You are on page 1of 26

Oil families and their inferred source rocks in the Barents Sea and northern Timan-Pechora Basin, Russia

Meng He, J. Michael Moldowan, Alla Nemchenko-Rovenskaya, and Kenneth E. Peters

AUTHORS Meng He  Stanford University, Department of Geological and Environmental Sciences, Stanford, California; hmenglxq@stanford.edu Meng He received his B.S. degree in petroleum geology and geochemistry from Moscow State University in 2005, and currently, he is a fourthyear Ph.D. student in the Department of Geological and Environmental Sciences at Stanford University. His broad research interests include the application of organic geochemistry to petroleum research. He is a student member of the Basin and Petroleum System Modeling (BPSM) and Molecular Organic Geochemistry Industrial Affiliates (MOGIA) at Stanford. Specifically, he is interested in the application of organic geochemistry as a means to understand the formation of sedimentary basins, degradation products under geochemical conditions within petroleum formation, and the origin of organic components in both sediments and petroleum. His research focuses on the applications of new technologies of organic geochemistry to three-dimensional basin and petroleum systems modeling at Stanford. He has extensive experience with the oil industry from his internships with Rosneft, YUKOS, ConocoPhillips, ExxonMobil, StatoilHydro ASA, PetroChina, and BP in the past 6 yr. J. Michael Moldowan  Stanford University, Department of Geological and Environmental Sciences, Stanford, California; Moldowan@stanoford.edu J. Michael Moldowan has been a research professor at Stanford University since 1993. Before he joined Stanford, he worked in the Chevron biomarker team since 1974. He founded the Biomarker Technology company in 1993, dedicated to advanced geochemical methods application in exploration and production. He published more than 100 refereed articles including The Biomarker Guides, first and second editions, and was a Treibs Medalist of the Organic Geochemistry Division of the Geochemical Society in 2011. Alla Nemchenko-Rovenskaya  The Foundation for East-West Cooperation, Moscow, 119234, Russia; tnemchenko@mail.ru Alla Nemchenko-Rovenskaya is a leading researcher at V.I. Vernadsky Institute of Geochemistry and Analytical Chemistry, Moscow, Russia. She has 50 yr of experience in scientific and academic institutes. She is an author of more than 100 publications, and the author of the scientific discovery

ABSTRACT A geochemical study of 34 oil samples was conducted to understand the types and distributions of effective source rocks and evaluate the geographic extent of the petroleum systems in the Barents Sea and northern Timan-Pechora Basin. Taxonspecific, age-related, and source-related biological markers (biomarkers) and isotope data provided information on the depositional environment of the source rock, source input, and source age of the oil samples. A relationship between biomarker and diamondoid concentration was used to identify mixed oils having both oil window and highly cracked components. Compound-specific isotope analyses of diamondoids and n-alkanes were used to deconvolute cosourced oils and identify deep source rocks in the basin. Results suggest five major source rocks in the Barents Sea and the northern TimanPechora Basin: Upper Jurassic shale, LowerMiddle Jurassic shale, Triassic carbonate and shale, Devonian marl, and Devonian carbonate. The Upper and LowerMiddle Jurassic source rocks are dominant in the Barents Sea. Triassic source rock consists of carbonate in the onshore part of northern TimanPechora Basin and marine shale in the Barents Sea. The Devonian Domanik Formation carbonate source rock extends offshore into the southern Barents Sea. The high-maturity Domanik Formation could also be a secondary source rock for most of the mixed oils in the northern Timan-Pechora Basin. This detailed geochemical study provides a new and detailed understanding of petroleum systems in the Barents Sea and northern Timan-Pechora Basin.

Copyright 2012. The American Association of Petroleum Geologists. All rights reserved. Manuscript received March 28, 2011; provisional acceptance June 16, 2011; revised manuscript received October 2, 2011; final acceptance October 18, 2011. DOI:10.1306/10181111043

AAPG Bulletin, v. 96, no. 6 (June 2012), pp. 1121 1146

1121

of the regular connection between the development of coal-bearing and the formation of giant gas fields in northern western Siberia. Alla Nemchenko is a senior scientist at the Russian Academy of Sciences V.I. Vernadsky Institute of Geochemistry and Analytical Chemistry, Moscow, Russia. She has 15 yr of experience in the industry and the academy, YUKOS, Exxon oil companies, and University of Texas at Austin, Faculty of Oil Economy. Kenneth E. Peters  Stanford University, Department of Geological and Environmental Sciences, Stanford, California; present address: Schlumberger, 18 Manzanita Place, Mill Valley, California; kepeters@stanford.edu Kenneth E. Peters is a science adviser for Schlumberger, having 33 yr of experience in the industry, government, and academia. He is a principal author of The Biomarker Guide (2005), a honorary teaching fellow at the University of Aberdeen, and a consulting professor at Stanford University where he co-leads the Basin and Petroleum System Modeling Industrial Affiliates Program. He was chairperson of the AAPG Research Committee (20072010) and an AAPG Distinguished Lecturer (20092010). In 2009, he received the Schlumberger Henri Doll Prize for Innovation and the Alfred E. Treibs Award presented by the Geochemical Society. Ken has a Ph.D. in geochemistry from University of California, Los Angeles.

INTRODUCTION The former Soviet Union is the largest producer of petroleum in the world, and it still has considerable potential for oil and gas exploration. The Timan-Pechora Basin has been a prolific producer of oil, condensate, and gas in the Arctic coastal region of northern Russia (Figure 1A, B). The well-established hydrocarbon trend onshore suggests an offshore extension that remains little studied and relatively unexplored. The Barents Sea includes the north Norwegian and Russian coasts, Novaya Zemlya, Franz Joseph Land, Svalbard archipelago, and the eastern margin of the Atlantic Ocean, an area covering about 1.25 million km2 (Figure 1A). Petroleum geologists and geochemists are becoming more interested in the Barents Sea Basin (especially the Russian Barents Sea) because of the increasing value of the petroleum and the great potential for oil and gas in this area. The literature from studies of this area is limited. Because the former Soviet Union only recently opened to the West, little information is available about the petroleum geochemistry of most Russian basins. Better evaluation of the prospects in this basin requires understanding of the types and distributions of effective source rocks (Peters and Cassa, 1994) and the geographic extent of the petroleum systems (Magoon and Dow, 1994). According to Dalland et al. (1988), Larsen et al. (1993), Ostisty and Cheredeev (1993), Bjory et al. (1978), Johansen et al. (1993), and Ulmishek (1982), numerous potential source rocks are in the Barents Sea region. These include marine Devonian, PennsylvanianPermian, Triassic, and Jurassic strata. However, most publications on petroleum systems focus on the Norwegian Barents Sea instead of the Russian Barents Sea. The previous studies do not fully consider or document deep source rocks and mixed oils composed of high- and lowmaturity contributions from multiple source rocks or from a single source rock. We conducted biological marker (biomarker) and diamondoid studies in the Barents Sea and northern TimanPechora Basin. In the biomarker study, we performed oil-oil correlation to identify oil families having one source rock using age- and source-related biomarkers. In addition, we also recognized mixed oils from multiple source rocks based on biomarker data and compound-specific isotope analysis of nalkanes. In the diamondoid study, we identified the mixed oil samples that contain oil window maturity oil and highly cracked oil components. Compound-specific isotope analysis of diamondoids (CSIA-D) is used to identify the deep source contributions to those mixed oil samples. Thirty-four

ACKNOWLEDGEMENTS This work was completed with financial support from the Stanford University Molecular Organic Geochemistry Industrial Affiliates (MOGIA) Program and A. I. Levorsen Research Grant. We greatly appreciate the assistance from staff in the sample separation, GC, GC-MS, and GC-MS/MS in the MOG Laboratory at Stanford University. We are extremely grateful to Robert Dunbar and Dave Mucciarone who provided us the opportunity to complete stable isotope analyses in the Stable Isotope Biogeochemistry Laboratory at Stanford University. We also thank GB Scientific Inc. for the compoundspecific isotope analyses of diamondoids and n-alkanes for all of the oil samples in this study. The AAPG Editor thanks the following reviewers for their work on this paper: Bret J. Fossum, Eric E. Michael, and Clifford C. Walters. EDITORS NOTE A color version of the figures may be seen in the online version of this paper.

1122

Barents Sea and Northern Timan-Pechora

oil samples from the Russian Barents Sea and northern Timan-Pechora Basin (Figure 1A, B; Table 1) were analyzed in the Molecular Organic Geochemistry Laboratory at Stanford University. Our interpretations assume that all field samples from the study area were properly labeled as to origin, and we found no evidence to the contrary. Source rock samples are not found in this study. All of the fluid samples have different physical properties, but we refer to them simply as oils. Our data provide an excellent opportunity to evaluate the hydrocarbon resource potential of the Barents Sea and northern Timan-Pechora Basin. The results of this study provide a better understanding of the petroleum systems in this region. In addition, this study identifies new petroleum systems related to deeply buried source rocks. Improved understanding of such source rocks can provide new exploration targets for oil, condensate, and gas in this frontier petroleum province. Such new insights may also be crucial for making exploration decisions both within the province and in neighboring blocks as they become available for exploration. Finally, we show that some of the nonroutine molecular geochemical methods used in this study can open new vistas for oil and gas exploration in any basin.

METHODS Compared with most bulk geochemical parameters, source-related and age-related biomarker ratios for oil samples are particularly useful to describe facies and depositional environment of the source rock while also successfully establishing oiloil and oil-source correlations. Biomarkers in oil samples can be applied to infer depositional environment, organic facies, thermal maturity, and even the age of the source rock (Peters et al., 2005). These data can be extremely helpful in oil-oil correlation studies. Oil cracking is a process where complex hydrocarbons are cracked to smaller molecules by heating (Dahl et al.,1999; Schoell and Carlson, 1999). We use a method that uses quantitative

analysis of both biomarkers and diamondoids (Dahl et al.,1999) to better understand oil cracking. The extent of thermal alteration can be assessed by quantitative analysis of diamondoids, which have a diamond structure that displays high thermal stability. During oil cracking, the concentration of diamondoids in oil increases. In contrast, most biomarkers have relatively low thermal stability and their abundance decreases during cracking. Biomarkers are also useful to describe the extent of oil cracking. The C29 aaa20R sterane (stigmastane) was selected for analyses of oil cracking by Dahl et al. (1999) because it decreases to near the detection limit during oil cracking, whereas the concentration of certain diamondoids (3- and 4methyldiamantanes) greatly increases. Thus, mixtures of noncracked (low-maturity) and cracked (high-maturity) oils show both significant stigmastane concentration and elevated diamondoid concentrations. To increase the accuracy of biomarker analysis, a series of separations was conducted before the analysis. First, silica-gel chromatography was applied to separate oil into saturates and aromatics. Second, the zeolite ZSM-5 was used to remove normal alkanes (n-alkanes or paraffins) from the saturate fraction. Before removing normal alkanes, saturate fractions were spiked with internal standards. For this project, a mixture of 5b-cholane and highly branched isoprenoids having an M/Z 238 mass spectral fragment was added to the saturates as an internal standard before gas chromatographymass spectrometry (GC/MS) (Seifert and Moldowan, 1979). Saturate fractions were treated with high Si/Al Z5M-5 zeolite (silicalite) to remove n-alkanes. All of the crude oils were quantitatively analyzed by gas chromatographyflame ionization detection (GC-FID). The biomarkers (terpanes) and diamondoids in the branched, cyclic, and aromatic fractions were analyzed by selected ion recording gas chromatographymass spectrometry (SIR-GC/ MS). Quantitative metastable reaction monitoring GC/MS (MRM-GC/MS) was used to analyze steranes (e.g., regular steranes, such as cholestanes, ergostanes and stigmastanes, diasteranes, dinosteranes), as well as certain terpanes (e.g., bicandinanes, tetracyclic polyprenoids, diterpanes) (Moldowan
He et al. 1123

Figure 1. (A) Location map for Barents Sea and Timan-Pechora Basin shows identified oil families. (B) Location map for Timan-Pechora Basin shows identified oil families. Structural regions from Ulmishek (1982) and Lindquist (1999b).

et al., 2004). Stable carbon isotope analyses of crude oils and saturate and aromatic fractions (Table 2) were completed in the Stable Isotope Biogeochemistry Laboratory at Stanford University. Analysis of diamondoids was conducted using the saturate fraction of each oil sample. A deuterated diamondoid internal standard containing D4 adamantane, D3 1-methyladamantane, D4 diamantane, D3 methyldiamantane, D5 ethyldiamantane, cholane, and D4 triamantane was added to each oil sample before the silica-gel chromatography separation. The SIR-GC/MS analysis of saturate fractions was conducted to measure the distribution of diamondoids in each saturate fraction. Adamantanes and diamantanes were quantitatively
1124 Barents Sea and Northern Timan-Pechora

analyzed by monitoring M/Z 135, 136, 149, and 187, 188, 201 mass chromatograms, respectively (Wingert, 1992). Quantification of diamondoids was achieved by integration of peak heights with reference to the standards. Compound-specific isotope analyses of diamondoids and n-alkanes were completed by GB Scientific Inc. on selected oil samples. Chemometric analysis was used to classify and assign confidence limits for identifying genetic oil families in the Barents Sea and northern TimanPechora Basin. Chemometrics is a collection of multivariate statistical methods for recognizing patterns in large data sets. It allows us to identify and remove noise from the data, show affinities

Figure 1. Continued. He et al. 1125

Table 1. Oil Samples by Field, Depth, Reservoir Age, and Oil Family
Depth (m) List 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 Sample ID 526a 52(1) 152 78 67 91 12 102(1) 22(1) 22(2) 22(3) 43 65 80(2) 81(2) 87 102(2) 50 80(1) 41 880 144 44 54 274 81(3) 81(4) 220 150 151 146 149 147 561 Field/Area Barentsburg Pechano-Ozerskoye Pechano-Ozerskoye Tarskoye Severo-Kharyaginkoye Khylchuyuskoye Vostochno-Vasilykovskoye Yareyyuskoye Zapadno-Lekkeyyaginskoye Zapadno-Lekkeyyaginskoye Zapadno-Lekkeyyaginskoye Myadseiskoye Severo-Saremboiskoye Toboiskoye Trebsa Ust-Talotinskoye Yareyyuskoye Passedskoye Toboiskoye Medynskoye Medynskoye N.Gulyaevskoye Naulskoye Prirazlomnoye Varandey Trebsa Trebsa Pechano-Ozerskoye N.Gulyaevskoye Prirazlomnoye Shtokmanovskoye Shtokmanovskoye Shtokmanovskoye Spitsbergen Region Spitsbergen Island Kolguyev Island Kolguyev Island Kolguyev Island Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Timan Pechora (northern part) Pechora Shelf Pechora Shelf Pechora Shelf Timan Pechora (northern part) Pechora Shelf Pechora Sea Timan Pechora (northern part) Timan Pechora (northern part) Kolguyev Island Pechora Shelf Pechora Shelf Barents Sea Barents Sea Barents Sea Spitsbergen Island Top Coal mine 1554.4 1538 1734 1554.4 1660 2457 1735 2560 3050 3118 Bottom hole 3056 4145 3669 3945 1990 3696.7 2640 2837 2857 2012 1240 2406 4464 3636 3855 1601 2112 2368 1812 1886 2237 N/A* Bottom Coal mine 1566.8 1561 1740 1566.8 1675 2539 1741 2571 3097 3183 Bottom hole 3110 4196 3683 3977 2000 3743 2710 2882 2882 2049 1250 2460 4490 3707 3910 1605 2249 2438 1880 1961 2257 N/A* Age of Reservoir Early Permian Early Triassic Early Triassic Early Triassic Early Triassic Early Triassic Pennsylvanian Early Permian Late Devonnian N/A* Early Devonian N/A* Early Devonian Early Devonian Late Devonian N/A* Early Permian Late Devonian Late Devonian Early Devonian Early Devonian Late Permian Late Permian Early Permian Early Devonian Late Devonian Late Devonian Early Triassic Late Permian Early Permian Middle Jurassic Middle Jurassic Early Jurassic N/A* Inferred Source Rock Triassic Triassic Triassic Triassic Triassic Triassic Triassic Triassic Devonian Marl Devonian Marl Devonian Marl Devonian Marl Devonian Marl Devonian Marl Devonian Marl Devonian Marl Devonian Carbonate Devonian Carbonate Devonian Carbonate Devonian Carbonate Devonian Carbonate Triassic/Devonian Carbonate Triassic/Devonian Carbonate Triassic/Devonian Carbonate Triassic/Devonian Carbonate Triassic/Devonian Carbonate Triassic/Devonian Carbonate Upper Jurassic Upper Jurassic Upper Jurassic Lower-Middle Jurassic Lower-Middle Jurassic Lower-Middle Jurassic Lower-Middle Jurassic Family I I I I I I I I II II II II II II II II III III III III III IV IV IV IV IV IV V V V VI VI VI VI

1126 Barents Sea and Northern Timan-Pechora

*N/A = not applicable.

among parameters, better understand the true structural relationship behind the data, and make accurate predictions about unknown samples. Two common techniques in chemometric analysis are hierarchical cluster analysis (HCA) and principal components analysis (PCA). For the chemometric analysis, we selected 20 biomarker and two isotope ratios (Table 2) based on our geochemical expertise to differentiate oil families in terms of age, organic input, and depositional environment of the source rocks. The HCA and PCA were completed using Pirouette (Infometrix, Inc.). Cluster distances between samples in HCA were calculated and visualized in a dendrogram (Figure 2A). The cluster distance is a relative measure of the degree of similarity among samples. In HCA for this study, each different parameter (biomarker or isotope ratio) has equal weight or significance. The Pirouette HCA used autoscale preprocessing, Euclidean metric distance, and incremental linkage options. The PCA reduces the dimensionality of a data set that is composed of many variables while retaining most of the information in the data set. It transfers several possibly related variables into a smaller number of uncorrelated variables, the principal components (Figure 2B). This transformation does not change the structural relationships underlying the data. The first three principal components generally describe most of the variance within the data set. Three-dimensional plots are built using the first three principal components as axes to generate the best three-dimensional separation of the samples into genetic families (Figure 2B).

PETROLEUM GEOLOGY The Barents Sea Basin was considered a foredeep basin of the Uralian tectonic belt and acted as a major catchment area for sediments shed from the Ural Mountains in the late PaleozoicMesozoic (Gramberg, 1988). With an average water depth about 350 m (1148 ft), it is one of the largest areas of continental shelf on the globe, covering about 230,000 km2 (88,800 mi2). The Barents Sea Basin formed by two major continental collisions

(400 and 240 Ma) and subsequently opened by continental separation. Because of the two collisions, the basement of this basin is dominated by Caledonian and Uralian trends, whereas extensional tectonic movements dominate its later Paleozoic and Mesozoic history. These tectonic events created the major rift basin traversing the Barents shelf and the platforms (Dore, 1995). Because of its special geographic location, this basin was strongly influenced by a marine depositional environment affected by the tectonic setting and climate changes. Carbonate deposition with some episodes of evaporite deposition prevailed over wide areas of the shelf in the late Paleozoic. Clastic deposition under more temperate conditions became dominant through the Mesozoic. The basin is characterized by an impressive thickness of approximately 12 km (7.5 mi) of Permian and younger sediments, especially Triassic deposits, which locally reach approximately 6 to 8 km (3.75 mi) in thickness (Dore, 1995). The western margin of the Barents Sea shelf was characterized by Cenozoic tectonics and sedimentation associated with the northern Atlantic continental separation. Thick Cenozoic sediments were deposited before and after the beginning of passive continental drifting in the Oligocene (Johansen et al., 1993). Sediment accumulation in this area of the basin was accelerated by uplift and erosion of the Barents Sea shelf to the east (Nyland et al., 1992). Erosion and redeposition were intense during the Pleistocene glaciations (Dore, 1995). LowerMiddle Jurassic sandstone is the major reservoir rock, having high porosity and permeability in the Barents Sea Basin (Dore, 1995). However, Triassic sandstone also exists in this area, although it has inferior reservoir quality compared with the Jurassic sandstone. Ninety-seven percent of the known recoverable reserves are from Jurassic reservoirs in the eastern Barents Sea Basin (Petroconsultants, 1996, personal communication). Thick and widespread Jurassic and Triassic shales provide good local to regional seal rocks, as do Lower Cretaceous shales (Johansen et al., 1993). Fault-bounded and domelike structural traps are dominant in the Barents Sea Basin because of the intense tectonics. On the
He et al. 1127

Table 2. Geochemical Parameters of Oil Samples in the Barents Sea and Northern Timan-Pechora Basin C29/C30 Sample C19/ ID Family C19+C23* C22/C21* C24/C23 C26/C25* Tet/C23* %C27* %C28* %C29* ETR* C31R/H* Hop* 152 102(1) 91 67 12 78 52(1) 526a 81(2) 87 22(1) 80(2) 65 22(3) 22(2) 43 50 880 41 102(2) 80(1) 274 44 81(4) 54 81(3) 144 151 220 150 561 147 146 149 I I I I I I I I II II II II II II II II III III III III III IV IV IV IV IV IV V V V VI VI VI VI 0.2 0.12 0.14 0.12 0.11 0.1 0.04 0.16 0.48 0.27 0.32 0.39 0.29 0.32 0.31 0.43 0.06 0.1 0.11 0.1 0.09 0.1 0.12 0.1 0.09 0.11 0.1 0.68 0.41 0.58 0.91 0.83 0.48 0.76 0.4 0.32 0.34 0.37 0.31 0.51 0.48 0.25 0.27 0.44 0.32 0.3 0.28 0.3 0.29 0.3 0.47 0.57 0.59 0.57 0.55 0.48 0.46 0.48 0.49 0.47 0.38 0.19 0.31 0.35 0.25 0.32 0.25 0.18 0.56 0.63 0.65 0.6 0.68 0.79 0.77 0.75 0.78 0.76 0.77 0.66 0.77 0.7 0.7 0.73 0.58 0.47 0.47 0.46 0.49 0.51 0.52 0.52 0.53 0.53 0.56 0.81 0.83 0.84 0.62 0.48 0.58 0.51 0.92 1.06 1.06 1.02 1.16 1.48 1.46 1.18 1.06 1.03 0.99 0.98 0.86 0.99 0.88 0.94 0.84 0.86 0.85 0.9 0.9 0.96 0.96 0.91 0.93 0.97 1.12 1.29 1.08 1.1 1.03 0.95 0.71 0.43 0.12 0.28 0.27 0.29 0.16 0.08 0.08 0.14 1.77 1.54 1.67 1.53 1.64 1.94 1.86 1.75 0.33 0.52 0.53 0.51 0.4 0.43 0.41 0.45 0.45 0.44 0.36 1.94 0.57 1.51 0.38 0.92 0.31 0.84 0.329 0.297 0.286 0.284 0.341 0.286 0.301 0.285 0.244 0.292 0.22 0.217 0.184 0.201 0.201 0.171 0.3 0.279 0.272 0.312 0.308 0.284 0.29 0.262 0.274 0.261 0.269 0.246 0.155 0.199 0.032 0.333 0.295 0.201 0.343 0.255 0.236 0.358 0.287 0.261 0.301 0.219 0.333 0.35 0.39 0.313 0.354 0.303 0.341 0.275 0.186 0.245 0.259 0.264 0.229 0.242 0.222 0.237 0.235 0.302 0.236 0.392 0.291 0.353 0.148 0.295 0.18 0.232 0.329 0.447 0.478 0.358 0.372 0.454 0.398 0.47 0.423 0.358 0.39 0.469 0.462 0.495 0.459 0.554 0.514 0.476 0.47 0.425 0.463 0.474 0.488 0.501 0.492 0.436 0.495 0.362 0.555 0.448 0.82 0.371 0.525 0.567 3.21 4.43 4.03 4.37 4.81 2.7 3.63 4.12 0.56 0.61 0.48 0.45 0.43 0.34 0.38 0.32 2.99 0.86 0.97 0.86 1.37 2.68 2.86 2.44 2.46 2.27 2.75 0.76 1.57 0.94 0.95 0.98 1.47 0.8 0.24 0.32 0.33 0.31 0.26 0.34 0.31 0.31 0.31 0.31 0.3 0.32 0.31 0.31 0.31 0.34 0.42 0.39 0.41 0.41 0.41 0.36 0.37 0.38 0.38 0.41 0.34 0.22 0.26 0.22 0.44 0.3 0.3 0.23 0.53 0.93 0.8 0.93 0.45 0.35 0.38 0.54 0.57 0.65 0.7 0.64 0.63 0.63 0.65 0.53 0.77 0.89 0.85 0.88 0.73 1.21 1.18 1.3 1.29 0.9 1.13 0.34 0.36 0.37 0.54 0.73 0.77 0.71 DiaH/ C35/ (H+diaH)* (C34+C35)Hop* GA/H* S/H* Bicad* 0.22 0.02 0.1 0.07 0.11 0.29 0.27 0.46 0.14 0.04 0.05 0.11 0.06 0.05 0.05 0.14 0.01 0.02 0.03 0.02 0.02 0.04 0.03 0.04 0.02 0.03 0.04 0.15 0.11 0.14 0.1 0.1 0.06 0.03 0.66 0.54 0.51 0.5 0.48 0.48 0.49 0.5 0.34 0.37 0.38 0.35 0.34 0.35 0.35 0.3 0.51 0.44 0.46 0.43 0.45 0.49 0.51 0.5 0.49 0.51 0.5 0.37 0.45 0.4 0.3 0.19 0.33 0.53 0.02 0.09 0.12 0.1 0.06 0.07 0.04 0.19 0.13 0.14 0.12 0.13 0.12 0.09 0.1 0.27 0.09 0.09 0.1 0.1 0.09 0.15 0.14 0.16 0.13 0.12 0.11 0.01 0.02 0 0.01 0.01 0.01 0.02 0.43 0.27 0.56 0.24 0.59 0.31 0.3 1.05 0.11 0.08 0.16 0.09 0.07 0.07 0.07 0.06 0.09 0.07 0.13 0.14 0.08 0.1 0.11 0.09 0.08 0.12 0.11 0.11 0.34 0.34 0.1 0.53 0.22 0.2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0.013 0.008 0.013 0.016 0.003 0 0 0

1128 Barents Sea and Northern Timan-Pechora

Table 2. Continued Sample ID 152 102(1) 91 67 12 78 52(1) 526a 81(2) 87 22(1) 80(2) 65 22(3) 22(2) 43 50 880 41 102(2) 80(1) 274 44 81(4) 54 81(3) 144 151 220 150 561 147 C30/ (C2730) diaS* 0.05 0.06 0.06 0.07 0.02 0.08 0.10 0.07 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.01 0.00 0.01 0.00 0.01 0.03 0.03 0.02 0.02 0.05 0.04 0.05 0.04 0.02 0.00 0.01 TADMD3/ C28S* 0.11 0.05 0.04 0.03 0.02 0.05 0.00 0.06 0.04 0.01 0.01 0.02 0.01 0.01 0.01 0.02 0.02 0.01 0.01 0.01 0.01 0.02 0.03 0.02 0.01 0.03 0.02 0.04 0.07 0.08 0.06 0.03 TADino* 0.55 0.11 0.17 0.12 0.20 0.18 0.44 0.43 0.23 0.10 0.12 0.08 0.06 0.07 0.07 0.07 0.10 0.07 0.04 0.08 0.24 0.07 0.04 0.05 0.06 0.57 0.07 0.60 0.72 0.71 0.59 0.69 C28/ C29 1.04 0.57 0.49 1.00 0.77 0.57 0.76 0.47 0.79 0.98 1.00 0.67 0.77 0.61 0.74 0.50 0.36 0.51 0.55 0.62 0.49 0.51 0.45 0.47 0.48 0.69 0.48 1.08 0.52 0.79 0.18 0.79 3/ (3- + 4MeS20R) 0.74 0.95 0.96 0.95 0.93 0.96 0.77 0.92 0.98 0.95 0.98 0.97 0.97 0.98 0.96 0.97 0.95 0.98 0.99 0.98 0.98 0.97 0.99 0.98 0.98 0.70 0.98 0.68 0.55 0.59 0.80 0.64 C24Tet/ C26 0.33 0.50 0.50 0.55 0.23 0.08 0.08 0.18 2.55 2.17 2.04 2.21 2.33 2.49 2.54 2.09 0.81 1.30 1.25 1.26 1.01 1.08 0.98 1.15 1.20 1.03 0.76 1.86 0.65 1.41 0.61 2.48 Ts/ (Ts + Tm) 0.60 0.54 0.59 0.50 0.60 0.80 0.73 0.61 0.57 0.43 0.51 0.56 0.48 0.50 0.49 0.68 0.19 0.32 0.32 0.32 0.31 0.31 0.31 0.29 0.27 0.31 0.36 0.52 0.54 0.49 0.11 0.22 C29bb/ (bb + aa) 0.65 0.62 0.65 0.71 0.74 0.74 0.65 0.66 0.71 0.65 0.51 0.60 0.63 0.59 0.58 0.76 0.68 0.67 0.69 0.72 0.67 0.67 0.69 0.68 0.69 0.63 0.70 0.66 0.74 0.67 0.64 0.24 C2920S (20S + 20R) 0.58 0.54 0.55 0.59 0.59 0.57 0.55 0.57 0.62 0.59 0.51 0.54 0.55 0.54 0.55 0.59 0.57 0.55 0.57 0.60 0.55 0.57 0.58 0.56 0.58 0.55 0.59 0.58 0.59 0.59 0.57 0.31 TA C20/ (C20 + C28,20R) 0.64 0.50 0.45 0.44 0.42 0.63 0.51 0.64 0.31 0.08 0.07 0.10 0.06 0.06 0.06 0.11 0.08 0.05 0.05 0.05 0.06 0.21 0.21 0.23 0.22 0.19 0.19 0.34 0.45 0.40 0.17 0.36 TA (I)/ TA (I + II) 0.57 0.49 0.45 0.44 0.42 0.55 0.45 0.60 0.37 0.08 0.07 0.12 0.07 0.07 0.07 0.14 0.10 0.07 0.06 0.06 0.07 0.23 0.22 0.24 0.23 0.19 0.19 0.38 0.44 0.41 0.22 0.31

Family I I I I I I I I II II II II II II II II III III III III III IV IV IV IV IV IV V V V VI VI

NDR* 0.18 0.14 0.18 0.14 0.23 0.18 0.18 0.15 0.14 0.12 0.14 0.12 0.13 0.11 0.11 0.11 0.11 0.13 0.14 0.11 0.12 0.14 0.14 0.13 0.13 0.14 0.15 0.15 0.19 0.17 0.25 0.19

d13Csat* 28.49 28.88 28.09 28.86 28.53 27.18 26.52 29.83 31.41 31.77 33.13 31.77 32.32 32.68 32.52 31.59 30.97 31.70 31.76 31.82 31.87 29.63 28.81 29.40 29.88 30.60 28.93 27.67 29.30 28.22 28.87 28.80

d13Caro* 28.75 28.56 28.38 28.50 28.40 27.55 27.57 29.66 30.25 30.85 31.24 31.41 31.63 31.92 32.03 30.63 28.64 30.30 30.25 30.47 30.27 29.30 29.29 29.10 29.49 30.22 28.93 26.13 28.27 26.92 26.79 27.31

d13Coil* 28.80 28.48 28.29 28.53 28.72 27.73 27.71 29.76 31.11 30.94 32.33 31.51 32.11 31.97 32.34 31.04 30.10 30.62 30.25 30.62 30.76 29.14 28.95 29.45 29.94 30.48 28.80 28.07 29.22 28.08 27.92 28.07

C26T/Ts 1.72 1.83 1.64 1.82 1.79 1.06 1.35 1.33 0.25 0.31 0.24 0.20 0.18 0.18 0.17 0.11 1.24 0.39 0.40 0.41 0.46 1.11 1.16 1.13 0.96 0.94 1.17 0.19 0.45 0.30 0.10 0.74

He et al. 1129

*Parameter was used in the chemometric analysis. C19/C19 + C23, C22/C21, C24/C23, and C26/C25 tricyclic terpanes; Tet/C23 = C24 tetracyclic terpane/C23 tricyclic terpane; %C27% = %C27% steranes/(%C27 to %C29 steranes); %C28 = %C28 steranes/(%C27 to %C29 steranes); %C29 = %C29 steranes/(%C27 to %C29 steranes); ETR = (C28 + C29 tricyclics)/ trisnorneohopane; C31R/H = C31 homohopane/hopane; C29/C30Hop = C29 30-norhopane/hopane; DiaH/(H + diaH) = C30 diahopane/(hopane + diahopane); C35/(C34 + C35)Hop = C35 homohopanes/(C34 + C35 homohopanes); GA/H = gammacerane/hopane; S/H = steranes/hopanes; Bicad = bicadinane/(bicadinane + hopane). NDR = 24-nordiacholestanes/(24- +27-nordiacholestanes); C30/(C2730)diaS = C30 diasteranes/(C27 to C30 diasteranes); TA-DMD3/C28S = C28 triaromatic demethyldinosterane/C28 stigmastane; TA-Dino = C29 triaromatic dinosteranes/(C29 dinosteranes + 3methylstigmastane 20R); C28/C29 = C28 steranes/C29steranes; C26T/Ts = C26 tricyclic terpane/trisnorneohopane; 3-/(3- + 4-MeS 20R) = 3-/(3-+4-methylstigmastane 20R); C24Tet/C26 = C24 tetracyclics/C26 trycyclics; TA C20/(C20 + C28, 20R) = desmethyl triaromatic steroids; TA(I)/TA(I + II) = (C21+C22)/(C21 + C22 + C26 + C27 + C28) desmethyl triaromatic steroids. (d13Csat, d13Caro, d13Coil = isotope ratios of saturate fraction, aromatic fraction and crude oil.) Many of these parameters are discussed in Peters et al. (2005).

TA (I)/ TA (I + II)

0.54 0.16

Russian side of the southern part of the Barents Sea Basin, the prolific northern Timan-Pechora Basin provided a major incentive to extend exploration offshore. The offshore part accounts for 30% of the total area of the Timan-Pechora Basin Province, approximately 131,700 km2 (50,850 mi2), which includes islands up to 5400 km2 (2085 mi2) in total area (Lindquist, 1999b). This offshore segment has not been studied as well as the onshore part of the Timan-Pechora Basin. However, with its wide shelf, it has great potential for oil and gas exploration. Paleozoic carbonate rocks comprise the principal reservoirs for the onshore TimanPechora Basin. Oil and gas are found in Paleozoic reservoirs ranging in age from Ordovician to Permian. Mississippian and Permian carbonate (limestone and dolomite) reservoirs formed in the offshore part of the northern Timan-Pechora Basin extending into the Barents Sea shelf, as in the Prirazlomnoye oil field. Trapping is generally within bioherm structures (reeflike organic buildups). The bioherms were exposed subaerially and underwent leaching by meteoric water, which can enhance reservoir properties (Wilson, 1975). Siliciclastic and carbonate reservoirs are the main reservoir types in the northern Timan-Pechora Basin. The siliciclastic reservoirs average 16% porosity and 154 md permeability, and the carbonate reservoirs average 13% porosity and 208 md permeability (Petroconsultants, 1996, personal communication). The Triassic sandstone reservoirs have superior reservoir properties compared with the other reservoir types in the northern TimanPechora Basin, but they are thin and discontinuous in the Barents Sea Basin. The Triassic sandstone reservoirs in the northern Timan-Pechora Basin have the highest porosity at 28% and average permeability from 75 to 345 md (Lindquist, 1999b). In the Barents Sea Basin, thick overburden rock and complicated tectonics created a variable thermal history. Mesozoic rifting produced the highest regional thermal gradient, and Cenozoic uplift resulted in significant thermal cooling with possibly hundreds of meters of erosion. Petroleum could have been generated by the Late Triassic from the rapidly buried Lower and Middle Triassic source rocks (Lindquist, 1999a).

TA C20/ (C20 + C28,20R) C2920S (20S + 20R) C29bb/ (bb + aa) Ts/ (Ts + Tm) C24Tet/ C26 3/ (3- + 4MeS20R) C26T/Ts d13Coil* d13Caro* d13Csat* TADino* TADMD3/ C28S* C30/ (C2730) diaS* NDR* Family Sample ID Table 2. Continued 1130 C28/ C29

Barents Sea and Northern Timan-Pechora

146 149

VI VI

0.19 0.24

0.00 0.00

0.07 0.17

0.5 0.86

30.07 29.93

29.06 29.14

29.36 29.47

0.34 0.41

1.47 0.29

0.74 0.27

1.19 3.87

0.36 0.38

0.61 0.54

0.47 0.51

0.58 0.14

SOURCE ROCK GEOLOGY Published source rock and organic geochemistry studies of the Barents Sea Basin are limited, especially for the Russian Barents Sea. The lower Paleozoic strata in the Russian Barents Sea are poorly studied. Four different source rock intervals in the Barents Sea and northern Timan-Pechora Basin are described in the literature, as discussed below. Upper Devonian (Domanik Facies) The Upper Devonian (Frasnian) Domanik Formation, organic-rich strata in the northern TimanPechora Basin, is generally considered to be the main hydrocarbon source rock in this region. The Domanik Formation is composed of black siliceous shales and limestones. It contains type II oilprone kerogen and has total organic carbon (TOC) up to 20%. The Domanik Formation was deposited in a tectonically controlled deep-water trough and was gradually filled by prograding clastic wedges between carbonate intervals during the Late DevonianTournaisian (Ulmishek, 1982). Because of its areal distribution and the history of formation of clastic wedges on margins of the deepwater trough, the Domanik facies might extend a short distance offshore along the continuation of the Devonian graben rift (Ulmishek, 1985). Because of its high economic importance, much speculation has been made as to whether the Domanik broadens westward and northward into the Barents Sea Basin. PennsylvanianLower Permian The PennsylvanianLower Permian strata have various facies in the Barents Sea and northern TimanPechora Basin. On Spitsbergen and Bear Islands, the PennsylvanianLower Permian layers are present as a carbonate sequence, which may have been deposited in local lagoons behind reefs. This carbonate sequence was deposited by a widespread marine transgression in arid climatic conditions. Restricted distribution and low TOC do not permit the PennsylvanianLower Permian strata to be significant source rocks. The PennsylvanianLower

Permian strata in the northeastern Timan-Pechora Basin are composed of black marine shale with TOC in the range of 0.75 to 1.00% (Ulmishek, 1982). This facies may extend a short distance into the southeastern corner of the Barents Shelf. LowerMiddle Triassic The Triassic section is the primary source rock interval in the Barents Sea. It is composed of black organic-rich shale with TOC in the range of 2 to 8% and a tendency to increase eastward. Almost 8 to 10% of the TOC can be extracted as bitumen from the source rock. The Triassic section contains mostly type II kerogen deposited under anoxic conditions in the gradually deepening sea (Ulmishek, 1982). The basinal deep-water facies continue far to the east and cover a significant part of the Barents Sea. Shallow-marine and nearshore Triassic facies occur in the South Barents and North Novaya Zemlya depressions. The organic matter in the source rock in this area probably has poor quality. The Triassic section in the northern Timan-Pechora Basin was likely deposited in a paleoshelf setting and contains both sapropelic and humic organic matter (Zakcharov and Kulibakina, 1996). However, Triassic source rock could be early mature in the onshore part of northern TimanPechora Basin (Zakcharov and Kulibakina, 1996; Kuranova et al., 1998). The Triassic source rock on northern Kolguyev Island and the adjoining shelf may be buried sufficiently to be mature. Lacustrine and estuarine facies are dominant in this area (Ulmishek, 1985). Upper Jurassic The Upper Jurassic black organic-rich shale is also a good source rock in the Barents Sea Basin. It was deposited in a deep-water anoxic environment. Low dilution of organic matter by sediment and restricted circulation of bottom water resulted in a high content of organic matter. These black shales are analogous to the Kimmeridgian Shale in the North Sea and Bazhenov shales in west Siberia that are both highly productive oil source rocks. The organic matter has a high content of type II
He et al. 1131

Figure 2. (A) Hierarchical cluster analysis for 34 oil samples based on 20 biomarkers and two stable carbon isotope ratios. (B) Principal components (PC) analysis for 34 oil samples based on 20 biomarkers and two stable carbon isotope ratios. 1132 Barents Sea and Northern Timan-Pechora

kerogen, and the organic-rich facies covers all of the Barents Shelf with the probable exception of narrow marginal areas. In the northern Barents Sea Basin, the Upper Jurassic shale ranges from hundreds to more than 1000 m (>3280 ft) thick, having a TOC of 1 to 9%. In contrast, the Jurassic shale in the central and southern Barents Sea is just 20 to 30 m (6698 ft) thick with TOC of 15 to 25% (Leith et al., 1993). This organic-rich shale is immature in the eastern Barents Sea Basin.

RESULTS AND DISCUSSION The HCA (Figure 2A) dendrogram identifies six oil families based on geochemical data for 34 oil samples from the Barents Sea and northern TimanPechora Basin. Highly thermally mature oil samples were excluded because secondary processes obscure genetic relationships by altering geochemical parameters. Most of the oil samples are nonbiodegraded and a few show light biodegradation. Twenty source-related biomarkers and two isotope parameters were used to construct the dendrogram (Figure 2A). The six identified oil families have broadly similar geochemical characteristics but may originate from different source rocks or different organofacies (Jones, 1987) of the same source rock. In addition to genetic information, age, lithology, and depositional environment of the source rock can be inferred from the detailed biomarker study, as discussed below. Six oil families were defined and can also be shown in the PCA plot (Figure 2B; Table 1). Oil-Oil Correlation and Geochemical Characteristics of Oil Families Biomarkers are complex molecular fossils derived from once-living organisms (Peters et al., 2005). Biomarkers in crude oil inherit the recognizable carbon structures of natural products synthesized by precursor organisms. Different assemblages of organisms live in different source rock depositional environments. Therefore, various biomarker parameters can be used to evaluate the deposi-

tional environment of the source rock. Biomarkers can be measured in both oil and source rock extracts, and they are helpful in source rock to oil correlation. In fact, biomarkers provide a method to interpret source rock organofacies when only oil samples are available. Biomarker fingerprints are primarily established by the source of organic matter and depositional environment, but the overprint of maturation is also important. Thus, numerous biomarker parameters should be taken into consideration to strengthen detailed correlations. We define a genetic oil family in this article as a group of oils that were generated and expelled from a single source rock organofacies. Some oil samples may show evidence of mixing but may be dominated by input from only one major source rock. Variations in oil characteristics caused by maturity or postexpulsion alteration processes, such as biodegradation and phase separation, should not alter genetic relationships. Representative gas chromatograms and terpane mass chromatograms for each oil family are shown in panels A to C of Figure 3. Key biomarker and nonbiomarker parameters used in this study to differentiate oil families include pristane/phytane (Pr/Ph); C29/C30 hopane; C33/C34 and C35/C34 homohopane; the distribution of C27, C28, and C29 steranes; and age-related biomarker ratios, such as the extended tricyclic terpane ratio (ETR) and triaromatic dinosteranes/(4-methylstigmastane or 3-methylstigmastane 20R). Maturity-related biomarkers, such as C29 20S/(20S + 20R), C29 bb/ (bb + aa), and Ts/ (Ts + Tm) (Ts = 18a-22, 29, 30-trisnorneohopane; Tm = 17a-22, 29, 30-trisnorhopane) ratios, are also examined in this study. Key biomarker parameters used to distinguish different oil families are summarized in Table 2, and details of the biomarker parameters that characterize each oil family are described below.

Family I Family I contains eight oil samples from Pennsylvanian, Lower Permian, and Triassic reservoirs from Spitsbergen Island, Kolguyev Island, and the northern Timan-Pechora Basin (Figure 1A, B; Table 1).
He et al. 1133

Figure 3. (A) Representative whole-oil gas chromatograms and 191 M/Z ion mass fragmentograms for family I and family II oils. (B) Representative whole-oil gas chromatograms and 191 M/Z ion mass fragmentograms for family III and family IV oils. (C) Representative whole-oil gas chromatograms and 191 M/Z ion mass fragmentograms for family V and family VI oils. Pr = pristane; Ph = phytane; Ts = 18-22, 29, 30-trisnorneohopane; Tm = 17-22, 29, 30-trisnorhopane.

The oils in this family are characterized by pristane greater than phytane (Figure 3A), which typically denotes a dysoxic to oxic depositional environment. The oils have relatively low ratios of pristane to n-C17 and phytane to n-C18, and stable carbon isotope values of crude oil 27.71 to 29.76, saturates 26.52 to 29.83, and aromatics 27.55 to 29.66. Oils in this family have normal to average tricyclic terpane concentrations, C29 hopane is less than C30, C33 homohopanes are more abundant than C34, and C34 homohopanes are greater than C35. Triaromatic 3-/(3- + 4-methylstigmastane 20R) is approximately one. A high ETR value (>2) is consistent with a Triassic source rock (Holba et al., 2001). However, the biomarkers suggest that two different lithofacies of source rock exist in this family. Samples 91, 67, and 102 (1) from northern Timan-Pechora Basin have high
1134 Barents Sea and Northern Timan-Pechora

C29/C30 hopane ratios (0.89), whereas samples 52(1), 78, 12, and 526(a) from Kolguyev Island, onshore of Timan-Pechora, and Spitsbergen Island show low C29/C30 hopane ratios (0.42). High TA-Dino/(TA-Dino + 3-methylstigmastane 20R) (i.e., TA-Dino = C29 triaromatic dinosteranes) and C28 triaromatic desmethyldinosterane/C28 stigmastane (TA-DMD3/C28S) ratios suggest Mesozoic and younger source rocks (Moldowan et al., 1996; Moldowan et al., 2001; Barbanti et al., 2011) (Table 2). Family II Family II consists of eight oils from Lower and Upper Devonian reservoirs in the northern TimanPechora Basin (Figure 1B; Table 1). This family of oils probably originated from Devonian Domanik source rocks. Oils in this family have very negative

Figure 3. Continued. He et al. 1135

stable carbon isotopic ratios of crude oil 30.94 to 32.34, saturates 31.59 to 33.13, and aromatics 30.25 to 32.03. Some classic geochemical parameters, such as Pr/Ph and the distribution of n-alkanes, indicate an oxic depositional environment and terrigenous input to the source rock (Peters et al., 2005) (Figure 3AC). Low C27/C29 and C28/C29 sterane ratios support higher plant input, and high C29/C30 hopane ratios suggest a marl source rock (Peters et al., 2005) (Table 2). The C34 homohopanes are more abundant than either the C33 or C35 homohopanes. Low ETR (0.45) and low TA-Dino/(TA-Dino + 3-methylstigmastane 20R) (TA-Dino = C29 triaromatic dinosteranes) and C28 triaromatic desmethyldinosterane/C28 stigmastane (TA-DMD3/C28S) ratios imply a Paleozoic source rock (Moldowan et al., 1996, 2001; Barbanti et al., 2011) (Table 2). The organic matter that generated family II oils has a peak oil window maturity based on the maturity-related biomarkers. Family III Family III consists of five oil samples from Devonian and Lower Permian reservoirs in the Pechora Sea and the northern Timan-Pechora Basin (Figure 1B; Table 1). This family of oils probably originated from a different organofacies of the Domanik Formation than family II. The oils in this family display biomarker distribution similar to family II oils. Pristane/phytane ratios less than one (Figure 3B) for this family indicate an anoxic depositional environment. The stable carbon isotopic ratios for crude oils in this family are 30.10 to 31.54, saturates 30.97 to 31.87, and aromatics 28.64 to 30.47. Low TA-Dino/(TA-Dino + 3-methylstigmastane 20R) (TA-Dino = C29 triaromatic dinosteranes) and C28 triaromatic desmethyldinosterane/C28 stigmastane (TA-DMD3/C28S) ratios imply Paleozoic source rock (Moldowan et al., 1996, 2001; Barbanti et al., 2011) (Table 2). However, high C29/C30 (>70%) hopane ratios strongly indicate a carbonate source rock for this oil family. The ETR ratios for this oil family (1) are higher than those of family II (0.45). The C19/(C19 + C23) and C24 tetracyclic terpane/C23 tricyclic terpane (Tet/C23) ratios for family III are lower than
1136 Barents Sea and Northern Timan-Pechora

those of family II. Family III has higher C22/C21 ratios than family II. Family IV Family IV contains six oil samples in Devonian and Lower Permian reservoirs from the northern Timan-Pechora Basin (Figure 1B; Table 1). The oils in this family are characterized by Pr/Ph ratios greater than one (Figure 3B), which indicates a dysoxic to oxic depositional environment. The isotope values for crude oils are 28.8 to 30.48, saturates 28.81 to 30.97, and aromatics 28.64 to 30.22. Most oils in this group were slightly biodegraded, as evidenced by an unresolved mixture (hump) on gas chromatograms. Low C27/ C29 and C28/C29 ratios indicate terrigenous source input. Low TA-Dino/(TA-Dino + 3-methylstigmastane 20R) (TA-Dino = C29 triaromatic dinosteranes) and C28 triaromatic desmethyldinosterane/C28 stigmastane (TA-DMD3/C28S) ratios also probably imply a Paleozoic source rock (Moldowan et al., 1996, 2001; Barbanti et al., 2011) (Table 2). Family V Family V consists of three oil samples from Permian and Lower Triassic reservoirs on the Pechora Shelf and Kolguyev Island (Figure 1A; Table 1). These oils are characterized by high Pr/Ph ratios that indicate a dysoxic to oxic depositional environment (Figure 3C). The stable carbon isotopic values of crude oils are 28.07 to 29.22, saturates 27.67 to 29.3, and aromatics 26.13 to 28.27. Bimodal gas chromatograms for these oils indicate a terrigenous source input. High C29/ C27 and C29/C28 sterane ratios also show higher plant input to this family of oils. Relatively low C33, C34, and C35 homohopane peaks on the M/Z 191 ion chromatogram also indicate dysoxic to oxic depositional conditions in the source rock of family V (Figure 3C). These oils have reached peak oil window maturity. Bicadinane exists in this family. Low ETR (<2) and high TA-Dino/ (TA-Dino + 3-methylstigmastane 20R) (TA-Dino = C29 triaromatic dinosteranes) and C28 triaromatic desmethyldinosterane/C28 stigmastane (TA-DMD3/ C28S) ratios indicate that this oil family could be

from Jurassic source rock (Moldowan et al., 1996, 2001; Barbanti et al., 2011) (Table 2). The low C26 tricyclic terpane ratios (C26T/Ts <1.0) for this family are diagnostic of Jurassic or younger source rocks (Peters et al., 2007). Family VI Family VI contains four oil samples from Lower and Middle Jurassic reservoirs. Three of the samples are from the Shtokmanovskoye field in the Barents Sea from Lower Jurassic (one) and Middle Jurassic (two) reservoir rocks, and one oil sample is from Spitsbergen Island (Figure 1A; Table 1). Family VI oil samples are characterized by high Pr/Ph ratios (>3) that indicate an oxic depositional environment (Figure 3C). The stable carbon isotope values of crude oils are 27.92 to 29.47, saturates 28.8 to 30.07, and aromatics 26.79 to 29.14. Gas chromatograms of these oils show a small hump indicating light biodegradation. This family has 13C-rich stable carbon isotope ratios for crude oil (28.07 to 29.47), saturate hydrocarbons (28.80 to 30.07), and aromatic hydrocarbons (27.31 to 29.14). The high C19/ (C19 + C23), low C22/C21, and high C24/C23 tricyclic terpane ratios and elevated %C29 steranes may indicate paralic deltaic marine shale source rock with higher plant input (Peters et al., 2005). The elevated C29/C30 hopane ratio suggests a marine marl source rock. Low ETR (<2) and high TA-Dino/(TA-Dino + 3-methylstigmastane 20R) (TA-Dino = C29 triaromatic dinosteranes) and C28 triaromatic desmethyldinosterane/C28 stigmastane (TA-DMD3/C28S) ratios indicate that this oil family could originate from Jurassic source rock (Moldowan et al., 1996, 2001; Barbanti et al., 2011) (Table 2).

Source Rock Interpretation from Oil Geochemistry Because no source rock samples were analyzed in this study, direct oil-source correlation could not be conducted. However, geochemical characteristics of the oil samples were used to infer the source rock organofacies (Peters et al., 2005). In addition,

understanding of the depositional environment of source rock can be improved by comparison of the data with published source rock and oil data from the general area of the study. The ETR can be used to distinguish crude oil generated from Triassic and MiddleUpper Jurassic source rocks (Holba et al., 2001). Families I and IV have ETR greater than or equal to 2 and therefore could be generated from Triassic source rock. Two lithofacies for the source rocks of family I are indicated by the biomarker parameters: a carbonate facies in the onshore part of northern TimanPechora Basin and a deep-marine shale facies in the Pechora Sea and Spitsbergen. Families II, III, V, and VI show ETR less than or equal to 2, with most less than 1.2, which could indicate a Jurassic source rock. Triaromatic dinosteroids are generally abundant in Mesozoic samples but are below detection limits for most of the Paleozoic marine samples (Moldowan et al., 2001). High TA-Dino/ (TA-Dino + 3-methylstigmastane 20R) (TA-Dino = C29 triaromatic dinosteranes) and C28 triaromatic desmethyldinosterane/C28 stigmastane (TA-DMD3 /C28S) ratios in families I, V, and VI suggest a Mesozoic and younger source rock. In contrast, low TA-Dino/(TA-Dino + 3-methylstigmastane 20R) (TA-Dino = C29 triaromatic dinosteranes) and C28 triaromatic desmethyldinosterane/C28 stigmastane (TA-DMD3/C28S) ratios in families II, III, and IV imply Paleozoic source rock (Moldowan et al., 1996, 2001; Barbanti et al., 2011) (Table 2). Bicadinanes may be derived from resins produced by some angiosperms (van Aarssen et al., 1992). They tend to be absent or below detection limits in oils generated from sub-Jurassic source rock and begin to occur and increase in younger oil samples (Barbanti and Moldowan, 2002, personal communication). Summons et al. (1995) report significant bicadinanes in all of the Jurassic oils that they studied from the Perth Basin. The high bicadinane index for family V indicates that they originated from Jurassic or younger source rock. In addition, the geochemical characteristics of family V oils are similar to those of oils and rock extracts from the Upper Jurassic Bazhenov Formation in nearby west Siberia, such as high C26 tricyclic terpane ratios, low C19/(C19 + C23), low
He et al. 1137

Figure 4. Comparison of compound-specific isotope analysis of n-alkanes in selected oil samples from each identified oil family in the Barents Sea and northern Timan-Pechora Basin.

C22/C21, high C24/C23, high % C29 steranes, low C28/C29 sterane ratios, and low C29/C30 hopane ratios (Peters et al., 1993, 1994, 2007). Although the oil samples in families V and VI are similar, distinct biomarker ratios of family VI, such as higher C26 tricyclic terpane ratios, % C29 steranes, C19/(C19 + C23), and C29/C30 hopane ratios, and lower C22/C21, C24/C23 and C28/C29 sterane ratios than those of family V (Table 2) are similar to published geochemical features of the Lower Middle Jurassic Tyumen Formation (Peters et al., 1993, 1994, 2007). Oils in families II and III show low Pr/Ph ratios (<2), dominant C28 and C29 steranes, C24 tetracyclic > C26 tricyclic terpanes, and C29 < C30 hopanes

(Table 2). These two oil families have smooth distributions of normal alkanes with maxima at C11 C15. These geochemical characteristics are similar to those of source rock extracts from the Upper Devonian Domanik Formation (Abrams et al., 1999). This suggests that families II and III originated from the Domanik Formation. Similar compoundspecific isotope analyses of n-alkanes for selected oil samples from both families support the Domanik source rock (Figure 4). However, the families could be from different organofacies of the Domanik Formation because family III has a high C29/C30 hopane ratios (>0.85), indicating carbonate source rock, whereas family II has low values, indicating a marl source rock.

Figure 5. Schematic relationship between concentrations of methyldiamantanes and stigmastane (5a, 14a, 17a[H]-24ethylcholestane 20R) for oil samples from the Barents Sea and northern Timan-Pechora Basin.

1138

Barents Sea and Northern Timan-Pechora

Age-related biomarker ratios show that family IV oil samples may originate from Paleozoic source rock. However, the geochemical features of family IV differ from those of families II and III. Compound-specific isotope analyses of n-alkanes and diamondoids for selected oil samples indicate that family IV represents mixed input from Domanik and Triassic or Domanik and Jurassic source rocks, as discussed below.

Identification and Deconvolution of the Mixed Oils and Deep Source Interpretations Mixing of oil types can be investigated using gas chromatography, mass chromatograms of biomarkers, stable carbon isotope ratios, and chemometrics. Age- and lithology-related biomarker parameters and oil-oil correlation ternary plots are also powerful tools to identify oil mixtures (Peters et al., 2005). In the PCA of the data set, samples 144, 81 (3), and 81(2) show features that suggest mixing (Figure 2B). The high TA-Dino and TA-DMD3/ C28S ratios for sample 81(3) indicate a Mesozoic cosource for this oil. Many of the biomarker ratios indicate that samples 80(1) and 81(2) were mostly generated from the Domanik Formation, but elevated TA-Dino and TA-DMD3/C28S for these samples 80(1) and 81(2) reflect a minor Mesozoic cosource (Moldowan et al., 1996, 2001; Barbanti et al., 2011) (Table 2). By combining age-related biomarker ratios with PCA results for these two oil samples, it can be concluded that sample 80(1) may have minor Triassic source input, and that sample 81(2) may have a minor Jurassic source input. This result is also supported by CSIA of the n-alkanes. End-member oil samples representing each oil family and proposed mixed oils were selected for CSIA of n-alkanes (Figure 4). These CSIA results show similar isotopic ratios for families II and III, supporting the interpretation that both families are from the Domanik Formation. The PCA together with the CSIA data show that for samples 144, 44, and 81(3) from family IV, the Triassic is most likely the major source rock, whereas the Devonian Domanik is a secondary source. For sample 50 from family III, the Do-

manik is a major source, and the Triassic is a secondary source (Figures 2B, 4). In addition, sample 144 appears to have a Jurassic cosource because of the presence of bicadinane. Based on the relationship between the two diamondoids (3- and 4-methyldiamantanes) and stigmastane (Dahl et al., 1999), 14 oil samples in this data set can be interpreted as mixtures of low- to moderate-maturity oil containing biomarkers and high-maturity cracked oil rich in diamondoids (Figure 5). The quantities of individual diamondoids in the oil samples are shown in Table 3. Before the extent of oil cracking is estimated, the diamondoid baseline for each oil family is determined using the concentration of 3- and 4-methyldiamantanes in a group of noncracked, nonbiodegraded, and nonfractionated oil samples from the same source rock. Because of the limited number of samples from each family, we determined the approximate diamondoid baseline using the minimum diamondoid concentration for oil samples generated from the same source rock. The extent of oil cracking for each sample was calculated using the formula (1-C0/Cc)100, where C0 is the diamondoid baseline and Cc is the measured methyldiamantane concentration in any oil sample. The diamondoid baseline for each oil family and the extent of oil cracking are listed in Table 3. Samples 91, 67, 102(1), and 12 from family I, samples 54, 274, 81(4), and 44 from family IV, and sample 151 from family V have high concentrations of diamondoids and relatively high concentrations of biomarkers, indicating strongly cracked oil mixed with low-maturity oil. Samples 526(a) from family I, 81(2) from family II, 81(3) from family IV, 150 from family V, and 146 from family VI could be mixtures of light to moderately cracked oil with low-maturity oil because of the relatively low concentrations of diamondoids and biomarkers in these samples. The remaining oil samples have not undergone cracking. In this study, CSIA-D was used to identify the highly cracked component in mixtures of highand low-maturity oil. Noncracked oil from each oil family (end-member oil) and some representative mixtures were selected for CSIA-D (Figure 6).
He et al. 1139

Table 3. Quantification of Diamondoids (ppm) and Biomarkers (ppm), Extent of Oil Cracking of the Oil Samples from the Barents Sea and Northern Timan-Pechora Basin with CSIA-D* Data of the Selected Oil Samples Sample ID Family I I I I I I I I II II II II II II II II III III III III III IV IV IV IV IV IV V V V VI VI VI VI 3- + 4Methyldiamantanes (ppm) 6.05 64.99 45.77 61.57 66.06 4.82 6.70 16.24 11.78 3.40 4.01 3.92 5.13 5.12 5.44 2.39 3.07 4.93 3.68 3.74 2.57 38.30 44.47 26.57 25.82 10.77 7.44 22.75 7.15 16.61 2.35 8.56 14.05 6.93 C29 Stigmastane (ppm) 2.23 15.57 11.69 18.92 15.89 3.67 2.87 13.99 31.35 121.81 107.92 38.88 129.53 114.88 128.33 45.52 72.91 160.69 98.63 101.70 82.01 67.56 48.99 41.90 73.06 44.20 21.94 18.87 11.26 32.14 95.98 2.30 1.34 0.00 Diamondoid Baseline (ppm) 4.82 4.82 4.82 4.82 4.82 4.82 4.82 4.82 2.39 2.39 2.39 2.39 2.39 2.39 2.39 2.39 2.57 2.57 2.57 2.57 2.57 7.44 7.44 7.44 7.44 7.44 7.44 7.15 7.15 7.15 2.35 2.35 2.35 2.35 Cracking Extent 20.36 92.58 89.47 92.17 92.70 0.00 28.09 70.32 79.72 29.77 40.37 38.96 53.39 53.30 56.05 0.19 16.36 47.85 30.26 31.28 0.00 80.57 83.27 72.00 71.19 30.91 0.00 68.57 0.00 56.96 0.00 72.55 83.28 66.07 1-Methyladamantane, d13C 19.56 19.75 18.72 19.38 20.51 1,4-Dimethyladamantane (cis), d13C 21.84 21.91 21.26 22.13 23.13

1140 Barents Sea and Northern Timan-Pechora

152 102(1) 91 67 12 78 52(1) 526a 81(2) 87 22(1) 80(2) 65 22(3) 22(2) 43 50 880 41 102(2) 80(1) 274 44 81(4) 54 81(3) 144 151 220 150 561 147 146 149

21.76

24.29

19.97 20.81

23.24 22.46

20.49 18.52 18.84

22.41 20.75 20.71

Table 3. Continued 1,41,31,3,41,3,41,3,61,3,53Dimethyladamantane Dimethyladamantane, Trimethyladamantane Trimethyladamantane Trimethyladamantane, Trimethyladamantane, Dimethyldiamantane, Sample ID Family (trans), d13C d13C (cis),d13C (trans), d13C d13C d13C d13C 152 102(1) 91 67 12 78 52(1) 526a 81(2) 87 22(1) 80(2) 65 22(3) 22(2) 43 50 880 41 102(2) 80(1) 274 44 81(4) 54 81(3) 144 151 220 150 561 I I I I I I I I II II II II II II II II III III III III III IV IV IV IV IV IV V V V VI 21.68 21.86 21.81 22.21 23.51 23.06 22.86 22.73 23.05 23.43 24.03 23.73 23.94 23.56 24.02 24.30 23.81 24.12 23.63 24.10 24.09 24.19 22.76 23.82 24.47 25.27 25.39 24.89 25.21 25.46 29.19 28.77 28.17 29.50 28.75

24.19

25.46

25.61

26.72

30.18

24.33 23.28 22.78

24.80 23.38 24.07

26.65 23.76 24.23

26.70 24.20 24.48

25.16 25.11 24.41

25.93 26.07 26.27

29.54 29.81 29.79

He et al. 1141

22.22 21.15 21.19

23.58 21.49 20.85

24.44 22.56 22.72

24.51 22.51 22.58

24.49 21.61 22.24

25.93 23.63 23.57

29.79 26.80 26.87

1,41,31,3,41,3,41,3,61,3,53Dimethyladamantane Dimethyladamantane, Trimethyladamantane Trimethyladamantane Trimethyladamantane, Trimethyladamantane, Dimethyldiamantane, Sample ID Family (trans), d13C d13C (cis),d13C (trans), d13C d13C d13C d13C

Jurassic oil samples show diamondoids that are enriched in 13C, whereas the proposed Domanik oil samples are relatively depleted in 13C (Figure 6; Table 3). Triassic oil samples show diamondoid isotope values intermediate between those of the putative Jurassic and Domanik oil samples. Sample 151, which contains a highly cracked oil and sample 220, a Jurassic-sourced end member, show similar diamondoid isotope values, which indicates that the highly cracked part of sample 151 is from a deep high-maturity Jurassic source rock. Different CSIA-D results between Triassic-sourced samples 78 (end-member oil) and 52(1) and mixed samples 102(1) and 67 from a Triassic source rock with highly cracked components indicate that a deep high-maturity source provides components to the Triassic low-maturity oil samples. The logical deep-source candidate is the Domanik Formation generating at high maturity. In addition, the Triassic is not buried deeply enough to be a deepsource candidate in the onshore part of the northern Timan-Pechora Basin where samples 102(1) and 67 were obtained (Zakcharov and Kulibakina, 1996). The CSIA-D data also support this conclusion. The CSIA-D data for samples 102(1) and 67 are between the Triassic end-member sample 78 and the Domanik end-member sample 22(2). In addition, PCA (Figure 2B) and CSIA of n-alkanes (Figure 4) results are also consistent with this conclusion. Samples 144, 44, and 274 from family IV are proposed as mixtures of Triassic and Domanik sources based on biomarker parameters. The CSIA-D results for those three samples are between the Triassic and Domanik end-members, in agreement with the biomarker results (Figures 2B, 6). Sample 144 has very low concentrations of diamondoids, and the CSIA-D result for sample 144 is very similar to that of Triassic oil samples (Figure 6), which supports the biomarker evidence above that this oil has a major contribution from Triassic source rock and minor contributions from Domanik and Jurassic source rock. Sample 81(2) is a mixture of Jurassic and Domanik input based on both biomarkers and the CSIA-D results (Figure 6). In sum, the high-maturity Domanik source rock could be the deep source for several mixed oils having both highly cracked and normal maturity components

1142

Barents Sea and Northern Timan-Pechora

*CSIA-D = compound-specific isotope analysis of diamondoids.

Table 3. Continued

147 146 149

VI VI VI

Figure 6. Comparison of compound-specific isotope analysis of diamondoids in selected oils from each identified oil family in the Barents Sea and northern Timan-Pechora Basin: (1) 1-methyladamantane; (2) 1,4dimethyladamantane (cis); (3) 1,4-dimethyladamantane (trans); (4) 1,3-dimethyladamantane; (5) 1,3,4-trimethyladamantane (cis); (6) 1,3,4trimethyladamantane (trans); (7) 1,3,6-trimethyladamantane; (8) 1,3,5-trimethyladamantane (cis); (9) 3-trimethyladamantane.

from families I and IV in the northern TimanPechora Basin.

Distribution of Oil Families and Oil Migration Integration of the geology and the geochemical data for each oil family can help us predict the areal and stratigraphic distribution of source rocks, which can help determine the geographic extent of petroleum systems in the Barents Sea and northern Timan-Pechora basins. The geographic distribution of the six oil families is shown in panels A and B of Figure 1 and summarized in Table 1. Based on the geochemical data, we propose that family I, which includes oils from the northern Timan-Pechora Basin, Kolguyev Island, and Spitsbergen Island, originated from a Triassic source rock. Our result supports the contention that the Triassic sequence covers all of the Barents Sea and northern Timan-Pechora Basin (Ulmishek, 1985). The Triassic source rock in the north onshore part of Timan-Pechora Basin was a carbonate deposited on a paleoshelf. Most of the source rocks for oil samples from the northern Timan-Pechora Basin and Pechora Shelf were deposited in the transition zones from continental to marine facies, whereas the Triassic source rock in the Barents Sea and Spitsbergen Island has a deep-marine shale facies (Zakcharov and Kulibakina, 1996). The thickness of Triassic deposits on the Pechora Shelf increases from south to north. The structure high of the

Timan-Pechora Basin resulted from repeated tectonic movement during the Late Permian and Mesozoic, but this activity did not bury the Triassic source rock deeply enough to allow cracking in the northern Timan-Pechora Basin (Zakcharov and Kulibakina, 1996). Thus, the Triassic source rock could not be a high-maturity deep source in the northern Timan-Pechora Basin. However, the Domanik Formation was active as a deep high-maturity source rock in this area. The Triassic section also contains good local clay seals, which tend to prevent vertical migration in the offshore part of northern Timan-Pechora Basin (including the Tarskoye and Pechano-Ozerskoye fields on Kolguyev Island) and on Spitsbergen Island (including Barentsburg field). However, vertical and lateral migration occurred frequently in the the Kolva Megaswell (including the Kharyaginskoye, Khylchuyuskoye, and Yareyyuskoye fields) and Adzva-Varandey structural area (including the Varandey, Trebsa, Naulskoye, and Toboiskoye fields), which is where most of the onshore Triassic (family I) and mixed TriassicDomanik (family IV) oils are located (Figures 1B, 2B; Table 1). Oil samples from families V and VI presumably originated from Upper Jurassic and Lower Middle Jurassic source rocks, respectively. The oils in family V are mostly from the Pechora Shelf. The Upper Jurassic claystone having the thickness of 50 to 150 m (164492 ft) in this area is a good regional seal (Zakcharov and Kulibakina, 1996), which makes vertical migration difficult. That
He et al. 1143

may explain why the family V oil samples from this area are not mixed with oil from other sources. Because the Upper Jurassic source rock is thinner (2030 m [6698 ft]) in the central and southern Barents Sea and has low maturity in the eastern Barents Sea, it is likely not a major source rock for the Shtokmanovskoye field in the southern Barents Sea Basin. Our interpreted results show that the three oil samples from the Shtokmanovskoye field belong to family VI. Family VI is interpreted to originate from a LowerMiddle Jurassic source rock, which may also extend farther north to Spitsbergen Island where one family VI oil sample was collected. Oils in families II and III originated from the Domanik Formation, which is generally thought to be the main source rock in the Timan-Pechora Basin. The Domanik Formation consists mostly of organic-rich, black, siliceous shales and limestones with a high TOC, which fits the geochemical characteristics of family II in this study. In addition, our study proposes a carbonate facies of the Domanik Formation. The oil samples in family III originated from this carbonate Domanik source rock and are primarily located on the Pechora Shelf and near offshore of the northern Timan-Pechora Basin. Because of the economic importance of the Domanik Formation, many speculate that it extends offshore and broadens westward and northward into the Barents Sea Basin. Our results suggest that the Domanik Formation extends offshore into the Pechora Sea, and that a corresponding facies change into carbonate occurs from the onshore siliceous shales. The CSIA-D results in this study indicate that the Domanik Formation on the Pechora Shelf has high maturity and is active as a deep source for the mixed oils in this area.

proximately 55.4% and C29 bb/(bb + aa) 20R sterane ratios of approximately 65.2%, indicating a peak to late oil window maturity (Table 2). The ratios of Ts/(Ts + Tm), TA (I)/TA (I + II), and TA C20/[(C20 + C28, 20R)] for family I oils (61, 45, and 44%, respectively) also support this conclusion. However, samples 67, 91, and 102(1) have relatively low maturity compared with other oil samples in family I based on these parameters. The relatively high tricyclic terpanes/hopanes, high Ts/(Ts + Tm), high diasteranes content, and low concentrations of biomarker compounds also indicate that some of the oils in family I have a late oil window maturity. The oil samples in families II and III have typical oil window maturities, and some have reached the peak to late oil generation stage. The oil samples in family IV exhibit early to peak oil window maturity range based on biomarker maturity parameters. The oil samples in family V show a peak oil maturity and those in family VI have an oil window maturity.

CONCLUSIONS Based on molecular and isotope evidence, we interpreted six oil families representing five major source rocks in the Barents Sea and northern TimanPechora Basin. The oil samples in family I are from Triassic source rock, which is mostly an organicrich black shale. The Triassic sequence covers most of the Barents Sea and northern Timan-Pechora Basin, and representatives of family I are located in the northern Timan-Pechora Basin, Kolguyev Island, and Spitsbergen Island. The Triassic source rock has a low maturity and a carbonate lithofacies in the north onshore part of the Timan-Pechora Basin. The Triassic source rock in the Barents Sea and Spitsbergen Island consists of deep-marine shale facies. The Triassic section also contains good local clay seals, which prevent vertical migration in the offshore part of northern Timan-Pechora Basin (including Tarskoye and Pechano-Ozerskoye fields on Kolguyev Island) and on Spitsbergen Island (including Barentsburg field) (Figures 1B, 2B; Table 1). The oil samples in families II and III

Thermal Maturity Maturity levels of the oils were assessed based on the following parameters: C29 sterane isomerization ratios, Ts/(Ts + Tm), TA C20/[(C20 + C28, 20R)] and TA (I)/TA (I + II). The family I oil samples have C29 20S/(20S + 20R) sterane ratios of ap1144 Barents Sea and Northern Timan-Pechora

originated from different organofacies of the Devonian Domanik Formation. Our geochemical results show that the Domanik Formation is an effective source rock in the northern Timan-Pechora Basin, which suggests that it extends offshore into the Pechora Sea, having a carbonate lithofacies. The oil samples in family V originated from Upper Jurassic source rocks located in the Pechora Shelf and southern Barents Sea. Good clay seals in the Jurassic section in this area militate against the mixing of Jurassic oil with other sources. The oil samples in the Shtokmanovskoye field are assigned to family VI and probably originated from a LowerMiddle Jurassic source rock that may extend farther north to Spitsbergen Island based on our geochemical study. Mixed oils were recognized in this study using biomarker and compound-specific isotope data. The oil samples in family IV are mostly cosourced by Triassic and Domanik source rocks and are located in the Pechora Shelf and Adzva-Varandey structural area (including the Varandey, Trebsa, Naulskoye, and Toboiskoye fields). Active vertical and lateral migration occurred in this area possibly because good local Triassic clay seals do not exist. We believe that the deep source rock for the oils in the northern Timan-Pechora Basin is the highmaturity Domanik Formation.

REFERENCES CITED
Abrams, M. A., A. M. Apanel, O. M. Timoshenko, and N. N. Kosenkova, 1999, Oil families and their potential sources in the northeastern Timan Pechora Basin, Russia: AAPG Bulletin, v. 83, p. 553577. Barbanti, S. M., J. M. Moldowan, D. S. Watt, and E. Kolaczkowska, 2011, New triaromatic steroids distinguish Paleozoic from Mesozoic oil: Organic Geochemistry, v. 42, p. 409424. Bjory, M., J. O. Vigran, and T. M. Rnningsland, 1978, Source rock evaluation of Mesozoic shales from Svalbard: IKU (Continental Shelf Institute) Open Report 160/1/78, p. 214. Dahl, J. E., J. M. Moldowan, K. E. Peters, G. E. Claypool, G. E. Rooney, G. E. Michael, M. R. Mello, and M. L. Kohnen, 1999, Diamondoid hydrocarbons as indicators of natural oil cracking: Nature, v. 399, p. 54 57, doi:10.1038 /19953. Dalland, A., D. Worsley, and K. Ofstad, 1988, A lithostratigraphic scheme for the Mesozoic and Cenozoic succes-

sion offshore for the Mesozoic and Cenozoic succession offshore mid- and northern Norway: Norwegian Petroleum Directorate Bulletin, v. 4, p. 65. Dore, A. G., 1995, Barents Sea geology, petroleum resources and commercial potential: Arctic, v. 48, p. 207221. Gramberg, I. S., 1988, The Barents Shelf Platform: PGO Sevmorgeologia, Leningrad, Nedra, 263 p. Holba, A. G., L. Ellis, I. L. Dzou, A. Hallam, W. D. Masterson, J. Francu, and A. L. Fincannon, 2001, Extended tricyclic terpanes as age discriminators between Triassic, Early Jurassic and MiddleLate Jurassic oils: 20th International Meeting on Organic Geochemistry, p. 464. Johansen, S. E., B. K. Ostisty, O. Birkeland, Y. F. Fedorovsky, VI. N. Martirosjan, O. B. Christensen, S. I. Cheredeev, E. A. Ignatenko, and L. S. Margulis, 1993, Hydrocarbon potential in the Barents Sea region: Play distribution and potentialArctic geology and petroleum potential: Amsterdam, Netherlands, Elsevier Science, NPF (Norwegian Petroleum Society) Special Publication 2, p. 273320. Jones, R. W., 1987, Organic facies, in J. Brooks and D. Welte, eds., Advances in petroleum geochemistry: London, Academic Press, v. 2, p. 191. Kuranova, L. VI., L. A. Plekhotkina, and N. N. Kosenkova, 1998, Conditions of sediment accumulation and formation of pools in Peschanoozer gas-condensate-oil field: Petroleum Geology, v. 32, p. 151156. Larsen, R. M., T. Fjaeran, and O. Skarpnes, 1993, Hydrocarbon potential of the Norwegian Barents Sea based on recent well results: Arctic geology and petroleum potential: Amsterdam, Netherlands, Elsevier Science, NPF (Norwegian Petroleum Society) Special Publication 2, p. 321331. Leith, T. L., et al., 1993, Mesozoic hydrocarbon source rocks of the Arctic region: Arctic geology and petroleum potential: Amsterdam, Netherlands, Elsevier, NPF (Norwegian Petroleum Society) Special Publication 2, p. 125. Lindquist, S. J., 1999a, South and North Barents Triassic Jurassic total petroleum system of the Russian offshore Arctic: U.S. Geological Survey Open-File Report 9950-G, 23 p. Lindquist, S. J., 1999b, The Timan-Pechora Basin province of northwest Arctic Russia: DomanikPaleozoic total petroleum system: U.S. Geological Survey Open-File Report 99-50-G, 24 p. Magoon, L. B., and W. G. Dow, 1994, The petroleum system: From source to trap: AAPG Memoir 60, 655 p. Moldowan, J. M., S. R. Jacobsen, B. J. Huizinga, D. S. Watt, K. E. Peters, and F. Fago, 1996, Chemostratigraphic reconstruction of biofacies: Molecular evidence linking cyst-forming dinoflagellates with pre-Triassic ancestors: Geology, v. 24, p. 159162, doi:10.1130/0091-7613 (1996)024<0159:CROBME>2.3.CO;2. Moldowan, J. M., S. R. Jacobsen, A. Al-Hajji, B. J. Huizinga, and F. Fago, 2001, Molecular fossils demonstrate Precambrian origin of dinoflagellates, in A. Zhuravlev and R. Riding, eds., Ecology of the Cambrian radiation: New York, Columbia University Press, p. 474492. Moldowan, J. M., Z. Chen, Z. Wei, F. Fago, J. Dahl, S. Schouten, J. S. Sinninghe Damste, W. J. Carrigan, P. Jenden, and M. A. Abu-Ali, 2004, High-resolution geochemical technologies (HRGTs) applied to understanding petroleum

He et al.

1145

systems in Saudi Arabia: 6th International Conference on Petroleum Geochemistry and Exploration in the AfroAsian Region, Beijing, China, p. 108109. Nyland, B., L. N. Jensen, J. Skagen, O. Skarpnes, and T. O. Vorren, 1992, Tertiary uplift and erosion in the Barents Sea: Magnitude, timing, and consequences: Amsterdam, Netherlands, Elsevier Science, Norwegian Petroleum Society (NPF) Special Publication 1, p. 153162. Ostisty, B. K., and S. I. Cheredeev, 1993, Main factors controlling regional oil and gas potential in the west Arctic, former USSR: Amsterdam, Netherlands, Elsevier Science, Norwegian Petroleum Society (NPF) Special Publication 3, p. 591597. Peters, K. E., and M. R. Cassa, 1994, Applied source rock geochemistry, in L. B. Magoon and W. G. Dow, eds., The petroleum system: From source to trap: AAPG Memoir 60, p. 93120. Peters, K. E., A. E. Kontorovich, J. M. Moldowan, VI. E. Andrusevich, B. J. Huizinga, G. J. Demaison, and O. F. Stasova, 1993, Geochemistry of selected oils and rocks from the central portion of the West Siberian Basin, Russia: AAPG Bulletin, v. 77, p. 863887. Peters, K. E., A. E. Kontorovich, B. J. Huizinga, J. M. Moldowan, and C. Y. Lee, 1994, Multiple oil families in the West Siberian Basin: AAPG Bulletin, v. 78, p. 893909. Peters, K. E., C. C. Walters, and J. M. Moldowan, 2005, The biomarker guide: Cambridge, United Kingdom: New York, Cambridge University Press, 1155 p. Peters, K. E., L. S. Ramos, J. E. Zumberge, Z. C. Valin, C. R. Scotese, and D. L. Gautier, 2007, Circum-Arctic petroleum systems identified using decision-tree chemometrics: AAPG Bulletin, v. 91, p. 877 913, doi:10 .1306/12290606097. Petroconsultants, 1996, Petroleum exploration and production database: Houston, Texas, Petroconsultants, Inc.

Schoell, M., and R. M. K. Carlson, 1999, Diamondoids and oil are not forever: Nature, v. 399, p. 1516, doi:10 .1038/19847. Seifert, W. K., and J. M. Moldowan, 1979, The effect of biodegradation on steranes and terpanes in crude oils: Geochimica et Cosmochimica Acta, v. 43, p. 111126, doi:10.1016/0016-7037(79)90051-6. Summons, R. E., C. J. Boreham, C. B. Foster, A. P. Murray, and J. D. Gorter, 1995, Chemostratigraphy and the composition of oils in the Perth Basin, Western Australia: Australian Petroleum Exploration Association Journal, v. 35, p. 613631. Ulmishek, G., 1982, Petroleum geology and resource assessment of the Timan-Pechora Basin, USSR, and the adjacent Barents-northern Kara shelf: Argonne, Illinois, Argonne National Laboratory, Energy and Environmental System Division, Report ANL/EES-TM-199, p. 197. Ulmishek, G., 1985, Geology and petroleum resources of the Barents-northern Kara shelf in light of new geologic data: Argonne, Illinois, Argonne National Laboratory, Energy and Environmental System Division, Report ANL/ES148, p. 89. Van Aarssen, B. G. K., J. K. C. Hessels, and J. W. de Leeuw, 1992, The occurrence of polycyclic sesqui-, tri-, and oligoterpenoids derived from a resinous polymeric cadinene in crude oils from southeast Asia: Geochimica et Cosmochimica Acta, v. 56, p. 30213031. Wilson, J. L., 1975, Carnonate facies in geologic history: Berlin, Springer-Verlag, 471 p. Wingert, W. S., 1992, GCMS analysis of diamondoid hydrocarbons in Smackover petroleum: Fuel, v. 71, p. 3743, doi:10.1016/0016-2361(92)90190-Y. Zakcharov, E. V., and I. B. Kulibakina, 1996, Oil-gas prospects of Triassic complex on shelf of Barents and Pechora seas: Petroleum Geology, v. 31, p. 260265.

1146

Barents Sea and Northern Timan-Pechora

You might also like