You are on page 1of 21

Marine and Petroleum Geology 22 (2005) 457477 www.elsevier.

com/locate/marpetgeo

Signicance of hydrocarbon seepage relative to petroleum generation and entrapment


Michael A. Abrams*
Energy and Geoscience Institute, University of Utah, 423 Wakara Suite 300, Salt Lake City, UT 84108, USA Received 1 June 2003; accepted 31 August 2004

Abstract Near-surface indications of migrating hydrocarbons provide the petroleum systems analyst critical information about source (organic matter type), maturation (organic maturity), migration (migration pathway delineation), and in selected geologic settings, specic prospect hydrocarbon charge. All petroliferous basins exhibit some type of near-surface signal, but the hydrocarbon leakage to surface is not always detectable with conventional seep detection methods. Understanding the Petroleum Seepage System, hence petroleum dynamics of a basin, is key to understanding and using near-surface geochemical methods for basin assessment and prospect evaluation. The relationships between near-surface hydrocarbon seepage and subsurface petroleum generation and entrapment are often complex. The petroleum seepage system contain four key elements: seepage activity (qualitative expressions of relative leakage rates, active vs passive, and episodic vs continuous), seepage type (concentration of migrated thermogenic hydrocarbon relative to in situ material, macro vs micro), migration focus (major direction of bulk ow leakage relative to the subsurface hydrocarbon generation and/or entrapment), and near-surface seep disturbances (near-surface processes which can greatly alter or block seepage signals). The rate and volume of hydrocarbon seepage to the surface greatly control near-surface geological and biological responses, and thus are the best method of sampling and analysis to detect hydrocarbon leakage effectively. To properly predict subsurface petroleum properties, interpretation of near-surface geochemical data must recognize many potential problems including recent organic matter input, transported hydrocarbons, bacterial alteration, mixing, contamination, and fractionation effects. Surface geochemical data should always be integrated with other geological data. Calibration datasets to determine the utility of nearsurface geochemical techniques within particular basinal settings are essential when evaluating prospects for hydrocarbon charge. q 2005 Elsevier Ltd. All rights reserved.
Keywords: Surface geochemistry; Near surface migration; Petroleum systems

1. Introduction Surface geochemistry is commonly called unconventional, although used by the largest oil companies as well as small independents, as a method to explore for oil and gas. Surface geochemistry methods have been used extensively for more than 70 years (Horvitz, 1980, 1981, 1985). So why are surface geochemical methods commonly called unconventional? Contractors and true believers in surface geochemistry demonstrate how effective the surface geochemical tools can be in handouts and publications (Schumacher and LeSchack, 2002), yet many petroleum explorationists have experienced failures.
* Tel.: C1 801 581 8856; fax: C1 801 585 3540. E-mail address: mabrams@egi.utah.edu.

0264-8172/$ - see front matter q 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.marpetgeo.2004.08.003

This perceived unreliability and lack of understanding of the petroleum seepage system contributes to the unconventional classication. A multi-year research study by Abrams (2002b) examined surface geochemical surveys that used a variety of nearsurface geochemical methods, in different geological settings. The study concluded that all petroliferous basins exhibit a near-surface signal, but the signal is not always directly above an accumulation and/or detectable with the conventional seepage detection methods. Also, this surface geochemistry calibration study demonstrates the importance of well-founded interpretations. Many of the failures were due to poor interpretations. Proper interpretation of nearsurface geochemical data requires the recognition of background vs anomalous populations (hydrocarbon concentrations signicantly higher than normal level found in localized near-surface sediments), recent organic matter (ROM, indigenous biological material) input, transported

458

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

hydrocarbons seepage (movement of anomalous hydrocarbons from site of origin to second location via sediment transport), reworked source rock (mature source rock with generated hydrocarbon included in sediment provenance), in situ or post-sampling bacterial alteration (biodegradation), eld or laboratory contamination, and possible migration or sampling fractionation-partitioning effects (Abrams et al., 2001a,b; Abrams, 2002a). The relationship between near-surface hydrocarbon seepage and subsurface petroleum generation and entrapment is often complex. The near-surface expression of hydrocarbon migration varies greatly due to the changes in leakage rates and concentration, major direction of bulk ow, and near-surface processes which alter or block seepage. Rates and volumes of hydrocarbon seepage to the surface control the near-surface geological and biological responses (Roberts et al., 1990) and, thus, the type of sampling required to detect hydrocarbon leakage effectively. Interpreters must rmly grasp these issues to understand signicance of migrated hydrocarbons within near-surface sediments. Surface geochemical measurements provide powerful empirical observations that must be integrated with geological and geophysical data. Near-surface to subsurface geochemical calibration datasets help to evaluate the utility of surface geochemical methods within specic basinal settings and surface sediments conditions.

a structure will be charged. Other active petroleum systems may be present within the exploration area. Trap integrity, reservoir, or migration issues may prevent charge and retention for specic traps. The presence of mature hydrocarbons in near-surface sediments conrms the existence of a mature source rock and migrated hydrocarbons if artifacts such as those caused by recycled sediments can be ruled out. Absence of petroleum seepage within a basin or above a prospect is not sufcient reason to eliminate the possibility that an active source rock is present or that there has been a charge to a specic prospect (with some notable exceptions). 2.2. Prospect evaluation Use of surface geochemical tools does not end when a viable petroleum system has been established. Surface geochemistry can also be used to: 1. delineate hydrocarbon bearing zone at depth when near vertical leakage has been established by calibration surveys; 2. evaluate hydrocarbon charge: gas vs oil by differential leakage-entrapment; 3. evaluate uid quality: reservoir oil gravity and elemental sulfur content; 4. detect presence of non-hydrocarbon gases (CO2 and N2); 5. identify by-passed zones in production settings. Horvitz (1969), Jones and Drozd (1985), Klusman (1993) and Schumacher and LeSchack (2002) document halo and apical near-surface geochemical anomalies above petroleum accumulations. However, not all petroleum accumulations contain a leakage anomaly above or nearby. Bolchert et al. (2000) examined the Green Canyon and Ewing Bank area in Northern Gulf of Mexico, and found no near by macro-seeps near 83% of the elds. The key is in understanding that near vertical migration hydrocarbons in low concentrations, microseepage, does not always occur. Thrasher et al. (1996) and Abrams (1996) demonstrate the importance of understanding the regional and local geology when attempting to evaluate near-surface geochemical anomalies relative to intra basin and/or prospect specic charge systems. Assuming vertical migration always occurs and is measurable may lead to failures. Surface geochemistry should be used as a prospect specic exploration tool only when a calibration surface geochemical dataset has demonstrated near vertical leakage is present in your study area.

2. Survey purpose Deciding whether a surface geochemical survey is appropriate requires understanding the petroleum system elements and processes you are trying to address. Surface geochemical surveys are usually conducted for regional source rock evaluation (conrm presence of a mature source rock) or prospect denition (trap charge). Surface geochemical methods can also be used to evaluate oil quality (Barwise, 1996), oil vs gas (Abrams, 2002a), by-passed oil (Schumacher et al., 1997), and presence of non-hydrocarbon gas (Abrams et al., 2001a,b). 2.1. Regional source rock characterization The presence of seepage, with sufcient amounts of hydrocarbon for detailed molecular characterization, can provide the following key petroleum systems information (Abrams et al., 2001a,b): 1. 2. 3. 4. source type (organic matter type), source age (if age diagnostic biomarkers are present), level of organic maturity (LOM), and primary (from source rock to carrier bed) and secondary (carrier bed to trap) migration pathways.

3. Petroleum seepage system The rates and volumes of hydrocarbon seepages to the surface modies the near-surface geochemical, geophysical, geological, and biological responses. A Petroleum Seepage

The interpreter must remember that favorable key petroleum systems elements and processes do not guarantee

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

459

System (PSS) is dened as the interrelationships among total sediment ll, tectonics (migration pathway), hydrocarbon generation (source and maturation), regional uid ow (pressure regime and hydrodynamics), and near-surface processes (zone of maximum disturbance, Abrams, 1992). The petroleum system analysts must rmly grasp the regional petroleum seepage system to fully appreciate the signicance of anomalous migrated hydrocarbons within near-surface sediments. Relationships among near-surface hydrocarbon seepages and subsurface petroleum generation and entrapment are often complex and their signicance commonly misinterpreted (Abrams, 2002a). Key elements of the petroleum seepage system include. 3.1. Seepage activity Dened as the qualitative expression of relative leakage rates (Abrams, 1989) with no specic relationship to migration mechanism. The seepage activity can be active, prolic ongoing leakage from subsurface to near-surface sediments. The offshore Gulf of Mexico is a good example of a region where many areas have experienced active seepage both at the present time and in the past (MacDonald et al., 1996). Seepage activity can also be passive, slow subtle leakage from subsurface to nearsurface sediments. The offshore Navarin and Saint George Basins (Bering Sea, Alaska) are good examples of passive seepage (Abrams, 1992). In areas of active seepage, hydrocarbon movement to the near surface is not always continuous, but can be episodic (Roberts and Carney, 1997; Quigley et al., 1999; MacDonald et al., 2000; Abrams and Boettcher, 2000). 3.2. Seepage type Dened as the concentration of migrating thermogenic hydrocarbon relative to in situ material. The in situ material includes recent organic matter (ROM) derived from pelagic, or reworked material from land based or subcrop sources. The common terms used to dene seepage type are macroand micro-seepage (Abrams, 1992). Macroseepage usually refers to large concentrations of migrating hydrocarbons, which are generally visible and related to bulk ow (Darcy ow). Macroseepage concentrations are generally in excess of 100,000 ppm (by volume) of total gas and 1000 ppm (by volume) of hydrocarbon sediment extract. Microseepage refers to low concentrations of migrating hydrocarbons, not visiblebut detectablewith standard analytical procedures. Migration mechanism commonly proposed for microseepage include buoyancy of micro-bubbles (Price, 1986; Klusman and Saeed, 1996; Saunders et al., 1999; Brown, 2000). Microseepage concentrations are generally less than 10,000 ppm (by volume) of total gas and 100 ppm (by volume) of hydrocarbon sediment extract.

3.3. Migration focus Dened as the major direction of near-surface leakage relative to the subsurface hydrocarbon generation and/or entrapment. The migration focus direction can be nearly vertical when there are major migration pathways such as faults and diapirs, or, lateral displacement related to basin uid ow dynamics (Thrasher et al., 1996). Schumacher and LeSchack (2002) argue that despite major lateral migration ow related to regional seals and key carrier beds, hydrocarbons also leak vertically via the buoyancy of micro-bubbles. 3.4. Near-surface seep disturbances Dened as near-surface processes (physical and biological) which alter or block the petroleum seepage signals. The Zone of Maximum Disturbance, known as ZMD (Abrams, 1992), is a shallow near-surface zone where pore water ushing; partitioning of migrated hydrocarbons between vapor, solute, and sorbed phases; bacterial alteration; and in situ hydrocarbon generation have altered the migrating themogenic hydrocarbon signatures beyond recognition. Additionally, shallow migration barriers such as hydrates, permafrost, and cohesive shales can partially block the migrating hydrocarbons.

4. Intepretation surface geochemical surveys 4.1. Dening background and anomalous All land or marine sediments contain some background level of light, low molecular weight (LMW) and heavier high molecular weight hydrocarbons (HMW). Identifying background vs anomalous signatures for near-surface hydrocarbon measurements can be complicated. An anomalous population is dened as a group of samples with total hydrocarbon concentrations signicantly above the established background. Surface geochemical measurements rarely follow a normal distribution but tend to be skewed, log normal distribution (Fig. 1a). The application of normal distribution descriptors such as mean, standard deviation, and variance have no statistical validity in a log normal surface geochemical dataset. Graphical data analysis provides a simple and visual process to evaluate sample distribution and assist in the identication of multiple populations. The two most common graphical methods best suited for surface geochemical datasets include frequency histograms and cumulative frequency. Both methods provide the interpreter with a quantitative method to identify the presence of an anomalous population. 4.1.1. Frequency histogram The histogram, or frequency distribution, separates the data values into bins, shown on the x-axis (Fig. 1a). Number

460

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

P Fig. 1. Population distribution of sediment headspace gas ( C1C5) from sediment cores collected within leakage zones above an offshore Gulf of Mexico eld. (a) Histogram. (b) Cumulative frequency plot.

of samples within each bin (frequency class or interval) is represented on the y-axis. Rectangles are constructed over each interval with the height being proportional to the number of measurements (class frequency) falling within each bin. Histograms are useful for depicting sample symmetry or skewness. The histogram shape is very dependent on the number of categories selected. 4.1.2. Cumulative frequency A cumulative frequency, also known as quantile plot, graphs cumulative frequency percent on the x-axis and the measured geochemical parameter on the y-axis (Fig. 1b). Three advantages of using a cumulative frequency plot over the histogram plot are arbitrary categories (bins) are

required, all the data are displayed, and every point has a distinctive position without overlap. Fig. 1a displays a histogram and Fig. 1b a cumulative frequency graphical plot for 36 core samples collected within leakage zones above a eld in offshore Gulf of Mexico. Sediment samples with headspace gas concentrations less than 200 ppm by volume are considered to be background (29 samples). Whereas sediment samples with headspace gas concentrations greater than 100,000 ppm by volume are considered to be anomalous (7 samples). In this example, the separation between background and anomalous is relatively straightforward. The difculties begin when the separation of background and anomalous is not as pronounced due to mixing, sampling problems, reworking

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

461

source rocks, transported hydrocarbon, fractionation, or in situ alteration-generation. Most interpreters assume the anomalous samples are non-indigenous (migrated) thermogenic hydrocarbons. However, some anomalous populations may reect indigenous or syngenetic hydrocarbons from bacterial activity or an artifact of differences in sediment type (lithology), sampling depths, and/or sampling times. 4.2. Determining origin of anomalous hydrocarbons 4.2.1. Light hydrocarbons (gas) The molecular characteristics of surface sediment gases vary with the type of gas present in the sediment (bacterial vs thermogenic), as well as with gas extraction method (Horvitz, 1985; Abrams, 1996; Bjoroy and Ferriday, 2002). This section concentrates on gas parameters commonly used to evaluate anomalous hydrocarbon source: bacterial, thermogenic, or mixture. Analytical method variability will be discussed in the Problems and Pitfalls Analytical Methods section. Conventional interpretation parameters developed for reservoired hydrocarbon gas samples are not as effective with surface geochemical screening datasets. Unfortunately, hydrocarbon seepage found in many marine surveys are fractionated (in situ, during migration, or sampling), partitioned (based on the properties of the hydrocarbon compound and physical environment), altered (bacterial), and/or mixed rendering conventional reservoired interpretation schemes relatively useless in most areas with low levels of seepage. Natural gases are characterized using three analyses: gas composition (C1C5 and non-hydrocarbon components such as CO2, O2, and N2), compound specic carbon isotopic ratio (d13Cn), and hydrogen isotopic ratio (dDCH4) (Schoell, 1983). Gas compositions and isotopic ratios depend on type and maturity of source. Alterations during migration or bacterial activity and mixing will also affect the gas
Table 1 Gas interpretation parameters commonly used for near-surface sediment gases Parameter Total HC gas: P C1C5 P C2 K C5 = P C1 K C5 ! 100 Information provided

compositions and isotopic ratios (Abrams, 1989). Table 1 lists the gas compositions and isotopic ratios most commonly used to evaluate sediment gases. The relative amounts of methane, ethane, propane, butane, and propane is the rst clue to origin. Methane may be derived from either thermogenic or bacterial processes. The wet gases (ethane, propane, butane, and pentane) are believed to be derived from only thermogenic sources. Studies over the years, both in the laboratory as well as empirical observations, indicate traces of ethane, propane, and a number of other light hydrocarbons can also be formed microbiologically (Hunt et al., 1980; also see Davis and Squires, 1954; Oremland, 1981; Oremland et al., 1988). Recent work, including compound specic isotopic studies, have shown that traces of a number of light hydrocarbons are produced by microbiological processes and a number of the organisms responsible have been identied (Whiticar, 1999). In spite of this recent work, this concept is still controversial within the petroleum geochemical community. What is agreed is samples with elevated total gas concentrations, and a wet gas fraction P P greater than 5.0% C2 C C3 = C1 K C3 ! 100, are most likely derived from a thermogenic process (Bernard, 1978; James, 1983;P Schoell, 1983). Thus, the gas wetness P ratio C2 K C5 = C1 K C5 ! 100, which includes not only ethane and propane, but butane and pentane, is a common parameter to help evaluate bacterial vs thermogenic. Ethene (ethylene) and propene (propylene) belong to a class of hydrocarbons known as olens. They contain one double bound and therefore are unsaturated with respect to hydrogen, and are almost always found in trace amounts in surface sediments (Whelan et al., 1988). These compounds appear to be rapidly hydrogenated via near-surface anaerobic microbial processes and so are not detected in measurable amounts in the vast majority of reservoired gases. These compounds are derived primarily from bacterial processes, not from conventional thermogenic

Interpretation process Histogram to identify background vs population Identify samples with elevated wet gas fraction and anomalous total HC gas Identify samples with elevated % wet gas and anomalous total HC gas Identify samples with anomalous total HC gas and elevated ethane/ethene and propane/propene ratios Identify samples with anomalous total HC gas and methane carbon isotopes and/or wet gas isotopic separations indicative of thermogenic origin Identify samples with anomalous total HC gas, determine C1 hydrogen isotopes for anomalous samples, and plot on d13C1 and dDCH4 crossplot

Gas wet percent % Methane C1 =

C1 C5 ! 100

Ethane/ethene and propane/propene Carbon isotopes: d13C1, Dd13C1 and d13C2, Dd13C2 and d13C3 Hydrogen isotope: dDCH4

Identify anomalous (above background) total HC gas Evaluate thermogenic (non-bacterial) contribution Evaluate thermogenic (non-bacterial) contribution Saturate (thermogenic) vs unsaturate (bacterial) Thermogenic vs bacterial

Thermogenic vs bacterial and type bacterial (fermentation vs carbonate reduction)

462

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

and catagenic reactions (Bernard et al., 2002; Ullom, 1988). Thus, the ratio of ethane (thermogenic) to ethene (bacterial) provides information on the gas origin. Bernard et al. (2002) used the ethane/ethene ratio, in combination with the isotopic ratio of methane, to evaluate the thermogenic contribution. One must use this ratio, as well as the propane/propene ratio, with caution because olens are more readily altered than alkanes in subsurface processes. Compound-specic isotopic ratios numbers have evolved as an important tool in surface geochemistry, with the IR-GC/MS (continuous ow isotope ratio gas chromatography/mass spectrometry). IR-GC/MS allow for isotopic measurements in sediments with relatively small concentrations of gas. Early studies relied only on methane carbon isotopes (Horvitz, 1981; Faber and Stahl, 1983; Abrams, 1989) due to equipment limitations. Thermogenic methane is enriched in 13C compared with bacterial derived methane, with values ranging from K50 to K20. The variation in ethane, propane, and butane carbon isotopic ratios, as well as hydrogen isotopic ratios, are additional ways to evaluate sediment gas origin and possible secondary fractionation (post-generation). The alteration story is much more complex than originally thought because of anaerobic microbial degradation of methane, which causes the isotopic ratio to get heavier (less negative). Hydrogen isotopes provide additional information which can help sort out a complex history (Whiticar, 1999). Compound-specic isotopic analysis (CSIA) also helps biochemists understand bacterial formation of methane. Methane generated in shallow marine sediments from methyl type fermentation, such as acetate reduction (shown below), has different carbon and hydrogen isotopic signatures than methane formed by CO2 (reduction): Carbonate reduction : CO2 C 8HC0 CH4 C 2H2 O Acetate fermentation : CH3 COOH 0 CH4 C CO2 In most marine sediments, sulfate rich zones curtail methanogenesis. Non-methanogens (sulfate reducing bacteria) metabolize available labile carbon. Methanogensis start using competitive substrates once available dissolved sulphate is exhausted (sulfate reduction zone). This is generally at a depth of 1 to 4 meters in most marine sediments. Carbonate reduction becomes the dominant methanogenic pathway under these conditions because methanogenic substrates such as acetate are depleted, and bicarbonate is availabile. Refer to Whiticar (1999) for greater detail on the systematics of bacterial formation and oxidation of methane. Oremlands and Whiticars work also show that other methyl precursors are possible: methylamine, methanol, methyl suldes in specic environments (Oremland et al., 1988). Chemotrophic methane oxidation is an important process that appears to be very widespread in anaerobic environments, which could include oil seeps and reservoirs (Larter et al., 2000). Unfortunately, the process

dramatically affects methane carbon isotopes, making them heavier in 13C, so that resulting values from this process overlap with the thermogenic petroleum range. It is important to understand this process because carbonate reduced methane is isotopically lighter than methane formed by fermentation. Bernards (1978) interpretation scheme for sediment gases conclude that methane with an isotopic ratio lighter than K60 would be derived from a bacterial source. A 50:50 mix of thermogenic with a methane isotopic ratio of K40, and carbonate reduced gas with methane isotopic ratio of K110 would result in a gas with a methane isotopic ratio of K75. Even a gas which is 80% thermogenic would result in a mixed gas with a methane isotopic ratio of K64 (Table 2). Based on the isotopic ratio of methane alone, this would be classied as bacterial using most interpretation charts. The presence of elevated wet gas concentrations (wet gas ratio), and ethane and propane isotopic ratios (if available), should provide clues that these mixed gases have a thermogenic component. Thus, understanding that carbonate reduced gases are very isotopically light and do mix with migrating thermogenic gases is important when evaluating gases extracted from marine sediments. Claypool makes the important point in one of his early papers that the way to predict the microbial reduction of CO2 to CH4 is to look at the differences in delta 13C between CO2 and methane in the system, which always gives a constant value if anaerobic reduction of CO2 is involved (Claypool and Kaplan, 1974). Secondary alteration such as anaerobic bacterial activity produces methane enriched in 13C. Resulting methane has an isotopic ratio very similar to thermogenic derived methane (Abrams, 1989, 1996). Bacterial consumption of methane proceeds signicantly faster than does that of ethane, propane, and butane (Whiticar, 1999). The result is a gas with elevated gas wetness. The ethane, propane, and butane carbon isotopic ratios (d13Cn) assist in evaluating thermogenic contribution. Because the ethane plus hydrocarbons are derived primarily from thermogenic sources, the isotopic separations between ethane and propane, and propane and butane, are strongly dependent on maturity level, when there is little or no secondary alteration (James, 1983). The hydrogen isotopic ratios (dDCH4) differentiate bacterial methane gas from thermogenic. Methanogens derive a signicant proportion of their hydrogen from interstitial water during methane formation by carbonate reduction (Whiticar, 1999). The dDCH4 for methane is
Table 2 Isotopic ratios for mixes of thermogenic and bacterial gases Thermogenic d13C1ZK40 80% 50% 40% Bacterial d13C1ZK110 20% 50% 60% Resulting mix d13C1 K64 K75 K82

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

463

derived from bacterial carbonate reduction range from K150 to K250 (SMOW). The above gas composition and compound-specic isotopic ratios are best used only for samples with elevated concentrations. Samples with low concentration levels (background) may not be representative of migrated gases. Background (low concentration) samples are subject to greater variation in composition among gas components due to various alteration and fractionation processes. Headspace extracted gas from 70 Gulf of Mexico seabed cores collected above or near a subsurface petroleum accumulation display four groups of samples (Fig. 2): Group 1 Background: total gas concentrations less than 1000 ppm by volume and wet gas fraction less than 0.05 (5.0%). Group 2 Fractionated: total gas concentrations less than 1000 ppm by volume and wet gas fraction greater than 0.05, up to 0.18 (18%). Group 3 Bacterial: total gas concentrations greater than 1000 ppm by volume and wet gas fraction less than 0.05 (5.0%). Group 4 Thermogenic: total gas concentrations greater than 1000 ppm by volume and wet gas fraction greater than 0.10 (10%). Group 1 and 2 provide no information on migrated thermogenic gases. Group 3 is most likely in situ derived bacterial gas. Group 4 appears to be migrated thermogenic gases (based on the elevated total gas concentration and wet gas fraction). Compound-specic carbon isotopic ratios from Group 4 samples conrm these samples are derived from migrating thermogenic gases. In conclusion, care must be taken when working with low concentration samples. They may appear to be thermogenic

based on wet gas percents and/or methane isotopic ratios. However, when only traces of gases are present, small alterations in one or more of the compounds present can produce very misleading results. The problem is amplied many times in depending on ratios involving very small values, as occurs in analyzing samples with only traces of hydrocarbons. 4.2.2. High molecular weight hydrocarbons (C15 plus) The most common screening procedures currently used for evaluating the presence of thermogenic high molecular weight (HMW) hydrocarbons (C12 plus), include whole extract gas chromatography (GC) and total scanning uorescence (TSF). A dried sediment sample is ground to a uniform size by weight, extracted with an organic solvent (Soxhlet or ASE, accelerated solvent extractor), and lastly concentrated. Many surface geochemical laboratories currently use low polarity solvents such as hexane. Low polarity solvents extract only low polarity compounds. Another approach would be to use a mixture of varying range of polarity solvents to extract all components and then separate the petroleum related compounds (saturated hydrocarbons, unresolved complex mixture (UCM), aromatics, and polars such as NSO and asphaltenes) from the ROM (saturated hydrocarbons, ketones, alcohols, and fatty acids) using multi-component silica gel column chromatography. 4.2.2.1. Whole extract gas chromatography. Sediments containing moderate levels of upward-migrating thermogenic high molecular weight (HMW) hydrocarbons are characterized by an UCM, discernible C 15C 32 n-alkanes and isoprenoids, and an overprint of odd nalkanes greater than C23 from terrigenous plant biowaxes.

Fig. 2. Total headspace extracted sediment gas ( above an offshore Gulf of Mexico eld.

P P P C1C4) vs wet gas fraction C2 K C4 = C1 K C4 from sediment cores collected within leakage zones

464

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

Table 3 Interpretation parameters for sediment whole extract gas chromatography (GC) Parameter Total UCM (mg/g; unresolved complex mixture) UCM (mg/g)!C23 UCM (mg/g)OC23 n-Alkanes (ng/g) !C23 n-Alkanes (ng/g) OC23 Total EOM ppm by weight (extractable OM) Provides information on Quantication of extractable hydrocarbons not resolvable in gas chromatography (mainly NSO and asphaltene compounds) UCM representative of migrated thermogenic portion UCM representative of recent organic matter portion n-Alkanes representative of migrated thermogenic portion n-Alkanes representative of ROM portion Quantication of total extractable HC

Samples containing elevated bitumens often are extensively biodegraded containing only a UCM (Brooks and Carey, 1986). Whole extract gas chromatograms can be subdivided into several major groups: Resolvable peaks: The total hydrocarbon fraction from a non-degraded petroleum seep is usually dominated by nalkanes, with a lesser amount of branched alkanes (including isoprenoids) as well as some cyclic alkanes, and alkyaromatics. Unresolved complex mixture (UCM): The unresolved complex mixture, otherwise known as UCM and naphthelene hump, is a quantication of unresolvable hydrocarbon and non-hydrocarbon compounds. Recent organic matter (ROM): Extract gas chromatograms from recent sediments generally display an odd carbon preference within the n-C25 and n-C33 range due to the elevated contribution from recent plant waxes. Key parameters commonly used to evaluate sediment extract chromatograms for the presence of migrated themogenic hydrocarbons include: total extractable material (EOM, total concentration of solvent extractable material in ng/g), total unresolved complex mixture (UCM: total amount of unresolved material in mg/g by weight), unresolved complex mixture greater or less than C23 (relative amounts of OC23 ROM vs !C23 migrated hydrocarbons), and total alkanes (total amount of alkanes greater than n-C15 in ng/g) (Table 3). 4.2.2.2. Fluorescence spectrometry. Total scanning uorescence (TSF) detects and measures organic compounds containing one or more aromatic rings. Oil seepages have distinctive uorescence ngerprints because they contain petroleum related compounds with one or more aromatic rings and their alky homologues. A solution of sediment extract is irradiated with light from about 250 to 500 nm at 10 nm intervals. The uorescence emissions spectrum is recorded for each excitation wavelength again scanning
Table 4 Interpretation parameters for sediment extract total scanning uorescence (TSF) Parameter MFI (units) Max_Em and Max_Ex R1: 360/320 nm at 270 nm Type of equipment Dilution Provides information on

from about 250 to 500 nm building a 3D spectrum (Brooks et al., 1983). The emissions maximum uorescence intensity (MFI) is recorded along with emissions wavelength (Max_Em) and excitation wavelength (Max_Ex) (Table 4). A second parameter commonly used to evaluate TSF data is the R1 ratio. R1 is the ratio of emissions at 360 nm compared to emissions at 320 nm when excitation at 270 nm is used (Table 4). This ratio characterizes the shape of uorescence spectra which can be related to API gravity using an empirical relationship derived from a calibration set (Barwise and Hay, 1996). Samples with relatively large concentrations of extract may require dilution. If so, measured MFI (maximum uorescence intensity) are adjusted by multiplying measured MFI by the dilution factor to obtain a corrected MFI: Corrected MFI Z Measured MFI ! Dilution factor 4.2.2.3. Gas chromatography/mass spectrometry. When anomalous high molecular weight hydrocarbons are found with the screening procedures, further molecular characterization is helpful. GC/MS, or gas chromatography/mass spectrometry provides detailed molecular information on biological markers. Biological markers are chemical compounds in the reservoired oils and sediment extracts with the basic molecular structure which can be linked to a known biological precursor. Different organic source facies contain different assemblages of organisms (bacteria, algae, marine algae, and higher plants). GC/MS biomarker data, in conjunction with non-biomarker parameters, resolve the organic source facies depositional environment, as well as level of thermal maturity. Key biomarker compounds are measured in oils and seep extracts, therefore, providing a method to correlate surface seep to subsurface oils and/or source rocks (Hunt, 1996; Peters and Moldowan, 1993).

(1) Magnitude/level of seepage (macro vs micro), (2) type of seepage (oil, condensate, recent organic matter) (1) Type of seepage (oil, condensate, recent organic matter), (2) API gravity when calibrated (see Barwise et al., 1996) Changes in MFI values due to equipment type Dilution factor used to correct MFI

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

465

4.3. Integration of LMW and HMW geochemical data Interpretation of surface geochemical data to evaluate subsurface hydrocarbon generation and migration rst requires that the gas (low molecular weight or LMW) and liquid (high molecular weight or HMW) hydrocarbon data be integrated. Not all analytical procedures provide similar results. The differences provide a way to classify anomalous hydrocarbons detected in near-surface sediments. There are ve general thermogenic signatures of LMW and HMW geochemical data: Active (fresh) migrated thermogenic oil. Sediment samples have elevated total hydrocarbon gas (anomalous) with thermogenic gas signatures (elevated gas wetness and thermogenic d13Cn) as well as elevated high extracted HMW hydrocarbons with a relatively undegraded thermogenic gas chromatogram and biomarker signatures. Relict (passive) migrated thermogenic oil. Sediment samples contain background total hydrocarbon gas with anomalously high extracted HMW hydrocarbons. The gas chromatogram is severely degraded with only an elevated UCM. In addition, the biomarker data displays strong indications of degradation with only the more resistant thermogenic compounds present. Relict migrated thermogenic oil with possible recharge. This is similar to the Relict (passive) Migrated Thermogenic Oil signature but with an addition of resolvable compounds in the C12C20 range on the gas chromatogram and above background thermogenic gas (elevated gas wetness and thermogenic d13Cn). These features indicate a relict degraded seep has been recharged with recent seepage. The fourth and fth thermogenic signatures, transported and reworked, do not indicate local seepage. They are discussed in detail in Section 4.45. 4.4. Pitfalls and problems 4.4.1. Recent organic matter interference Surface sediments contain ROM derived from rock fragments with organic content and/or biological remains unrelated to hydrocarbons migrating from depth. The type of in situ extractable organic material present in the sediment sample will be dependent on the origin (provenance) and local biological setting. The indiscriminate extraction process dissolves all organic matter, including both the migrated seep hydrocarbons from subsurface reservoirs or mature source rocks, and in situ organic materials. The presence of ROM may mask and/or modify peaks on the extract gas chromatograms (GC) and gas chromatography-mass spectrometry (GC/MS) fragmentograms, and alter uorescence spectrometry results relative to migrated thermogenic hydrocarbons from depth. Whole extract GC, TSF, and GC/MS display predictable changes depending on the relative

amount of recent organic matter and migrated thermogenic hydrocarbons (derived from the thermal breakdown of organic matter) (Abrams et al., 2001a,b). 4.4.1.1. Gas chromatograms. Extract GC for surface sediment cores dominated by ROM display abundant unsaturated compounds, low isoprenoids, and C23C35 nalkanes with a predominance of odd-over-even carbon numbers (Fig. 3a). By contrast, extract GC for cores dominated by thermogenic hydrocarbons display abundant saturate and isoprenoid peaks, raised baseline from UCM, and C23C35 n-alkanes with some predominance of even over odd carbon numbers (Fig. 3b). 4.4.1.2. Total scanning uorescence. TSF for surface sediment cores with a thermogenic oil signature display maximum uorescence intensity excitation (Max_Ex) and emission wavelengths within the thermogenic petroleum range; excitation between 280 and 330 nm and emission between 380 and 400 nm (Fig. 4a) By contrast TSF uorograms for cores with a ROM signature display maximum uorescence intensity excitation (Max_Ex) and emission (Max_Em) wavelengths within the perlylene range; excitation 330 C nm and emission 400 C nm (Fig. 4b). 4.4.1.3. Gas chromatography-mass spectrometry. The interpretations of biomarker data from surface sediment seepages requires a different approach than conventional oiloil and oil-source correlation. Because ROM interferes and biodegradation is common, most conventional biomarker ratios do deviate from true migrated value as the relative amounts of in situ ROM increase relative to the migrated hydrocarbon (Fig. 5). The key is not to use the conventional biomarker ratios found in publications such as Peters and Moldowans (1993) Biomarker Guide but instead use lesser known, yet previously established biomarker parameters (diasterane equivalents and aromatic parameters) and substitute traditional biomarker ratios with novel ratios: extended tricyclics ratio, gammacerane to diahopane index, and oleanane to diahopane index (Holba and Huizinga, 2002). These novel ratios are less susceptible to interferences by recent organic matter and biodegradation than found in the conventional ratios. 4.4.2. Identifying transported and reworked thermogenic hydrocarbons Thermogenic HMW hydrocarbons extracted from shallow sediments may not be derived from local seepages but could be derived from materials within the sediment provenance (reworked mature source rock) or carried by displaced sediments which contain migrated mature hydrocarbons (transported).

466

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

Fig. 3. Solvent extract GC from a Gulf of Mexico seabed geochemical survey: (a) migrated thermogenic hydrocarbon dominance, (b) recent organic matter dominance.

4.4.3. Reworked signature Recently deposited thermally mature source rock derived from near-by uplifted and eroded sediment provenances can be confused with localized migrated hydrocarbon seepage (Piggott and Abrams, 1996). Key geochemical characteristics which indicate reworked mature source rock may be present includes strong thermogenic signal with little or no ROM character; extract GC with a full compliment of normal parafns (no evidence of bacterial alteration); relatively low levels of high molecular weight hydrocarbon extract (solvent extract less than 100 mg/g extract); little or

no associated sediment gas, elevated total organic carbon; thermogenic seep signatures present in more than 30% of cores; and cores with a thermogenic signature within and away from migration pathways zones. If this geochemical signature is present, the following additional information will provide conrmation that a reworked source rock is present: biostratigraphic evaluation of core samples to look for palynological and paleontological evidence of reworked detrital kerogen; Rock-eval pyrolysis and pyrolysis-GC (elevated S1 free hydrocarbons), geologic maps with mature source outcrops present in study area; and comparison of

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

467

Fig. 4. Solvent extract TSF uorograms from a Gulf of Mexico seabed geochemical survey: (a) migrated seep hydrocarbons dominance, (b) recent organic matter dominance.

molecular characteristics of seep to local provenance source rock outcrop. 4.4.4. Transported signature Near-surface thermogenic high molecular weight hydrocarbons derived from subsurface leakage can be carried along with displaced sediments and transported to locations downdip. The displaced thermogenic hydrocarbons will contain relatively low concentrations of thermogenic hydrocarbons relative to localized hydrocarbon seepage. A case history of the eastern deepwater Gulf of Mexico by Cole et al. (2001) found two anomalous populations of hydrocarbons (above general background based on multiple surveys in study area): Group 1 Sediment cores with very large concentrations of thermogenic hydrocarbons and migrated mature hydrocarbon signal much greater than in situ ROM; Group 2 Sediment cores with low concentrations of thermogenic hydrocarbons and in situ ROM signal greater than thermogenic. The organic facies based on biomarker results for Group 1 (macroseepage) match the subsurface hydrocarbons based on drilling results. The organic facies for Group 2 (micro seepage) did not match the subsurface hydrocarbons

but was similar to the organic facies found in shelfreservoired hydrocarbons. Shallow high resolution seismic proles indicate slope unstability and the movement of sediments from updip locations to the transported signal area. Furthermore, uid ow models strongly suggested the Group 2 anomalous seabed cores are unrelated to the subsurface petroleum system, where as the migration ow paths only go to the Group 1 seep sites (Cole et al., 2001a,b). Thus, the interpretation of transported hydrocarbons is not

Fig. 5. Changes in conventional biomarker ratios with increased ROM input relative to migrated thermogenic hydrocarbons (adapted from Abrams and Boettcher, 2000).

468

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

only based on geochemical data, but an evaluation of sediment transport and migration pathway analysis. The identication of transported thermogenic hydrocarbons is subtleand at times difcult to interpretbut extremely important. Several ways to identify transported hydrocarbon seepage include: 4.4.4.1. Seepage magnitude. Transported hydrocarbon seepage will have lower concentrations of hydrocarbons relative to in situ migrated hydrocarbon seepage. Cole et al. (2001) have recognized a group of samples called low condence based on an extensive Gulf of Mexico database. The low condence cores contain MFI values from TSF analysis less than 30,000 units and UCM values from GC analysis less than 100 mg/g, and have been shown by biomarker analysis to be transported not in situ seepage. These cut-off values will vary for different seepage systems (seepage activity and type). 4.4.4.2. Location seepage. Transported hydrocarbons will be present in areas away from major migration pathways as well as within migration zones. Thus, sampling programs should contain some core locations away from potential migration pathways and within areas of major sediment transport. 4.4.4.3. Seepage activity. Transported hydrocarbons are most likely to occur in basins with prolic active hydrocarbon seepage (Abrams et al., 2001a,b). Examine geophysical, geological, and geochemical data to document seepage activity and assist in evaluating transported hydrocarbon seepage risk. 4.4.4.4. Variation in seepage with core depth. Compare geochemical signal from different parts of the core. Do the thermogenic hydrocarbon anomalies correspond to specic depositional packages? Transported hydrocarbons should correspond to specic sediment packages re-deposited by major uvial and/or slope failure systems.

4.4.5. Variation in anomalous signature due to different analytical procedures Multiple analytical procedures have been developed to remove migrated gases from sediments (Abrams, 1996; Abrams et al., 2001a,b). The terminology used for each sediment gas extraction method generally refers to the physical removal process and not the phase. Migrated gases are either in the interstitial pore spaces as a free or dissolved phase, bound to mineral or organic surfaces, or entrapped in crystal inclusions (Fig. 6). The exact nature (physical binding state) of the gases removed by each procedure is still poorly known; therefore, these procedures should be considered to be operational denitions and not representative the actual in situ physical state of the gases in the sediment. Consequently, it is very important, in comparing results from different laboratories or times, to make sure that the same experimental procedures were used. The removal procedures range from simple shaking, mechanical break-up, vacuum, and chemical (acid treatment of sediment). Interstitial gases can be sampled by either non-mechanical or mechanical methods: Non-mechanical: The headspace gas collects interstitial gases which can be released by simple shaking. The sample can contains an aliquot of sediment, degassed distilled or ltered seawater, and air or inert gas (helium or nitrogen). The can is vigorously shaken using a paint shaker and heated prior to sampling (laboratory dependent). The interstitial gases within sediment pore spaces move to the headspace within the can which is then sampled through silicone septum on the top of specially modied can (Bernard, 1978). Mechanical. The ball mill gas extraction method utilizes a steel ball within a stainless steal container to mechanically break up a measured aliquot of unconsolidated sediments. The container is vigorously shaken. The steel ball pulverizes sediment sample, releasing gases which are collected from the container headspace through a septum (Bjoroy and Ferriday, 2002).

Phase
Non-mechanical Non-mechanical Non-mech anica l

Extraction
Headspace (shake can)

Interstitial gas Interstitial g Interstitial g as as


interstitial pore gas as dissolved or free phase

Ball & mill (ball & vessel)


Mechani call Mech anica
1 1 Mechanical1

Blender (rotating blade) Disrupter (fixed blade)

3 3 Acid Acid extraction Aci dextraction ext ractio n 3

Sorbed (acid-vacuum)

Bound gas Bound gas Bound gas


bound to mi mineral and organic surfaces; s s; or entrapped in crystal carbonate inclusions in
44 Vacuu m desorp desorp tion desorption Vacuum tion 4

Adsorbed (vacuum)
1 physical break up mechanism 2 also called occluded 3 Horvitz method 4 See Zhang(2003)

Fig. 6. Sediment gas extraction procedures.

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

469

The blender gas method, also known as loosely bound or cuttings, utilizes a blender to mechanically break up aliquot of sediment and release interstitial gases within unconsolidated sediments. The released gases are sampled through a septum on the blender cap (Abrams, 1996). The disrupter gas method uses a xed blade to break up sediment within a sealed chamber. The sediment sample is moved through xed blades by vigorous unidirectional shaking. The released interstitial gases are sampled through a septum on the top of disrupter chamber (Abrams, 2004). Bound gases can be sampled by two basic methods: The acid extractionalso known as Horvitz, adsorbed, sorbed, and boundmethod captures gas bound to the ne grain sediments (Horvitz, 1981), captured within authigenic carbonate (Abrams, 1996), or bound by structured water mineral surfaces (Whiticar, 2002). The coarse-grained fraction (greater than 63 mm) is removed by wet sieving a bulk sediment sample. The ne-grained portion (63 mm and smaller) is heated in phosphoric acid in a partial vacuum to remove bound hydrocarbons. A modied version, called called microdesorption, of the original Horvitz methodology is used by Whiticar (2002). Vacuum desorption also removes the bound gases and is similar to the acid extraction method but does not include the addition of acid (Zhang, 2003). Horvitz (1981) reported that different gas sediment extraction methods (shaking-headspace, blender, and acid extraction) provide different results on replicate samples. Studies by the author in the early 1980s provided similar results (Fig. 7a and b) (Abrams, 1989). Bjoroy and Ferriday (2002) data support similar conclusions. Horvitz (1981, 1985) and Bjoroy and Ferriday (2002) both concluded the adsorbed sediment gas extraction methods were superior since resulting sediment extracted gases contain higher wet gas concentrations. Bjoroy and Ferriday (2002) concluded that the ball mill was also superior since the resulting gases contain higher wet gas concentrations. Without any comparison to the migrated gas composition, it is impossible to verify their conclusions. One needs to compare the sediment and reservoir gas composition in order to make such a conclusion. Abrams (1989) did compare the sediment extracted methane gas isotopic composition from two different sediment extraction methods (adsorbed and headspace) to the reservoir gases in a leaky system (bulk ow via a leaky fault). Abrams concluded the adsorbed extracted sediment gases compared well with the reservoir gases (Fig. 7b). Additional studies are currently underway by the author to evaluate the various sediment gas extraction methods under controlled laboratory conditions (Abrams, 2002b). In the authors opinion, the only statement which can be made with condence is that different sediment extraction methods do provide different gas compositions and compound specic isotopic ratios. To determine which method best represents the in situ gas composition and isotopic ratio requires additional study. It is almost impossible with any

Fig. 7. Comparison of gas composition and isotopic compositions using different gas sediment extraction methods: (a) headspace and blender hydrocarbon gas compositions from replicate samples (adapted from Abrams, 1992), (b) adsorbed and headspace methane carbon isotopic (d13C1) compositions from replicate samples (adapted from Abrams, 1989).

geochemical extraction procedure to ever know the absolute amount or state of compounds extracted. The best that can be hoped for is to obtain consistent results from sample to sample, so comparisons are possible. More evaluation to relate extraction procedures to in situ physical state of hydrocarbons recovered is required. 4.4.6. Variation in anomalous gas signature due to nearsurface sediment type Sediment gas concentrations and compositions will vary by sediment type (grain size and composition). Partitioning of migrating gases in near-surface sediments depends on many factors: migration phase (vapor or solute), gas characteristics (solubility, Henrys constant, and sorption kinetics), sediment characteristics (grain size, type minerals, organic matter content, and type of organic matter), and sediment gas extraction method (headspace, blender, mechanical, or acid extraction).

470

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

Surface geochemical surveys rarely evaluate the sediment characteristics and assume different distributions of near-surface sediment gases reect only the presence of migrated thermogenic gases. In cases of very large-volume seepage (macro), the effect of sediment is most likely second order, and has minimal effect on nal interpretation. But in areas of lower-volume seepage, the variation in total sediment gas concentrations and gas compositions are greatly affected by the type of analysis and sediment type. The headspace and blender methods examine free or interstitial gases. Abrams (1989) demonstrated the variability of free (interstitial) sediment gases by sediment size (percent sand, fraction greater than 63 mm). Horvitz (1980) noted the effect of sediment grain size on ethane plus adsorbed hydrocarbons in the Gulf of Mexico seabed coring surveys. He concluded that the differences were the result of highly adsorptive clays present in the surface sediments. Adsorbed (acid extraction) gases may also vary by composition (mineralogy). Shallow sediment samples were collected using a grid survey over two onshore elds during a recent calibration study (Abrams, 2002b). The Area 1 samples contain adsorbed gas concentrations between 10 and 100,000 ppm with one value around 300,000 ppm. Area 2 contain samples with adsorbed gas concentrations from 100,000 to 300,000 ppm with one value around 700,000 ppm (Fig. 8). The contractor report indicated Area 2 is above a petroleum bearing reservoir whereas Area 1 was not. In reality both sets of samples were collected above petroleum bearing traps. Area 1 has CaCO3 less than 30% and Area 2 has CaCO3 greater than 30%. The differences in gas volumes using the adsorbed sediment gas extraction method appears to be unrelated to presence of hydrocarbon accumulation but controlled by CaCO3. This is not a unique observation, and sediment type should be considered when evaluating any surface geochemical survey data, especially in areas of low volume seepage. If the anomalous population is strongly correlative to sediment

type or grain size, there is a good chance the anomalous population is not related to local seepages. 4.4.7. Comparison of near-surface sediment and reservoir gas composition Few published studies compare surface gas compositions and compound specic isotopic ratios to subsurface trapped gases. A recent paper given by Whiticar at the 2004 AAPG conference concluded that the sorbed surface gases compared well with the subsurface reservoired gases (Whiticar, 2004). Studies by the author in the offshore Gulf of Mexico, using 15 piston cores sampled at varying depths collected along key migration pathways (as determined by seismic) and over a recent discovery, did not look similar. The headspace extracted sediment gases ranged from 0.01 to 32.7% gas wetness for samples with above background concentrations (greater than 10,000 ppm by volume), whereas the reservoired gas wetness ranged from 10.5 to 18.5% (Fig. 9; Abrams, 2004). In another example, methane isotopic ratios from surface sediment cores range from d13C1 K68 to K37 (Fig. 10). The reservoir methane gas ranged from d13C1 K57 to K54. Only 5 of the 15 near surface samples provided results within the reservoir gas boundaries (Abrams, 2004). The samples with methane isotopic compositions lighter are the result of mixing with in situ derived bacterial gases. The samples with methane isotopic compositions heavier are the result of secondary alteration due to bacterial oxidation. There are several possible explanations for the differences between near surface sediment and reservoir gases: Sediment extraction method fractionation (Abrams, 1989, 1996; Abrams et al., 2004); Migration fractionation (chromatographic effect, Krooss and Leythaeuser, 1996); Secondary alteration (bacterial activity); Mixing with in situ derived bacterial gases.

Fig. 8. Variation in adsorbed (acid extraction) sediment gas concentrations relative to percent calcium carbonate (CaCO3) of sediment.

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

471

Fig. 9. Comparison of normalized hydrocarbon gas compositions obtained from headspace extracted seabed cores collected during a surface geochemical survey and MDT exploration tests.

In reality all of the four processes play some role in the measured near-surface sediment extracted gases. Which process is important in your study area will depend on the local petroleum seepage system.

5. Migration pathway analysis Migration pathway analysis is critical in understanding the near-surface seepage in terms of petroleum system dynamics (Macgregor, 1993; Thrasher et al., 1996). Fluid ow modeling, seismic attribute evaluation (mapping vertical noise trails), and surface morphology analysis are independent non-geochemical ways to interpret nearsurface geochemical anomalies and how they relate to subsurface hydrocarbon generation and entrapment. Petroleum seepage along major migration pathways is well documented and targets surface geochemical core locations

(Abrams, 1992, 1996; Reilly et al., 1996; Abrams et al., 2001a,b). Early surface geochemical surveys relied on the vertical leakage concept and collected cores using grid patterns (Matthews, 1996). Few studies have sought to correlate seeps with specic migration pathways identied from reection seismic, seaoor bathymetry, and geochemical measurements. Mapping thermogenic hydrocarbon seeps (oil and gas) relative to potential cross-stratal migration pathways is one way to establish effective migration pathways to charge potential traps. Fluids either ow predominantly along major stratal surfaces or they cross stratal surfaces via faults, dipairs, or major fracture systems. Fluids migrating along stratal surfaces are well documented and relatively well understood (Toth, 1980, 1996). In contrast, cross stratal migration requires sufcient pore pressures to overcome capillary entry pressure in lowered capillary pressures zones. Entry pressure is a function of rock pore throat size and grain

Fig. 10. Comparison of methane carbon isotopic ratios obtained from headspace extracted seabed cores collected during a surface geochemical survey and MDT exploration tests.

472

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

wettability. Factors reecting pore throat size and wettability include rock net-to-gross (sandshale ratio), variable fault slip, stress regime, and pore uid composition. The increased pore pressures can be due to several factors such as rapid deposition, petroleum generation, and local hydrocarbon column heights. These factors may change quickly as evidenced by the episodic nature of petroleum leakage in near-surface sediments (Roberts and Carney, 1997; Quigley et al., 1999; MacDonald et al., 2000; Abrams and Boettcher, 2000). 5.1. Fluid ow modeling Multidimensional uid ow, both along strata and cross stratal, are simulated by modeling programs, e.g. PRA BasinFlow, IES PetroFlow, and IFP Temis. Water and hydrocarbon ows are modeled in two or three dimensions. These modeling programs depend on capillary entry pressure, pore uid dynamics (pore pressure and type), and regional hydrodynamics, which are largely unknown in most exploration areas. Nevertheless, the programs provide generalized understandings of major uid ow directions. Cole et al. (2001) use 2D uid ow modeling, in conjunction with seabed geochemical measurements, to document a transported thermogenic sediment petroleum signal in the deepwater Gulf of Mexico. They concluded that the low concentration thermogenic samples are likely the redistributed oil-bearing sediments from shelf-slope failures. The models strongly suggest that a group of cores with low concentration thermogenic hydrocarbons, identied as low condence samples, are unrelated to the subsurface petroleum charges. Reservoir oils collected after the geochemical surveys conrm the low condence samples are sourced from a different organic facies, and thus unrelated to local charge. Identication of transported hydrocarbons (low condence) by Cole et al. (2001) provides an explanation for discrepancies between predicted (pre-drill) and post-drilling source facies maps published by Wenger et al. (1994). These maps are based on both high and low level seepage, thus mixing the transported (low condence) and in situ hydrocarbon (high condence) signals. 5.2. Seismic attribute analysis Single-trace attributes such as amplitude and frequency can be used to document acoustic anomalies believed to be related to migrating or shallow-generated gas. Gas clouds, gas chimneys, bright spots, pull downs, wipe outs, gas disturbed zones, and blank-out zones are well described in the literature (Sweet, 1973; Anderson and Hampton, 1980; Siddiquie et al., 1981; Edrington and Calloway, 1984; Abrams, 1992). Many geophysicists consider these features as seismic noise that degrades the quality of seismic reections. They devote great efforts to this problem and lter out gas signals.

In 3D seismic cubes and sophisticated attribute analysis, these noise features assist surface geochemists in mapping migration pathways to near-surface (Loseth et al., 2002; Heggland, 1998). A semiautomatic method that highlights vertical noise trails in seismic data uses assemblies of multitrace seismic attributes and neural networking. A chimney cube is a 3D volume of seismic data that highlights vertical chaotic seismic feature (Aminzadeh et al., 2002). 5.3. Surface morphology reects of hydrocarbon migration Leaking hydrocarbons, and accumulated uid ow affects shallow sediments and sea oor character. Seabed morphology depends on several factors: rates and volume of leakage, type of migrating uid (water, oil, and/or gas), sediment environment, time frame (long vs short term), and oceanographic conditions (salinity, temperature, and bottom water currents) (Hovland and Judd, 1988). Morphology features may be positive relief (constructive) or negative (destructive). Constructional features can result from slow accumulation of uidized mud, hydrates (depends on pressure and temperature regime), and/or carbonate hardground (authigenic carbonate from bacterial activity), whereas depressions are produced from the release of geopressured uids or the collapse of uidized sediments. These seabed features range from very small (less than a meter), up to 1 km wide and 50 m high, and therefore are often recognized on seismic and sonar data (Hovland and Judd, 1988; Roberts et al., 1990; Kaluza and Doyle, 1996). The seaoor at uid-expulsion sites generally have an acoustic character signicantly different than that of adjacent areas, displaying localized amplitude anomalies. Dip maps on the seaoor and articially illuminated timestructure maps with amplitude overlays from 3D seismic are effective for locating bathymetric variation, which may be related to leakage. In the absence of near-surface 3D data, high resolution sub-bottom proling (CHIRP), side-scan sonar, 2D seismic proles (sparker, air-gun, etc.), and swath (multi-beam) bathymetry-backscatter maps also may locate uid expulsion features, possibly related to migration pathways. The seabed morphology features described above provide evidence of near-surface uid expulsions where the uids may include gas (biogenic or thermal) and water, as well as oil. Fluid expulsion features result from uid releases due to geopressure (pore pressure in excess of hydrostatic) along a major cross-stratal migration feature (faults, fractures, and diapers). Thus, near-surface uid expulsion features by themselves do not conrm a mature source is present. Sediment samples must be collected and analyzed to conrm that these potential migration pathways are or have been associated with hydrocarbon generation and migration.

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

473

6. Hydrocarbon charge vs surface signal 6.1. Presence mature source rock Localized seepage indicates a generating source is present. Source rock character can be examined if sufcient seep material is available for detailed molecular characterization (GC and GC/MS). Source type (organic matter type), source age (if age diagnostic biomarkers are present), and organic maturity (hydrocarbon generation temperature) may be interpreted, keeping in mind that migrating hydrocarbons in near-surface sediments do not guarantee a nearby structure will be charged with an economic accumulation (Abrams, 2002a). Bolchert et al. (2000) found that many of the major seeps in the northern Gulf of Mexico are in areas with no elds or discoveries, and that many of the elds do not have a near-by surface seep. Thus, the signicance of seepage at or near-by a potential prospect is not straight forward. Thermogenic seepage in near-surface sediments at or near a prospect conrms the existence of a mature source rock. Tying a seep to a specic trap should include migration pathways analysis, using both seismic data and uid ow modeling. Nor does the lack of hydrocarbon seepage condemn a basin or prospect area. Not all petroleum bearing basins have detectable seepage. 6.2. Hydrocarbon charge and charge type Another key goal of surface geochemistry is to predict the likely petroleum phase and composition. Predicting oil, condensate, and gas using near-surface geochemistry has been discussed for years (Horvitz, 1981, 1985). Jones and Drozd (1985) collected gas samples from shallow holes using an inatable packer and pump system. They compared the soil gas composition to reservoir charge type, and developed empirically determined ranges for sediment gas hydrocarbon measurements over different reservoirs. This was a rst and important step in demonstrating that the nearsurface signal may be related to the reservoir composition. The gas composition data was plotted on Pixler plots (Pixler, 1969), using empirically derived ratios to dene probable phase. These empirically derived guidelines may work for the type of sampling used by Jones and Drozd (1985), but should not be directly applied to other nearsurface gas collection methods. A comparison of anomalous sediment gases with reservoir gases led to similar conclusions. Headspace hydrocarbon gas compositions from shallow sediment samples collected at major migration pathways above several Gulf of Mexico offshore elds show systematic changes with reservoir charge (Fig. 11). Samples with gas compositions similar to Line A trend usually contain oil and gas. Samples with gas compositions similar to Line B trend usually are dry. Note the distinction between dry hole (Trend B) and oilgas reservoir (Trend A) is based on very small gas compositions changes, less than 0.1%. Thus, using

Fig. 11. Hydrocarbon charge evaluation using headspace sediment gas composition.

headspace extracted sediment gas data to dene reservoir charge must be used with great caution. The charge phase (gas vs oil) may also be evaluated using seepage magnitude. Gas vs oil may be a function of source and/or retention (differential entrapment) in selected geological settings. The gas vs oil eld distribution in the South Caspian Basin could be a function of differential trap retention more than charge (van Grass et al., 2000; Abrams et al., 2001a,b). Traps with similar oil and gas charges, and very recent tectonic activity, tend to retain oil over gas by differential entrapment resulting in a more oily accumulation. In contrast, other traps with similar oil and gas charges, but no recent tectonic activity, tend to retain both the oil and gas resulting in a more gassy accumulation. The resulting surface signatures differ in the two scenarios. The accumulations with recent tectonic activity are more leaky and have prolic seepage signatures, very high total sediment gas and C12 plus solvent extract hydrocarbon concentrations (Fig. 12). Accumulations with little or no recent tectonic activity are less leaky and have lower total sediment gas and C12 plus solvent extract concentrations

Fig. 12. Gas vs oil evaluation using seepage magnitude in basins with differential entrapment controlling gas-condensate to oil trap distribution.

474

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

(Fig. 12). The presence of a seismic gas chimney above or near the trap may provide physical evidence of differential gas loss (Phipps and Carson, 1982). Factors to remember when predicting reservoir charge using near-surface gas composition and/or seepage magnitude include: Relatively direct and active migration pathway from the reservoir to near-surface sediments; Limitations of method of sediment hydrocarbon gas extraction (headspace, blender, ball mill, or adsorbed/sorbed); Limitations of method of sediment hydrocarbon oil extraction (extraction method, solvent, or sieved vs bulk); Variability (noise to signal ratio, precision, and accuracy) of seep detection method being utilized. If near-vertical leakage is well documented from previous surface geochemical surveys, the pattern of micro-seepage over a eld may reect reservoir heterogeneity, and distinguish hydrocarbon charged compartments not in production (by passed) from drained compartments (hydrocarbon already produced). Microbiological surveys above active and abandoned elds in the Sacramento Basin (California) and Vassar Waterood (Oklahoma) were undertaken to help evaluate drainage patterns, inll locations, step-out potential, and hydrocarbon potential in abandoned elds (Schumacher et al., 1997). Comparison of microbiological surface anomalies appear to match production and drill well activity results according to Schumacher et al. (1997). Again, caution should be used when utilizing surface geochemistry methods for evaluating by-passed oil due to possibilities of: Insufcient direct communication with surface (no vertical migration); Variation among sampling methods and sediment conditions; Possible surface contamination in a producing region. 6.3. Non-hydrocarbon gas Most surface geochemical surveys measure hydrocarbon gas concentrations in areas where petroleum is the main objective. Non-hydrocarbon gases are equally important in selected geologic settings, such as southeast Asia. Large concentrations of non-hydrocarbon gases, such as carbon dioxide and nitrogen, greatly affect the production economics. Nitrogen and carbon dioxide gas predictions usually involve temperature history analyses because basins with large concentrations of reservoired carbon dioxide and nitrogen are often associated with elevated temperature gradients (Giggenbach, 1997). This method should not be used for prospect specic predictions because other factors, such as migration and entrapment may also be important. A more empirical method for non-hydrocarbon gas prediction is near-surface geochemistry. When concentrations

of thermally derived non-hydrocarbon gases are elevated in near-surface sediments, there is a high risk of nonhydrocarbon gases in near-by reservoirs. Headspace sediment gases from a recent site-specic seabed sampling survey above a eld contain 7080% CO2. The resevoired gas is 6070% CO2. Stable carbon isotopic ratios from sediment extracted gases with elevated CO2 gases conrms most of the anomalous seabed carbon dioxide gas is thermal, not bacterial, with values similar to the reservoir gases (Abrams, 1996). 6.4. Oil quality prediction Prediction of oil properties using surface geochemistry uses several methods. Horvitz (1985) noted the uorescence spectra shape varied with oil type. The general shape is dened by the R1 ratio: intensity of the emission band at 360 nm divided by that of the 320 nm band using a 270 nm excitation in both cases. Barwise and Hay (1996) derived an empirical relationship between the uorescence R1 and uid API gravity using 130 oils. Barwise concluded R1 could be used to predict oil gravity. However, thermal maturity, biodegradation, and recent organic matter input will also affect the uid gravity prediction relationship. Molecular characteristics (as determined by high resolution gas chromatography and gas chromatographymass spectrometry) can be directly or indirectly related to API gravity in both oils and surface seeps. Kennicutt and Brooks (1988) chose two molecular parameters to evaluate API gravitypristane/phytane and bisnorhopane/hopanein southern California oils. API gravities predicted using pristane/phytane and bisnorhopane/hopane ratios were in good agreement with those measured in oils from near-by reservoired uids. These molecular ratios should be used cautiously because they depend on the same oil being present throughout an area, and these ratios are affected to varying degrees by a number of oil alteration processes, including biodegradation, water washing, and thermal degradation, as well as recent organic matter interference. It is important to choose molecular parameters which strongly correlate to API gravity and are not easily altered by any of these processes or factors.

7. Summary Near-surface indications of migrating hydrocarbons provide critical information on source (organic matter type), maturation (organic maturity), migration (migration pathway delineation), and in selected geologic settings, prospect specic hydrocarbon charge (phase, composition, and quality). Petroleum seepage system evaluation integrates nearsurface geochemical analyses for basin assessment and prospect evaluation. The PSS has four elements: seepage activity (qualitative evaluation of leakage rates: active vs

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477

475

passive, and episodic vs continuous), seepage type (concentration of migrating thermogenic hydrocarbons relative to in situ materials, macro- vs micro-seepage); migration focus (direction of leakage relative to subsurface hydrocarbon generation and/or entrapment); and near-surface seep disturbances (near-surface processes which alter or block the seepage signals). The rate and volume of hydrocarbon seepage to the surface affects near-surface geological and biological responses, and thus is the optimal type of sampling procedures for detecting hydrocarbon leakage. Best practices for the interpretation of near-surface geochemical measurements should include: Recognition of background vs anomalies in surface geochemical surveys. Identify presence of multiple populations (background vs anomalous) using statistical procedures and graphs, i.e. histogram and cumulative frequency plots. Use of multiple parameters and integrated interpretation. Natural gases are generally analyzed in three ways: gas composition (C1C5 and non-hydrocarbon component such as CO2); compound specic carbon isotopic ratios (d13Cn); and hydrogen isotopic ratios (dDCH4). Gas composition and isotopic ratio depend on type and maturity of sources and on the degree of secondary alteration. Basic screening of high molecular weight (HMW) hydrocarbons includes whole extract gas chromatography (GC) and total scanning uorescence (TSF), followed by gas chromatography/mass spectrometry (GC/MS, biomarkers) for samples with anomalous HMW hydrocarbons. Interpretation of biomarker data from surface sediment seepages differs from oiloil and oilsource correlation studies because of recent organic matter and bacterial alteration which will change the molecular character. It is important to utilize biomarker ratios, which are less susceptible to interferences rather than those commonly used for reservoir analyses. Recognition of pitfalls and problems with surface geochemical data. It is important to recognize and avoid pitfalls and problems caused by transported hydrocarbons seepage (displaced seepage), reworked mature source rocks (source rock contained within sediment provenance), and potential eld or laboratory contamination. Additional problems include variation due to differing analytical methods, changing sediment types, and migration or sampling fractionation-partitioning effects. Migration pathway analyses. Fluid ow modeling, seismic attribute, and surface morphology analyses provide independent non-geochemical ways to interpret nearsurface geochemical anomalies. Petroleum seepages along major migration pathways can potentially be tied to specic traps or sources. All petroliferous basins exhibit near-surface signals, but some geochemical signatures are not distinguishable from background sediment signals with methods currently used by industry. Relationships among near-surface hydrocarbon seepages and subsurface petroleum generation and entrapment are often complex. We need to recognize complexities,

make sense of observations, and reduce exploration risk by better integration of near-surface geochemical measurements. Surface geochemistry is a potentially powerful exploration and production tool which can greatly assist the petroleum systems analyst reduce charge risk if properly used. Acknowledgements The author would like to acknowledge the many coworkers who over the years have provided both technical and moral support in understanding the Petroleum Seepage System: Leo Horvitz (deceased), Al Young, and Deet Schumacher. I also thank the University of Utah Energy and Geoscience Institutes industry sponsors for supporting the Surface Geochemistry Calibration research program: ConocoPhillips (Dan Burgraff, Brad Huizinga, and Albert Holba), Anadarko (Ahmed Chaouche and Harry Dembicki), Total (Denis Levache and Jean Pierre Houzay), Marathon (Mantez McDonald and Eric Barsch), Petrobras (Joao Francolin and Dennis Miller), and Hunt (Bob Webster). The manuscript was reviewed by Al Young (retired EXXON), Jean Whelan (WHOI), and Lucinda Gathercole (WHOI). References
Abrams, M.A., 1989. Interpretation of surface methane carbon isotopes extracted from surcial marine sediments for detection of subsurface hydrocarbons. Association Petroleum Geochemical Exploration Bulletin 5, 139166. Abrams, M.A., 1992. Geophysical and geochemical evidence for subsurface hydrocarbon leakage in the Bering Sea, Alaska. Marine and Petroleum Geology Bulletin 9, 208221. Abrams, M.A., 1996. Distribution of subsurface hydrocarbon seepage in near surface marine sediments. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 114. Abrams, M.A., 2002a. Signicance of hydrocarbon seepage relative to subsurface petroleum generation and entrapment. In: American Association of Petroleum Geologist Convention Abstracts, Annual AAPG Convention, Houston, TX, March 1013, 2002. Abrams, M.A., 2002b. Surface geochemical calibration research study: an example of research partnership between academia and industry. In: New Insights Into Petroleum Geoscience Research Through Collaboration Between Industry and Academia, Geological Society, London, UK, May 89, 2002. Abrams, M.A., 2004. Signicance of gas extracted from marine sediments to evaluate subsurface hydrocarbon generation and charge. In: American Association Petroleum Geology Convention Abstracts, Annual AAPG Convention, Dallas, TX, April 1821, 2004. Abrams, M.A., Boettcher, S.B., 2000. Mapping migration pathway using geophysical data, seabed core geochemistry, and submersible observations in the central Gulf of Mexico. In: American Association of Petroleum Geologist Convention Abstracts, Annual AAPG Convention, New Orleans, Louisiana, April 1619, 2000. Abrams, M.A., Segall, M.P., Burtell, S.G., 2001. Best practices for detecting, identifying, and characterizing near-surface migration of hydrocarbons within marine sediments. In: Offshore Technology Conference, Houston, TX. Proceedings Volume, OTC Paper 13039.

476

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477 Heggland, R., 1998. Gas seepage as an indicator of deeper prospective reservoirs; a study based on exploration 3-D seismic data. Marine and Petroleum Geology 5 (1), 19. Holba, A.G., Huizinga, B.J., 2002. Depositional environment indicators: how to optimize your indicator to avoid the pitfalls of conventional indicators. American Association of Petroleum Geologist Convention Abstracts, Annual AAPG Convention, Houston, TX, March 1013, 2002. Horvitz, L., 1980. Near surface evidence of hydrocarbon movements from depth. In: Roberts III., W.H., Cordell, R.J. (Eds.), Problems of Petroleum Migration. American Association of Petroleum Geologist Memoir No. 10, , pp. 241269. Horvitz, L., 1981. Hydrocarbon prospecting after forty years. In: Gottleib, B.M. (Ed.), Unconventional Methods in Exploration for Petroleum and Natural Gas II. Southern Methodist University Press, Dallas, TX, pp. 8395. Horvitz, L., 1985. Geochemical exploration for petroleum. Science 229, 812827. Hovland, M., Judd, A.G., 1988. Seabed Pochmarks and Seepage. Graham and Trotman, London p. 293. Hunt, J.M., 1996. Petroleum Geochemistry and Geology. W.H. Freeman and Company, New York p. 743. Hunt, J.M., Miller, R.J., Whelan, J.K., 1980. Formation of C4C7 hydrocarbons from bacterial degradation of naturally occurring terpenoids. Nature 288, 577578. James, A.T., 1983. Correlation of natural gases by use of carbon isotopic distribution between gas components. American Association Petroleum Geologist Bulletin 67, 11761191. Jones, V.T., Drozd, R.J., 1985. Predictions of oil and gas potential by near surface geochemistry. American Association Petroleum Geologist Bulletin 67, 932952. Kaluza, M.J., Doyle, E.H., 1996. Detecting uid migration in shallow sediments: continental slope environment, Gulf of Mexico. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near-Surface Expression. American Association of Petroleum Geologist Memoir No. 66, pp. 1526. Kennicutt II., M., Brooks, J.M., 1988. Surface geochemical exploration studies predict API gravity off California. Oil and Gas Journal 12. Klusman, R.W., 1993. Soil Gas and Related Methods for Natural Resource Exploration. Wiley, New York, NY p. 483. Klusman, R.W., Saeed, M.A., 1996. A comparison of light hydrocarbon microseepage mechanisms. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 157168. Krooss, B.M., Leythaeuser, D., 1996. Molecular diffusion of light hydrocarbons in sedimentary rocks and its role in the migration and dissipation of natural gas. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 173184. Larter, S, Head, I., Wilhelms, A., 2000. Implications of slow biodegradation rates in oilelds for crustal biospheres. In: Goldschmidt 2000, September 38, 2000 Oxford UK, Journal of Conference Abstracts, vol. 5, p. 619. Loseth, H, Wensaas, L., Arntsen, B., 2002. Gas chimneysindicating factured cap rock. In: American Association of Petroleum Geologist Near-Surface Hydrocarbon Migration: Mechanisms and Seepage Rates, Vancouver, BC, Canada, September 1619 2001. MacDonald, I.R., Reilly, J.F., Best, S.E., Venkataramaiah, R., Sassen, R., Guinasso, N.L., Amos, J., 1996. Remote sensing inventory of active oil seeps and chemosynthetic communities in the Northern Gulf of Mexico. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 2739. MacDonald, I., Buthman, D.B., Sager, W.W., Peccini, M.B., Guinasso Jr., N.L., 2000. Pulsed oil discharge from a mud volcano. Geology 28 (10), 907910.

Abrams, M.A., Guliev, I.S., Collister, J., 2001. Geochemical evaluation of rapidly subsiding basin: South Caspian Basin. In: American Association Petroleum Geology Convention Abstracts, Annual AAPG Convention, Denver, Colorado, June 36, 2001. Abrams, M.A., Dahdah, N.F., Francu, E., 2004. Evaluating petroleum systems elements and processes in frontier exploration areas using seabed geochemistry. World Oil 225 (6), 5360. Aminzadeh, F., Groot, P.D., Berge, T., Oldenziel, T., Ligtenberg, H., 2002. Determining migration path from seismically derived gas chimney. In: Near-Surface Hydrocarbon Migration: Mechanisms and Seepage Rates, American Association of Petroleum Geologist Hedberg Conference, Vancouver, BC, Canada, April 710 2002. Anderson, A.L., Hampton, L.D., 1980. Acoustics of gas bearing sediments: measurements and models. Journal of the Acoustical Society of America 67, 18901903. Barwise, T., Hay, S., 1996. Predicting oil properties from core uorescence. In: Abrams, M.A., Schumacher, D. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 363371. Bernard, B.B., 1978. Light hydrocarbons in marine sediments. PhD Dissertation, Texas, p. 144. Bernard, B, Brooks, J.M., Zumberge, J., 2002. Determining the origin of gases in near surface sediments. In: Near-Surface Hydrocarbon Migration: Mechanisms and Seepage Rates. American Association of Petroleum Geologist Hedberg Conference, Vancouver, BC, Canada, April 710 2002. Bjoroy, M., Ferriday, I., 2002. Surface geochemistry as an exploration tool: a comparison of results using different analytical techniques. In: American Association of Petroleum Geologist Hedberg Conference Near-Surface Hydrocarbon Migration: Mechanisms and Seepage Rates, Vancouver, BC, Canada, April 710 2002. Bolchert, G., Weimer, P., McBride, B.C., 2000. Structural and straigraphic controls on petroleum seeps, Green Canyon and Ewing Bank, Northern Gulf of Mexico: implications for petroleum migration. Gulf Coast Association of Geological Societies, Transactions I, 6574. Brooks, J.K., Carey, B.D., 1986. Offshore surface geochemical exploration. Oil and Gas Journal 84, 6672. Brooks, J.M., Kennicutt II, M.C., Bernard, L.A., Genoux, G.J., Carey, B.D., 1983. Applications of total scanning uorescence to exploration geochemistry. Offshore Technology Paper, OTC-4624, pp. 393400. Brown, A., 2000. Evaluation of possible gas microseepage mechanisms. Association Petroleum Geochemical Exploration Bulletin 84 (11), 17751789. Claypool, G.E., Kaplan, I.R., 1974. The origin and distribution of methane in marine sediments. In: Kaplan, I.R. (Ed.), Natural Gases in Marine Sediments. Plenum, New York, pp. 99139. Cole, G., Yu, A., Peel, F., Requejo, R., DeVay, J., Brooks, J., Bernard, B., Zumberge, J., Brown, S., 2001a. Constraining source and charge risk in deepwater areas. World Oil 222, 6977. Cole, G., Requejo, R., DeVay, J., Yu, A., Peel, F., Taylor, C.H., Brooks, J., Bernard, B., Zumberge, J., Brown, S., 2001b. The deepwater GOM petroleum system: insights from piston coring, dening seepage, anomalies, and background. In: 21st Annual GCS-SEPM Research Conference, 2001, pp. 315342. Davis, J.B., Squires, R.M., 1954. Detection of microbially produced gaseous hydrocarbons other than methane. Science 119, 381382. Edrington, T.S., Calloway, T.M., 1984. Sound speed and attenuation measurements in gassy sediments in the Gulf of Mexico. Geophysics 49 (3), 297299. Faber, E., Stahl, W., 1983. Analytical procedure and results of an isotopic geochemical surface survey in an area of the British North Sea. In: Brooks, J. (Ed.), Petroleum Geochemistry and Exploration in Europe The Geological Society, pp. 5163. Giggenbach, W.F., 1997. Relative importance of thermodynamic and kinetic processes in governing the chemical and isotopic composition of carbon gases in high heatow sedimentary basins. Geochimica et Cosmochimica Acta 61 (17), 37633785.

M.A. Abrams / Marine and Petroleum Geology 22 (2005) 457477 Macgregor, D.S., 1993. Relationships between seepage, tectonics, and subsurface petroleum reservoirs. Marine and Petroleum Geology 10, 606619. Matthews, D.M., 1996. Importance of sampling design and density in target recognition. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 243254. Oremland, R.S., 1981. Microbial formation of ethane in anoxic estuarine sediments. Applied and Environmental Microbiology 42 (1), 122129. Oremland, R.S., Whiticar, M.J., Strohmaier, F.F., Kliene, R.P., 1988. Bacterial ethane formation from reduced, ethylated sulfur compounds in anoxic sediments. Geochimica et Cosmochimica Acta 51, 18951904. Peters, K.E., Moldowan, J.P., 1993. The Biomarker Guide. Prentice Hall, Englewood Cliffs, NJ p. 63. Piggott, N., Abrams, M.A., 1996. Near surface coring in the Beaufort and Chukchi Seas. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 385400. Pixler, B.O., 1969. Formation evaluation by analysis of hydrocarbon ratios. Journal of Petroleum Technology 21, 665670. Phipps, G.G., Carson, T.G., 1982. The exploration applications of seismic DHI analysis in the Malay Basin. Offshore Technology Paper 4257, pp. 391395. Price, L.C., 1986. A critical overview and proposed working model of surface geochemical exploration. In: Davidson, M.J. (Ed.), Unconventional Methods Exploration for Petroleum and Natural GasIV. Southern Methodist University Press, Dallas, TX, pp. 245304. Quigley, D., Hornaus, J.S., Luyendyk, B.P., Francis, R.D., Clark, J., Washburn, L., 1999. Decrease in natural marine hydrocarbon seepage near Coal Oil Point, California, associated with offshore oil production. Geology 27 (11), 10471050. Reilly II., J.F., MacDonald, I.R., Biegert, E.K., Brooks, J.M., 1996. Geologic controls on the distribution of chemosynthetic communities in the Gulf of Mexico. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 3962. Roberts, H., Carney, R.S., 1997. Evidence of episodic uid, gas, and sediment venting on the northern Gulf of Mexico continental slope. Economic Geology 92, 863879. Roberts, H.H., Aharon, P., Carney, R., Larkin, J., Sassen, R., 1990. Sea oor response to hydrocarbon seeps, Louisiana continental slope. Geomarine Letters 10, 232243. Saunders, D., Burson, K.R., Thompson, C.K., 1999. Model for hydrocarbon microseepage and related near-surface alterations. Association Petroleum Geochemical Exploration Bulletin 83 (1), 170185. Schoell, M., 1983. Genetic characterization of natural gases. American Association of Petroleum Geologist Bulletin 67, 22252238. Schumacher, D., LeSchack, L.A., 2002. Surface Exploration Case Histories: Applications of Geochemistry, Magnetics, and Remote Sensing, American Association of Petroleum Geologist Memoir No. 48 2002 p. 504. Schumacher, D., Hitzman, D.C., Tucker, J., Rountree, B., 1997. Applying high-resolution surface geochemistry to assess reservoir compartmentalization and monitor hydrocarbon drainage. In: Kruizenga, R.J.,

477

Downey, M.W. (Eds.), Applications of Emerging Technologies: Unconventional Methods in Exploration for Petroleum and Natural Gas, pp. 309322. Siddiquie, H.N., Gopala Roa, D., Vora, K.H., Topgi, R.S., 1981. Acoustic masking in sediments due to gases on the western continental shelf of India. Marine Geology 39, 2737. Sweet, W.E., 1973. Marine acoustical hydrocarbon detection. In: Offshore Technology Conference Houston, TX. Proceedings Volume Offshore Technology Conference Paper 1803. Thrasher, J.A., Fleet, A.J., Hay, S.J., Hovland, M., Duppenbecker, S., 1996. Understanding geology is the key to using seepage in exploration: the spectrum of seepage styles. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 223242. Toth, J., 1980. Cross formational gravity ow of groundwater: a mechanism of the transport and accumulation of petroleum. In: Roberts III., W.H., Cordell, R.J. (Eds.), Problems of Petroleum Migration. American Association of Petroleum Geologist Memoir No. 10, pp. 121169. Toth, J., 1996. Thoughts of a hydrologist on vertical migration and near surface geochemical exploration for petroleum. In: Schumacher, D., Abrams, M.A. (Eds.), Hydrocarbon Migration and its Near Surface Effects. American Association of Petroleum Geologist Memoir No. 66, pp. 79284. Ullom, W.L., 1988. Ethylene and propylene in soil gas: occurrences: sources, and impact on interpretation of exploration geochemical data. Association of Petroleum Geochemical Explorationists, Bulletin 4, 6281. van Grass, G., Abrams, M.A., Bilbo, M., Narimanov, A.A., 2000. The use of integrated seepage detection tools in the South Caspian. In: American Association of Petroleum Geologist International Regional Conference Abstracts, Abstracts, Oil and Gas Business of the Greater Caspian AreaPresent and Future Exploration and Production Operations, Istanbul, Turkey, July 912, 2000. Wenger, L.M., Goodoff, L.R., Gross, O.P., Harrison, S.C., Hood, K.C., 1994. Northern Gulf of Mexico: an integrated approach to source, maturation, and migration. In: Geologic Aspects of Petroleum Systems, American Association of Petroleum Geologist/AMGP Hedberg Research Conference, Mexico City, Mexico, October 26, 1994. Whiticar, M.J., 1999. Carbon and hydrogen isotopes systematics of bacterial formation and oxidation of methane. Chemical Geology 161, 291341. Whiticar, M.J., 2002. Characterization and application of sorbed gas by microdesorption CF-IRMS. In: Near-Surface Hydrocarbon Migration: Mechanisms and Seepage Rates, American Association of Petroleum Geologist Hedberg Conference, Vancouver, BC, Canada, April 710 2002. Whiticar, M.J., 2004. Characterization and application of sorbed gas by microdesorption CF-IRMS. In: American Association of Petroleum Geology Convention Abstracts, Annual AAPG Convention, Dallas, TX, April 1821, 2004. Zhang, L., 2003. Vacuum desoprtion of light hydrocarbons adsorbed on soil particles: a new method in geochemical exploration for petroleum. American Association Petroleum Geochemical Exploration Bulletin 87, 8997.

You might also like