br />

16/1-1 2530 Silt/shale<br />

16/1-1 2667 Silt/s<strong>and</strong>stone<br />

16/1-1 2743 S<strong>and</strong>stone/shale<br />

vitrinitic macerals, <strong>and</strong> an aromatic-rich composition.<br />

Assessment of <strong>the</strong> oil vs. gas potential of humic coals<br />

is somewhat more complex. Middle Jurassic humic<br />

coals in <strong>the</strong> North Sea are primarily gas prone but do<br />

have <strong>the</strong> capability to expel aliphatic-rich, volatile oil<br />

(Isaksen et al., 1998a, b). Such coals are known from<br />

<strong>the</strong> Sleipner/Hugin <strong>and</strong> Brent delta systems to <strong>the</strong><br />

south <strong>and</strong> north of our study area, respectively. No<br />

humic coals are thought to be present in <strong>the</strong> area of<br />

<strong>the</strong> Viking Graben to <strong>the</strong> west of <strong>the</strong> nor<strong>the</strong>rn Utsira<br />

High.<br />

The primary source <strong>rock</strong>in <strong>the</strong> Utsira High area is<br />

<strong>the</strong> organic-rich, oil-prone, Kimmeridgian to Volgian–<br />

Ryazanian Draupne Formation. A secondary source<br />

<strong>rock</strong>, which has poor to fair potential for oil, is <strong>the</strong><br />

Oxfordian Hea<strong>the</strong>r Formation. The Permian Kupferschiefer<br />

source is known to extend this far north <strong>and</strong><br />

was penetrated by well 25/10–2. Although <strong>the</strong> Kupferschiefer<br />

shale contains a predominance of oil-prone<br />

algal <strong>and</strong> amorphous organic matter, <strong>the</strong> lackof sufficient<br />

source volume precludes it from contributing<br />

with any significance to <strong>the</strong> reservoired oils in this area.<br />

No contribution from <strong>the</strong> Kupferschiefer Formation<br />

was observed among <strong>the</strong> biomarkers of <strong>the</strong> reservoired<br />

oils. Fur<strong>the</strong>rmore, molecular signatures in oils <strong>and</strong><br />

Isaksen <strong>and</strong>Ledje 867


gases in <strong>the</strong> area do not suggest presence of Middle<br />

Jurassic woody/coaly source <strong>rock</strong>s.<br />

During <strong>the</strong> early Callovian a marine transgression<br />

tookplace throughout <strong>the</strong> Viking Graben. The relative<br />

rise in sea level continued into <strong>the</strong> earliest Cretaceous<br />

with deposition of time-transgressive marine shales of<br />

<strong>the</strong> Draupne Formation. Pyrolysis, solvent extraction,<br />

<strong>and</strong> visual kerogen data (Table 1) show that <strong>the</strong>se<br />

shales contain oil-prone kerogen that has predominantly<br />

liptintic organic matter in <strong>the</strong> form of marine,<br />

algal bodies <strong>and</strong> lipid-rich amorphous material. The<br />

oil-prone nature of <strong>the</strong> amorphous kerogen is supported<br />

by elevated hydrogen indices (400 mg HC/g<br />

org C) <strong>and</strong> paraffinic compositions for those samples<br />

where <strong>the</strong> amorphous material accounts for <strong>the</strong> bulk<br />

of <strong>the</strong> kerogen. We do not imply a common origin for<br />

<strong>the</strong> algal <strong>and</strong> amorphous material. Although some percentage<br />

of <strong>the</strong> amorphous material may be linked to<br />

an algal origin through bacterial degradation (seen as<br />

severely degraded algal bodies), <strong>the</strong> bulkof <strong>the</strong> amorphous<br />

material may indeed have been enriched by<br />

o<strong>the</strong>r mechanisms, such as a predominantly bacterial<br />

origin or a preferential adsorption of amorphous material<br />

on certain clays (Mayer, 1994). For <strong>the</strong> most part,<br />

<strong>the</strong>se shales are interpreted to have accumulated under<br />

anoxic bottom-water conditions that helped preserve<br />

<strong>the</strong>ir oil-prone nature. The presence of finely laminated<br />

shales suggests a lackof, or low abundance of,<br />

grazing or burrowing organisms. Oil-prone organic<br />

matter may also be preserved under suboxic conditions<br />

(Isaksen <strong>and</strong> Bohacs, 1995).<br />

From its wire-line log response (gamma-ray, resistivity,<br />

<strong>and</strong> sonic log signatures), we subdivided <strong>the</strong><br />

Draupne Formation into an upper <strong>and</strong> lower unit (Figure<br />

4). The Upper Jurassic sequence on <strong>the</strong> west side<br />

of <strong>the</strong> Viking Graben contains extensive mass-flow<br />

sediments in <strong>the</strong> Brae Formation, representing rapid<br />

syntectonic sedimentation from <strong>the</strong> graben escarpment<br />

(Figure 5). The s<strong>and</strong>s of <strong>the</strong> Brae Formation are interbedded<br />

with open-marine sediments of <strong>the</strong> Draupne<br />

Formation <strong>and</strong> pinch out eastward (Figure 3). As a result,<br />

<strong>the</strong> Upper Jurassic has a complex lithofacies distribution<br />

<strong>and</strong> has seen a dilution of <strong>the</strong> oil-prone kerogen<br />

of <strong>the</strong> basinal Draupne shale by l<strong>and</strong>-derived<br />

gas-prone kerogen, especially along <strong>the</strong> western graben<br />

margin. Although such mass-flow deposits can oxygenate<br />

anoxic sediments or anoxic <strong>and</strong> stratified bottom<br />

waters, <strong>the</strong> presence of low amounts of oxygen is not<br />

considered detrimental to preservation of organic matter<br />

(Heinrichs <strong>and</strong> Reeburgh, 1987; Lee, 1992). In our<br />

opinion, <strong>the</strong> type of organic matter exerts a far greater<br />

868 Utsira High Area (Norwegian North Sea)<br />

control on oil vs. gas potential than does <strong>the</strong> oxidation<br />

of organic matter <strong>within</strong> <strong>the</strong> dysoxic zone.<br />

Interpretations of seismic data <strong>and</strong> wire-line log responses<br />

for <strong>the</strong> Upper Jurassic sediments enabled <strong>the</strong><br />

construction of an Upper Jurassic isopach map as<br />

shown by Figure 6. This has been extended with earlier<br />

work (Isaksen et al., 1997) to include <strong>the</strong> South Viking<br />

Graben in <strong>the</strong> greater Sleipner–Tiffany–Brae area. In<br />

this context, <strong>the</strong> Upper Jurassic isopach thicknesses<br />

represent <strong>the</strong> sedimentary section from base Callovian<br />

to base Cretaceous. In <strong>the</strong> graben area, between <strong>the</strong><br />

Balder <strong>and</strong> Gryphon fields, <strong>the</strong> Upper Jurassic attains<br />

thicknesses of 3 km, of which a considerable thickness<br />

constitutes <strong>the</strong> coarser clastics shed off from <strong>the</strong> East<br />

Shetl<strong>and</strong> Platform.<br />

Rock-Eval pyrolysis <strong>and</strong> TOC analyses (Table 1)<br />

were carried out on 14 wells <strong>within</strong> <strong>the</strong> greater Utsira<br />

High area to evaluate <strong>the</strong> organic richness, <strong>quality</strong>, <strong>and</strong><br />

<strong>the</strong>rmal maturity of source intervals. To compare <strong>and</strong><br />

assess source <strong>rock</strong>s, which are currently at different<br />

maturities, we used a technique derived from <strong>the</strong> work<br />

of Cooles et al. (1986) <strong>and</strong> Pepper (1991), whereby<br />

measured hydrogen index (HI) <strong>and</strong> TOC values are<br />

back-calculated to <strong>the</strong>ir original (immature) HI o <strong>and</strong><br />

TOC o values according to <strong>the</strong>ir <strong>the</strong>rmal history. The<br />

oil vs. gas potential of <strong>the</strong>se <strong>rock</strong>s (Figure 7) follow <strong>the</strong><br />

definitions discussed previously.<br />

The HI is commonly used as an approximation of<br />

<strong>the</strong> hydrogen content of <strong>the</strong> kerogen <strong>and</strong> thus as an<br />

indicator of <strong>the</strong> source <strong>rock</strong><strong>quality</strong> in terms of oilgeneration<br />

vs. gas-generation potential. This is not necessarily<br />

<strong>the</strong> case for source <strong>rock</strong>s that have a predominance<br />

of terrigenous higher plant organic matter (type<br />

III) because resinite macerals can contribute to <strong>the</strong> HI<br />

value but have a poor potential for generation of C 8<br />

aliphatic <strong>hydrocarbon</strong>s (Isaksen et al., 1998a). The HI<br />

can be plotted against <strong>the</strong> T max to classify kerogen<br />

types <strong>and</strong> evaluate <strong>the</strong> level of <strong>the</strong>rmal (organic) maturity<br />

(Figure 8). This chemical kerogen classification<br />

shows that both upper <strong>and</strong> lower Draupne source intervals<br />

along <strong>the</strong> western graben margin contain more<br />

type III <strong>and</strong> III/II kerogen than in <strong>the</strong> eastern part of<br />

<strong>the</strong> graben (Figure 9A, B). This is in agreement with<br />

<strong>the</strong> expected strong terrigenous influx of woody material<br />

associated with <strong>the</strong> rapid syntectonic sedimentation<br />

from <strong>the</strong> graben scarps. The eastward extension<br />

of type III kerogen seems to be associated with <strong>the</strong> easternmost<br />

extension of mass-flow sediments from <strong>the</strong><br />

western graben margin.<br />

To fur<strong>the</strong>r investigate <strong>the</strong> indications of dilution of<br />

<strong>the</strong> oil-prone kerogen of <strong>the</strong> Draupne shale along <strong>the</strong>


western graben margin, <strong>the</strong> content of visual organic<br />

matter type (OMT) was plotted for <strong>the</strong> analyzed wells<br />

(Figure 10A, B). Wells along <strong>the</strong> eastern margin consistently<br />

show a higher content of oil-prone liptinite<br />

kerogen macerals for both <strong>the</strong> upper <strong>and</strong> lower<br />

Draupne formations. Wells along <strong>the</strong> western margin<br />

of <strong>the</strong> graben conversely have more vitrinite (gas-<br />

Figure 4. Subdivision of <strong>the</strong><br />

Jurassic sedimentary section<br />

based on wire-line log signatures.<br />

The logs shown are<br />

gamma-ray, sonic, <strong>and</strong> resistivity.<br />

The right column shows <strong>the</strong><br />

calculated percent TOCbased<br />

on <strong>the</strong> technique by Passey et<br />

al. (1990).<br />

prone) kerogen content. Incorporation of visual kerogen<br />

data allows <strong>the</strong> interpreter to better assess whe<strong>the</strong>r<br />

<strong>the</strong> oil-prone vs. gas-prone nature of <strong>the</strong> source <strong>rock</strong>is<br />

primarily due to organic matter type or organic matter<br />

degradation through severe oxidation.<br />

The ratings of both upper <strong>and</strong> lower Draupne<br />

source intervals (Figure 11A, B) indicate that a<br />

Isaksen <strong>and</strong>Ledje 869


Figure 5. Main depositional<br />

environments for <strong>the</strong> Upper Jurassic<br />

Draupne Formation. The<br />

extensive mass-flow sediments<br />

represent syntectonic sedimentation<br />

from <strong>the</strong> western graben<br />

escarpment. As a result, offshore,<br />

marine shales are interbedded<br />

with mass-flow sediments.<br />

Shoreface sedimentation<br />

dominated around <strong>the</strong> Utsira<br />

High.<br />

predominantly gas-prone organic facies was deposited<br />

in <strong>the</strong> graben west of blocks 16/1 <strong>and</strong> 25/10. We interpret<br />

<strong>the</strong> decrease in oil potential to be mainly due<br />

to <strong>the</strong> syntectonic influx of turbidites with terrigenous<br />

organic matter during <strong>the</strong> Kimmeridgian period. The<br />

870 Utsira High Area (Norwegian North Sea)<br />

magnitude of a potential decrease in <strong>the</strong> oil-prone nature<br />

of <strong>the</strong> kerogen during accumulation under more<br />

oxic conditions (i.e., oxygen-rich turbidite currents)<br />

remains unresolved. Laminated blackshales sampled<br />

from wells drilled on structural highs suggest a sparse


Figure 6. Isopach map (contours<br />

given in meters) of <strong>the</strong><br />

Upper Jurassic representing <strong>the</strong><br />

section between <strong>the</strong> base Callovian<br />

<strong>and</strong> <strong>the</strong> base Cretaceous.<br />

Isaksen <strong>and</strong>Ledje 871


Figure 7. <strong>Source</strong> richness <strong>and</strong><br />

source <strong>quality</strong> chart for <strong>the</strong> upper<br />

<strong>and</strong> lower Draupne formations.<br />

Each data point represents<br />

<strong>the</strong> average weighted<br />

values from cuttings or core<br />

samples from <strong>the</strong> interval. The<br />

four samples from <strong>the</strong> Lower<br />

Jurassic at <strong>the</strong> base of <strong>the</strong> chart<br />

denote HI values less than 100<br />

mg/g organic carbon.<br />

benthic fauna at <strong>the</strong>se locations. It seems probable,<br />

however, that <strong>the</strong> Draupne shales experienced a higher<br />

degree of dilution by coarser clastics as well as periodic<br />

oxygenation from turbidite currents in basinal positions<br />

connected with debris <strong>and</strong> turbidite flows from<br />

<strong>the</strong> western basin margin. For <strong>the</strong> upper Draupne, <strong>the</strong><br />

gas-prone organic facies is confined to block24/12,<br />

whereas for <strong>the</strong> lower Draupne gas-prone organic facies<br />

also extends into block15/3 to <strong>the</strong> south. The<br />

potential for <strong>the</strong> Draupne Formation to generate oil<br />

improves toward <strong>the</strong> east, from a mixed gas-prone <strong>and</strong><br />

oil-prone organic facies in <strong>the</strong> eastern parts of blocks<br />

24/12 <strong>and</strong> 15/3 to an oil-prone organic facies closest<br />

to <strong>the</strong> Utsira High. At <strong>the</strong> time of our study, no source<br />

<strong>rock</strong>samples were available from <strong>the</strong> Beryl field (UK<br />

9/13) source kitchen (north of 5930N).<br />

<strong>Source</strong> Rock Thermal Maturity<br />

T max measurements from Rock-Eval pyrolysis span a<br />

range from immature (430C) to fully mature (455C)<br />

(Figure 8). Care was taken to avoid T max data derived<br />

from shales stained by in-migrated <strong>hydrocarbon</strong>s (typ-<br />

872 Utsira High Area (Norwegian North Sea)<br />

ically seen as a bimodal S 2 peak). As a fur<strong>the</strong>r control<br />

on <strong>the</strong> <strong>the</strong>rmal maturity vs. depth trend, comparison<br />

was made with vitrinite reflectance (% R o) measurements<br />

on Middle Jurassic coals <strong>and</strong> coaly shales from<br />

locations throughout <strong>the</strong> entire Viking Graben where<br />

such <strong>rock</strong>s are present (Sleipner/Hugin <strong>and</strong> Brent delta<br />

systems). The resultant <strong>the</strong>rmal maturity vs. depth relation<br />

is illustrated in Figure 12. T max <strong>and</strong> % R o data<br />

suggest that early to peakoil generation from <strong>the</strong><br />

Draupne shale occurs between an R o equivalent of<br />

0.62% <strong>and</strong> 0.88% corresponding to depths of 3400–<br />

4400 m below sea floor. Late oil generation occurs at<br />

R o 1.1%, corresponding to a depth of approximately<br />

5000 m below sea floor.<br />

SECONDARY MIGRATION OF<br />

OIL AND GAS<br />

Evidence for Migration into Tertiary Strata<br />

Proof of <strong>hydrocarbon</strong> <strong>migration</strong> from Jurassic source<br />

<strong>rock</strong>s to Tertiary reservoir <strong>rock</strong>s is given by <strong>the</strong> accu-


mulations at <strong>the</strong> Balder, Hanna, Grane, Jotun, <strong>and</strong><br />

Ringhorne fields. In o<strong>the</strong>r areas, evidence for <strong>migration</strong><br />

into <strong>the</strong> Tertiary <strong>and</strong> Cretaceous sections is shown by<br />

<strong>the</strong> presence of oil shows <strong>and</strong> stains or <strong>the</strong>rmogenic gas<br />

<strong>within</strong> an o<strong>the</strong>rwise <strong>the</strong>rmally immature section. Core<br />

<strong>and</strong> cuttings samples were analyzed for <strong>the</strong> presence of<br />

inclusions with possible <strong>hydrocarbon</strong>s. Molecular characterization<br />

of such <strong>hydrocarbon</strong> inclusions can provide<br />

important insights to paleo<strong>migration</strong> events (Isaksen<br />

et al., 1998b). In this study, no fluid inclusions<br />

were found (Table 3). The absence of fluid inclusions<br />

is due to paleotemperatures <strong>and</strong> present-day temperatures<br />

cooler than required for <strong>the</strong> formation of inclusions.<br />

One such process, quartz overgrowth, does not<br />

start until temperatures of approximately 80C are<br />

reached.<br />

Thermogenic gases were discovered in <strong>the</strong><br />

Paleocene–Eocene of wells 16/1–1 (Figure 13), 16/1–<br />

Figure 8. Evaluation of source<br />

<strong>rock</strong> <strong>the</strong>rmal maturity <strong>and</strong> <strong>quality</strong>.<br />

Roman numerals denote <strong>the</strong><br />

predominant organic-matter<br />

type, whereas <strong>the</strong>rmal maturity<br />

levels are given by Rock-Eval<br />

T max values <strong>and</strong> vitrinite reflectance<br />

(% R o).<br />

2 (Figure 14), <strong>and</strong> in <strong>the</strong> <strong>the</strong>rmally immature Upper<br />

Jurassic sections of wells 16/1–2, 16/1–3 (Figure 15),<br />

<strong>and</strong> 25/8–2 (Figure 16). Thermogenic <strong>hydrocarbon</strong> gas<br />

is also present <strong>within</strong> <strong>the</strong> Upper Cretaceous to Upper<br />

Jurassic section (2200–3200 m) in well 25/10–2RE<br />

<strong>and</strong> is especially pronounced <strong>within</strong> <strong>the</strong> Rotliegende<br />

s<strong>and</strong>stones (Permian) at 2900–3000 m (Figure 17).<br />

Three <strong>rock</strong>samples from <strong>the</strong> Permian section (one<br />

from 3007 m <strong>within</strong> <strong>the</strong> Kupferschiefer shale <strong>and</strong> two<br />

<strong>within</strong> <strong>the</strong> Rotliegende conglomerates at 3009 <strong>and</strong><br />

3037 m) were analyzed by <strong>the</strong>rmal evaporation–gas<br />

chromatography) (Figure 18), pyrolysis–gas chromatography,<br />

solvent extraction, liquid chromatography,<br />

<strong>and</strong> gas chromatography of <strong>the</strong> saturate <strong>hydrocarbon</strong><br />

fraction (Figure 19) to checkfor presence of<br />

free <strong>hydrocarbon</strong>s <strong>and</strong> to ascertain whe<strong>the</strong>r a secondary<br />

<strong>migration</strong> pathway may exist <strong>within</strong> <strong>the</strong> Rotliegende<br />

at this location. The <strong>the</strong>rmal evaporation–gas<br />

Isaksen <strong>and</strong>Ledje 873


Figure 9. Classification of kerogen <strong>quality</strong> <strong>within</strong> <strong>the</strong> (A) upper Draupne <strong>and</strong> (B) lower Draupne formations based on <strong>the</strong> HI data<br />

shown in Figure 7. Note how areas that have higher predicted contents of type III kerogen conform to areas affected most by<br />

syntectonic mass-flow sediments along <strong>the</strong> western graben margin.<br />

chromatography data do not show significant quantities<br />

of free <strong>hydrocarbon</strong>s. Solvent extractable (free) <strong>hydrocarbon</strong>s<br />

observed <strong>within</strong> <strong>the</strong> conglomerates are<br />

likely <strong>the</strong> results of short-distance <strong>migration</strong> (2–30 m)<br />

from <strong>the</strong> Kupferschiefer organic-rich shales.<br />

Migration Model<br />

Marine source <strong>rock</strong>s of <strong>the</strong> Upper Jurassic Draupne<br />

Formation thicken from <strong>the</strong> Utsira High westward into<br />

<strong>the</strong> Viking Graben (Figure 3). Hydrocarbons are likely<br />

to have been generated in large amounts west of <strong>the</strong><br />

Utsira High. Maturation studies indicate that <strong>hydrocarbon</strong><br />

expulsion <strong>and</strong> <strong>migration</strong> started during <strong>the</strong> latest<br />

Cretaceous <strong>and</strong> continues to <strong>the</strong> present day<br />

(Thomas et al., 1985; Dahl et al., 1987; Isaksen et al.,<br />

874 Utsira High Area (Norwegian North Sea)<br />

1998b). Regional <strong>migration</strong> cell mapping at <strong>the</strong> Middle<br />

Jurassic level in <strong>the</strong> Viking Graben indicates a general<br />

basin to margin pathway. These <strong>migration</strong> <strong>pathways</strong><br />

intersect faults at which point vertical leakage through<br />

<strong>the</strong> Cretaceous section is key to charging reservoir intervals<br />

in <strong>the</strong> Tertiary.<br />

Figure 3 illustrates that <strong>the</strong> Jurassic source intervals<br />

are currently mature for generation of <strong>hydrocarbon</strong>s<br />

(% R o 0.6) in <strong>the</strong> center of <strong>the</strong> graben. The<br />

proposed model for migrating <strong>hydrocarbon</strong>s into <strong>the</strong><br />

Tertiary plays in <strong>the</strong> Utsira High area involves <strong>migration</strong><br />

at <strong>the</strong> Draupne source level into <strong>the</strong> adjacent Middle<br />

Jurassic Hugin Formation <strong>and</strong> Sleipner Formation<br />

s<strong>and</strong>stones. Longer distance secondary <strong>migration</strong> involves<br />

eastward lateral <strong>migration</strong> through <strong>the</strong> Jurassic<br />

s<strong>and</strong>stone conduits. Across <strong>the</strong> relatively small faults


Figure 10. Classification of kerogen <strong>quality</strong> <strong>within</strong> <strong>the</strong> (A) upper Draupne <strong>and</strong> (B) lower Draupne formations based on <strong>the</strong> content<br />

of liptinitic, oil-prone organic matter from visual kerogen evaluations.<br />

in <strong>the</strong> Jurassic section, <strong>hydrocarbon</strong>s migrate up along<br />

<strong>the</strong> fault plane <strong>and</strong> into Jurassic conduits on <strong>the</strong> highside<br />

blockor migrate through <strong>the</strong> fault plane <strong>and</strong> into<br />

juxtaposed s<strong>and</strong>stones. In general <strong>hydrocarbon</strong>s migrate<br />

laterally in this fashion until succeeded by vertical<br />

<strong>migration</strong> up fault planes along <strong>the</strong> Utsira High westbounding<br />

fault system.<br />

Major <strong>migration</strong> entry points available were faults<br />

cut through <strong>the</strong> entire Cretaceous succession in <strong>the</strong><br />

eastern part of <strong>the</strong> fault system as it steps up toward<br />

<strong>the</strong> Utsira High. Where Jurassic carrier beds <strong>and</strong> Tertiary<br />

reservoirs are connected, <strong>hydrocarbon</strong>s continue<br />

to move updip into available independent closures or<br />

into combined stratigraphic <strong>and</strong> structural traps at <strong>the</strong><br />

Tertiary level. The <strong>migration</strong> model is supported by<br />

Tertiary <strong>hydrocarbon</strong> shows (e.g., 16/1–1 <strong>and</strong> 16/1–<br />

2) <strong>and</strong> accumulations (e.g., Balder field) located east<br />

of such <strong>migration</strong> entry points. Migration <strong>pathways</strong><br />

through low-permeability conduits (e.g., ratty s<strong>and</strong>s in<br />

<strong>the</strong> Upper Jurassic <strong>and</strong> fractured chalks in Cretaceous)<br />

are likely to be less efficient. The absence of faulting in<br />

<strong>the</strong> Cretaceous section is believed to effectively prevent<br />

<strong>migration</strong> of <strong>hydrocarbon</strong>s into tertiary reservoirs<br />

in <strong>the</strong> central part of <strong>the</strong> graben (e.g., 15/3–1 <strong>and</strong> 15/<br />

3–3).<br />

EXPLORATION SIGNIFICANCE<br />

The key observations from this study are <strong>the</strong> following:<br />

• Both upper <strong>and</strong> lower Draupne formations are effective<br />

source <strong>rock</strong>s <strong>within</strong> <strong>the</strong> greater Utsira High<br />

area.<br />

Isaksen <strong>and</strong>Ledje 875


Figure 11. Expected <strong>hydrocarbon</strong> type generated from mature (A) upper Draupne <strong>and</strong> (B) lower Draupne source <strong>rock</strong>s. The area<br />

outlines are based on TOC o <strong>and</strong> HI o calculations.<br />

• The source <strong>rock</strong><strong>quality</strong> of <strong>the</strong> Draupne Formation<br />

ranges widely from gas-prone along <strong>the</strong> western margin<br />

of <strong>the</strong> graben to oil-prone in <strong>the</strong> east.<br />

• The gas-prone organic facies is believed to be associated<br />

with terrigenous input from westerly derived<br />

mass-flow sediments, which have both diluted <strong>the</strong><br />

oil-prone organic matter <strong>and</strong> introduced a more oxic<br />

depositional environment.<br />

• Migration is considered a significant riskfactor for<br />

any Tertiary prospect west of <strong>the</strong> Utsira High mainbounding<br />

fault.<br />

The exploration significance of <strong>the</strong>se observations<br />

lies in <strong>the</strong>ir use as an aid to predict whe<strong>the</strong>r an area is<br />

likely to be dominated by oil or gas reserves. For a prospect<br />

on <strong>the</strong> Utsira High to be sheltered from gas influx<br />

876 Utsira High Area (Norwegian North Sea)<br />

it may be concluded that three requirements need to<br />

be satisfied: (1) <strong>the</strong> Draupne Formation <strong>within</strong> <strong>the</strong><br />

drainage area needs to be of an oil-prone nature, (2) <strong>the</strong><br />

oil-prone source <strong>rock</strong>must not be <strong>within</strong> <strong>the</strong> gasgeneration<br />

window, <strong>and</strong> (3) <strong>migration</strong> access to <strong>the</strong><br />

gas-prone source kitchen west of <strong>the</strong> Utsira High area<br />

needs to be denied. It follows that it is critical to underst<strong>and</strong><br />

<strong>the</strong> <strong>migration</strong> entry points into <strong>the</strong> Tertiary<br />

section. In <strong>the</strong> graben, downdip from <strong>the</strong> Utsira High,<br />

a thickCretaceous section <strong>and</strong> lackof penetrating<br />

faults prevent effective <strong>migration</strong> from <strong>the</strong> Jurassic into<br />

Tertiary traps. Exploration efforts at Tertiary levels<br />

should <strong>the</strong>refore be focused close to <strong>the</strong> Utsira High in<br />

areas that have thin Cretaceous sediments <strong>and</strong> where<br />

basin-bounding faults are present to provide <strong>the</strong> necessary<br />

<strong>migration</strong> route.


Figure 12. Regional assessment of <strong>the</strong>rmal maturity vs. depth<br />

based on measurements of vitrinite reflectance from humic<br />

coals <strong>and</strong> shales that have terrigenous macerals.<br />

CONCLUSIONS<br />

• Chemical kerogen classification <strong>and</strong> source <strong>rock</strong> pyrolysis<br />

show that both upper <strong>and</strong> lower Draupne<br />

source intervals along <strong>the</strong> western graben margin<br />

contain more terrigenous (type III <strong>and</strong> III/II) kerogen<br />

than in <strong>the</strong> eastern part of <strong>the</strong> graben. Such<br />

change in organic facies <strong>within</strong> <strong>the</strong> Draupne source<br />

<strong>rock</strong>naturally results in a higher proportion of gas<br />

generation upon maturation.<br />

• Good potential for oil generation from Draupne Formation<br />

source <strong>rock</strong>s exists along <strong>the</strong> entire western<br />

margin <strong>and</strong> nor<strong>the</strong>rn nose of <strong>the</strong> Utsira High.<br />

• Thermogenic-<strong>hydrocarbon</strong> gas was encountered in<br />

<strong>the</strong> Upper Cretaceous <strong>and</strong> Tertiary in wells 16/1–1,<br />

16/1–2, 16/1–3, 25/8–2, <strong>and</strong> 25/10–2. These gases<br />

have migrated into <strong>the</strong> Cretaceous <strong>and</strong> Tertiary sedimentary<br />

section from deeper, mature, sections of<br />

<strong>the</strong> Draupne <strong>and</strong> Hea<strong>the</strong>r formations.<br />

• Migration entry points into <strong>the</strong> Tertiary section are<br />

provided by faults cutting through <strong>the</strong> base Paleocene.<br />

Unfaulted Cretaceous sections more likely act<br />

as a barrier to <strong>migration</strong> into Tertiary reservoirs.<br />

• No <strong>hydrocarbon</strong> fluid inclusions were detected in<br />

<strong>the</strong> samples analyzed. The lackof fluid inclusions is<br />

most likely due to <strong>the</strong> low <strong>the</strong>rmal maturity level of<br />

<strong>the</strong> Tertiary sediments. At most locations <strong>the</strong> Paleocene<br />

<strong>and</strong> Eocene s<strong>and</strong>s are poorly consolidated <strong>and</strong><br />

have not been exposed to <strong>the</strong> diagenetic temperatures<br />

required ( 80C) for mineral mobilization<br />

<strong>and</strong> <strong>the</strong> formation of fluid inclusions.<br />

• Thermogenic liquid <strong>hydrocarbon</strong>s <strong>within</strong> <strong>the</strong> Permian<br />

conglomerates in well 25/10–2 are derived<br />

from <strong>the</strong> organic-rich Kupferschiefer shales by shortdistance<br />

<strong>migration</strong>.<br />

APPENDIX: ANALYTICAL METHODS<br />

Total organic carbon (TOC) <strong>and</strong> Rock-Eval pyrolysis measurements<br />

were made on a Leco IR 312 carbon analyzer <strong>and</strong> Delsi RockEval<br />

II, respectively. Vitrinite reflectance <strong>and</strong> visual kerogen analyses were<br />

analyzed according to st<strong>and</strong>ard coal methods. Thermoevaporation<br />

<strong>and</strong> pyrolysis–gas chromatography was performed using custombuilt<br />

systems based on <strong>the</strong> HP 5790 gas chromatograph (GC)<br />

equipped with flame-ionization detectors. The <strong>the</strong>rmal extraction<br />

was done iso<strong>the</strong>rmally at 280C for 3 min. The sample was <strong>the</strong>n<br />

heated at 60C/min to 610C, <strong>and</strong> <strong>the</strong> volatiles were trapped in liquid<br />

nitrogen on <strong>the</strong> head of <strong>the</strong> GC column (split 60:1). The GC<br />

had a 50 m capillary column (5% phenyl-methylsilicone). Initial temperature<br />

was 20C followed by a heating rate of 4C/min to a final<br />

temperature of 280C. Cuttings gas, <strong>the</strong>rmal evaporation–gas chromatography,<br />

<strong>and</strong> pyrolysis analyses were carried out at Exxon Production<br />

Research Company, Houston, Texas.<br />

Isaksen <strong>and</strong>Ledje 877


Figure 13. Well 16/1–1 profiles<br />

of cuttings-gas data showing<br />

gas-wetness <strong>and</strong> summed<br />

concentrations of C 1 through C 4<br />

(vol. %). Gas wetness is calculated<br />

as 100 [(C 2 C 3 C 4)/<br />

(C 1 C 2 C 3 C 4)]. Thermogenic<br />

<strong>hydrocarbon</strong> gas is<br />

present at 2000–3000 m, <strong>within</strong><br />

<strong>the</strong> Paleocene–Eocene. No free<br />

liquid <strong>hydrocarbon</strong>s were detected<br />

(samples were analyzed<br />

from 1920, 1935, 2530, 2667,<br />

<strong>and</strong> 2743 m [Table 4]), nor<br />

were any fluid inclusions detected<br />

(samples were analyzed<br />

from <strong>the</strong> Paleocene at depths<br />

of 2320–2648 m [Table 3]).<br />

Levels of organic metamorphism<br />

(LOM) 6 <strong>and</strong> 8 correlate<br />

to vitrinite reflectance values of<br />

0.43% <strong>and</strong> 0.55%, respectively.<br />

Figure 14. Well 16/1–2 profiles<br />

of cuttings-gas data showing<br />

gas wetness <strong>and</strong> summed<br />

concentrations of C 1 through<br />

C 4. Thermogenic <strong>hydrocarbon</strong><br />

gas is present <strong>within</strong> Eocene,<br />

Paleocene, Danian, <strong>and</strong> Late Jurassic<br />

at 1700–2500 m. No free<br />

<strong>hydrocarbon</strong> liquids were detected<br />

(samples were analyzed<br />

from 2099, 2130, <strong>and</strong> 2250 m),<br />

nor were any fluid inclusions<br />

observed (analyses <strong>within</strong> <strong>the</strong><br />

Paleocene at 2102–2490 m).<br />

878 Utsira High Area (Norwegian North Sea)


Figure 15. Well 16/1–3 profiles of cuttings-gas data showing<br />

gas wetness <strong>and</strong> summed concentrations of C 1 through C 4. Thermogenic<br />

<strong>hydrocarbon</strong> gas is present <strong>within</strong> <strong>the</strong> Upper Jurassic<br />

interval at 2600–2800 m.<br />

Figure 16. Well 25/8–2 profiles of cuttings-gas data showing<br />

gas wetness <strong>and</strong> summed concentrations of C 1 through C 4. Thermogenic<br />

gas is present with <strong>the</strong> immature Upper Jurassic from<br />

1500 m to total depth drilled.<br />

Figure 17. Well 25/10–2 profiles<br />

of cuttings-gas data showing<br />

gas wetness <strong>and</strong> summed<br />

concentrations of C 1<br />

through C 4.


Figure 18. Well 25/10–2RE.<br />

Thermoevaporation (S 1) of core<br />

samples from (A) Kupferschiefer<br />

shale at 3007 m (sample<br />

202087A), (B) Rotliegende conglomerate<br />

at 3009 m (sample<br />

202087B), <strong>and</strong> (C) Rotliegende<br />

conglomerate at 3037 m (sample<br />

202087C). S 1 represents <strong>the</strong><br />

free <strong>hydrocarbon</strong>s in <strong>the</strong> <strong>rock</strong>.<br />

880 Utsira High Area (Norwegian North Sea)


Figure 19. Well 25/10–2RE.<br />

Gas chromatograms of <strong>the</strong> saturate<br />

<strong>hydrocarbon</strong> fractions<br />

from (A) Kupferschiefer shale at<br />

3007 m (sample 202087A), (B)<br />

Rotliegende conglomerate at<br />

3009 m (sample 202087B), <strong>and</strong><br />

(C) Rotliegende conglomerate<br />

at 3037 m (sample 202087C).<br />

Isaksen <strong>and</strong>Ledje 881


REFERENCES CITED<br />

Barnard, P. C., <strong>and</strong> B. S. Cooper, 1981, Oils <strong>and</strong> source <strong>rock</strong>s of<br />

<strong>the</strong> North Sea area, in L. V. Illing <strong>and</strong> G. D. Hobson, eds.,<br />

Petroleum geology of <strong>the</strong> Continental Shelf of northwest<br />

Europe: Heyden, London, Institute of Petroleum, p. 169–<br />

175.<br />

Bernard, P. C., A. G., Collins, <strong>and</strong> B. S. Cooper, 1981, Identification<br />

<strong>and</strong> distribution of kerogen facies in a source <strong>rock</strong> horizon—examples<br />

from <strong>the</strong> North Sea basin, in J. Brooks, ed.,<br />

Organic maturation studies <strong>and</strong> fossil fuel exploration: London,<br />

Academic Press, p. 271–282.<br />

Cooles, G. P., A. S. Mackenzie, <strong>and</strong> T. M. Quigley, 1986, Calculation<br />

of petroleum masses generated <strong>and</strong> expelled from source<br />

<strong>rock</strong>s, in D. Leythaeuser <strong>and</strong> J. Rullkötter, eds., Advances in<br />

organic geochemistry: Oxford, Pergamon, p. 235–245.<br />

Cooper, B. S., <strong>and</strong> P. C. Barnard, 1984, <strong>Source</strong> <strong>rock</strong>s <strong>and</strong> oils of<br />

<strong>the</strong> central <strong>and</strong> nor<strong>the</strong>rn North Sea, in G. Demaison <strong>and</strong> R. J.<br />

Murris, eds., Petroleum geochemistry <strong>and</strong> basin evaluation:<br />

AAPG Memoir 35, p. 303–314.<br />

Cooper, B. S., P. C. Bernard, <strong>and</strong> N. Telnaes, 1995, The Kimmeridge<br />

Clay formation of <strong>the</strong> North Sea, in B. J. Katz, ed., Petroleum<br />

source <strong>rock</strong>s: Berlin, Springer-Verlag, p. 89–110.<br />

Cornford, C., 1984, <strong>Source</strong> <strong>rock</strong>s <strong>and</strong> <strong>hydrocarbon</strong>s of <strong>the</strong> North<br />

Sea, in K. W. Glennie, ed., Introduction to <strong>the</strong> petroleum geology<br />

of <strong>the</strong> North Sea: London, Blackwell, p. 171–204.<br />

Curtin, D. P., <strong>and</strong> S. Ballestad, 1986, South Viking Graben: habitat<br />

of Mesozoic <strong>hydrocarbon</strong>s, in A. M. Spencer, E. Holter, C. J.<br />

Campbell, S. H. Hanslien, P. H. H. Nelson, E. Nysae<strong>the</strong>r, <strong>and</strong><br />

E. G. Ormaasen, eds., Habitat of <strong>hydrocarbon</strong>s on <strong>the</strong> Norwegian<br />

Continental Shelf: London, Norwegian Petroleum Society,<br />

Graham & Trotman, p. 153—157.<br />

Dahl, B., E. Nysae<strong>the</strong>r, G. C. Speers, <strong>and</strong> A. Yukler, 1987, Oseberg<br />

area—integrated basin modeling, in J. Brooks <strong>and</strong> K. Glennie,<br />

eds., Petroleum Geology of north west Europe: London, Graham<br />

<strong>and</strong> Trotman, p. 1029–1038.<br />

Dungworth, G., 1972, Organic geochemistry of <strong>the</strong> Marl Slate <strong>and</strong><br />

o<strong>the</strong>r organic-rich sediments: Ph.D. dissertation, University of<br />

Newcastle upon Tyne, United Kingdom, 502 p.<br />

Field, J. D., 1985, Organic geochemistry in exploration of <strong>the</strong><br />

nor<strong>the</strong>rn North Sea, in B. M. Thomas, S. S. Eggen, R. M.<br />

Larsen, P. C. Home, <strong>and</strong> A. G. Doré, eds., Petroleum<br />

geochemistry in exploration of <strong>the</strong> Norwegian Shelf: London,<br />

Norwegian Petroleum Society, Graham & Trotman, p. 39–<br />

57.<br />

Gibbons, M. J., 1978, The geochemistry of Sabkha <strong>and</strong> related deposits:<br />

Ph.D. dissertation, University of Newcastle upon Tyne,<br />

United Kingdom, 430 p.<br />

Gibbons, M. J., 1987, The depositional environment <strong>and</strong> petroleum<br />

geochemistry of <strong>the</strong> Marl Slate-Kupferschiefer, in J. Brooks<br />

<strong>and</strong> A. J. Fleet, eds., Marine petroleum source <strong>rock</strong>s: Geological<br />

Society Special Publication 26, p. 249.<br />

Goff, J. C., 1983, Hydrocarbon generation <strong>and</strong> <strong>migration</strong> from Jurassic<br />

source <strong>rock</strong>s in <strong>the</strong> East Shetl<strong>and</strong> Basin <strong>and</strong> Viking Graben<br />

of <strong>the</strong> nor<strong>the</strong>rn North Sea: Journal of <strong>the</strong> Geological Society,<br />

v. 140, p. 445–474.<br />

Harris, J. P., <strong>and</strong> R. M. Fowler, 1987, Enhanced prospectivity of<br />

<strong>the</strong> Mid-Late Jurassic sediments of <strong>the</strong> South Viking Graben,<br />

nor<strong>the</strong>rn North Sea, in J. Brooks <strong>and</strong> K. Glennie, eds., Petroleum<br />

geology of north west Europe: London, Graham & Trotman,<br />

p. 879–898.<br />

Heinrichs, S. M. <strong>and</strong> W. S. Reeburgh, 1987, Anaerobic mineralization<br />

of marine sediment organic matter: rates <strong>and</strong> <strong>the</strong> role<br />

of anaerobic processes in <strong>the</strong> ocean carbon economy: Geomicrobiology<br />

Journal, v. 5, p. 191–237.<br />

882 Utsira High Area (Norwegian North Sea)<br />

Hirst, D. M., <strong>and</strong> K. C. Dunham, 1963, Chemistry <strong>and</strong> petrography<br />

of <strong>the</strong> Marl Slate of SE Durham, Engl<strong>and</strong>: Economic Geology,<br />

v. 58, p. 912–940.<br />

Huc, A. Y., H. Irwin, <strong>and</strong> M. Schoell, 1985, Organic matter <strong>quality</strong><br />

changes in an Upper Jurassic shale sequence from <strong>the</strong> Viking<br />

Graben, in B. M. Thomas, S. S. Eggen, R. M. Larsen, P. C.<br />

Home, <strong>and</strong> A. G. Doré, eds., Petroleum geochemistry in exploration<br />

of <strong>the</strong> Norwegian Shelf: London, Norwegian Petroleum<br />

Society, Graham & Trotman, p. 179–183.<br />

Isaksen, G. H., <strong>and</strong> K. M. Bohacs, 1995, Geological controls of<br />

source <strong>rock</strong>geochemistry through relative sea-level; Triassic;<br />

Barents Sea, in B. Katz, ed., Petroleum source <strong>rock</strong>s: New<br />

York, Springer-Verlag, p. 25–50.<br />

Isaksen, G. H., R. Patience, G. van Graas, <strong>and</strong> A. I. Jenssen,<br />

1997, Elements of <strong>hydrocarbon</strong> system analysis in a rift-basin<br />

with mixed marine <strong>and</strong> non-marine source <strong>rock</strong>s: <strong>the</strong> South<br />

Viking Graben, North Sea (abs.): International Meeting on<br />

Organic Geochemistry, extended abstracts volume, p. 575–<br />

576.<br />

Isaksen, G. H., D. J. Curry, J. D. Yeakel, <strong>and</strong> A. I. Jenssen, 1998a,<br />

Controls on <strong>the</strong> oil <strong>and</strong> gas potential of humic coals: Organic<br />

Geochemistry, v. 29, no. 1–3, p. 23–44.<br />

Isaksen, G. H., R. J. Pottorf, <strong>and</strong> A. I. Jenssen, 1998b, Correlation<br />

of fluid inclusions <strong>and</strong> reservoired oils to infer trap fill history<br />

in <strong>the</strong> South Viking Graben, North Sea: Petroleum Geoscience,<br />

v. 4, p. 41–55.<br />

Lee, C., 1992, Controls on organic carbon preservation: <strong>the</strong> use of<br />

stratified water bodies to compare intrinsic rates of decomposition<br />

in oxic <strong>and</strong> anoxic systems: Geochimica et Cosmochimica<br />

Acta, v. 56, p. 3323–3335.<br />

Mayer, L. M., 1994, Surface area control of organic carbon accumulation<br />

in continental shelf sediments: Geochimica et Cosmochimica<br />

Acta, v. 58, p. 1271–1284.<br />

Miller, R. G., 1990, A palaeoceanographic approach to Kimmeridge<br />

Clay formation, in A. Y. Huc, ed., Deposition of organic facies:<br />

AAPG Studies in Geology 30, p. 13–26.<br />

Oszczepalski, S., 1986, On <strong>the</strong> Zechstein Copper Shale lithofacies<br />

<strong>and</strong> palaeoenvironments in SW Pol<strong>and</strong>, in G. M. Harwood <strong>and</strong><br />

D. B. Smith, eds., The English Zechstein <strong>and</strong> related topics:<br />

Oxford, Blackwell, p. 171–182.<br />

Passey, Q. R., S. Creaney, J. B. Kulla, F. J. Moretti, <strong>and</strong> J. D.<br />

Stroud, 1990, A practical model for organic richness from<br />

porosity <strong>and</strong> resistivity logs: AAPG Bulletin, v. 74, p. 1777–<br />

1794.<br />

Paul, J., 1986a, Stratigraphy of <strong>the</strong> lower Werra Cycle (Zl) in West<br />

Germany, in G. M. Harwood <strong>and</strong> D. B. Smith, eds., The English<br />

Zechstein <strong>and</strong> related topics: Oxford, Blackwell, p. 149–<br />

155.<br />

Paul, J., 1986b, Environmental analysis of basin <strong>and</strong> schwellen facies<br />

in <strong>the</strong> lower Zechstein of Germany, in G. M. Harwood, <strong>and</strong><br />

D. B. Smith, eds., The English Zechstein <strong>and</strong> related topics:<br />

Oxford, Blackwell, p. 143–147.<br />

Pepper, A., 1991, Estimating <strong>the</strong> petroleum expulsion behavior of<br />

source <strong>rock</strong>s: a novel quantitative approach, in W. A. Engl<strong>and</strong><br />

<strong>and</strong> A. J. Fleet, eds., Petroleum <strong>migration</strong>: Geological Society<br />

Special Publication 59, p. 9–31.<br />

Skarpnes, O., G. P. Hamar, K. H. Jakobsson, <strong>and</strong> D. E. Ormaasen,<br />

1981, Regional Jurassic setting of <strong>the</strong> North Sea north of <strong>the</strong><br />

central highs: <strong>the</strong> sedimentation of <strong>the</strong> North Sea reservoir<br />

<strong>rock</strong>s: Norwegian Petroleum Society, article 13, 18 p.<br />

Smith, D. B., 1980, The evolution of <strong>the</strong> English Zechstein basin:<br />

Contributions in Sedimentology, v. 9, p. 7–34.<br />

Thomas, B. M., P. Moller-Pedersen, M. F. Whitaker, <strong>and</strong> N. D.<br />

Shaw, 1985, Organic facies <strong>and</strong> <strong>hydrocarbon</strong> distributions in<br />

<strong>the</strong> Norwegian North Sea, in B. M. Thomas, S. S. Eggen, R. M.<br />

Larsen, P. C. Home, <strong>and</strong> A. G. Doré, eds., Petroleum geo-


chemistry in exploration of <strong>the</strong> Norwegian Shelf: London,<br />

Norwegian Petroleum Society, Graham & Trotman, p. 3–<br />

26.<br />

Ungerer, P., A. Chiarelli, <strong>and</strong> J. L. Oudin, 1985, Modelling of petroleum<br />

genesis <strong>and</strong> <strong>migration</strong> with a bidimensional computer<br />

model in <strong>the</strong> Frigg sector, Viking Graben, in B. M. Thomas,<br />

S. S. Eggen, R. M. Larsen, P. C. Home, <strong>and</strong> A. G. Doré, eds.,<br />

Petroleum geochemistry in exploration of <strong>the</strong> Norwegian<br />

Shelf: London, Norwegian Petroleum Society, Graham &<br />

Trotman, p. 121–129.<br />

Ziegler, P. A., 1988, Evolution of <strong>the</strong> Arctic-Atlantic <strong>and</strong> <strong>the</strong> western<br />

Tethys: AAPG Memoir 43, 198 p.<br />

Isaksen <strong>and</strong>Ledje 883

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!