--> Origin of Early Overpressure in the Upper Devonian Catskill Delta Complex, Western New York State, by Gary G. Lash and David R. Blood, #30049 (2007).

Datapages, Inc.Print this page

Click to view posters in PDF format.

Poster 1 (1.9 mb)               Poster 2 (0.9 mb)               Poster 3 (1.6 mb)   

Poster 4 (3.8 mb)               Poster 5 (0.4 mb)               Poster 6 (1.2 mb)

Right click on poster to save.

 

PSOrigin of Early Overpressure in the Upper Devonian Catskill Delta Complex, Western New York State*

By

Gary G. Lash1 and David R. Blood2

 

Search and Discovery Article #30049 (2007)

Posted May 23, 2007

 

*Adapted from poster presentation at AAPG Annual Convention, Long Beach, CA, April 1-4, 2007

 

1Dept. of Geosciences, SUNY Fredonia, Fredonia, NY 14063, USA ([email protected])

2Chesapeake Appalachia, Charleston, WV  25302 USA

 

Abstract 

The Upper Devonian Rhinestreet black shale of the western New York state region of the Appalachian Basin has experienced multiple episodes of overpressure generation manifested by at least two sets of natural hydraulic fractures. These overpressure events were thermal in origin and induced by the generation of hydrocarbons during the Alleghanian orogeny close to or at the Rhinestreet’s ~3.1 km maximum burial depth. Analysis of differential gravitational compaction strain of the organic-rich shale around embedded carbonate concretions that formed within a meter or so of the seafloor indicates that the Rhinestreet shale was compacted ~58%. Compaction strain was recalculated to a paleoporosity of 37.8%, a value well in excess of that expected for burial > 3 km. The paleoporosity of the Rhinestreet shale suggests that porosity reduction caused by normal gravitational compaction of the low-permeability carbonaceous sediment was arrested at some depth shy of its maximum burial depth by pore pressure in excess of hydrostatic. The depth at which the Rhinestreet shale became overpressured, the paleo-fluid retention depth, was estimated by use of published normal compaction curves and empirical porosity-depth algorithms to fall between 850 and 1,380 m. Early and relatively shallow overpressuring of the Rhinestreet shale likely originated by disequilibrium compaction induced by a marked increase in sedimentation rate in the latter half of the Famennian stage (Late Devonian) as the Catskill Delta Complex prograded westward across the Appalachian Basin in response to Acadian tectonics. The regional Upper Devonian stratigraphy of western New York state indicates that the onset of overpressure occurred at a depth of ~1,100 m, well in advance of the Rhinestreet shale’s entry into the oil window during the Alleghanian orogeny.

 

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

   

uAbstract

uIntroduction

  uFigures 1-1 – 1-3

uField and petrographic data

  uFigures 2-1 – 5-1

uInterpretation

  uFigures 5-2 – 6-2

uConclusions

uReferences

 

 

 

 

 

 

 

 

 

 

  

Introduction 

Figures 1-1 – 1-3 

Figure 1-1. Relationship of (A) porosity and depth and (B) pressure and depth for a shale that becomes overpressured by disequilibrium compaction at the fluid retention depth (modified after Harrold et al., 2000). (C) porosity-depth profile for Brunei Darussalam (modified after Tingay et al., 2000).

Figure 1-2. Stratigraphic section, showing Rhinestreet Shale, along with overlying Angola Shale and underlying Cashaqua Shale, at Eighteenmile Creek.

Figure 1-3. Geologic map of part of western New York, as location map for study area.

 

Burial-induced mechanical compaction of argillaceous sediment is accomplished by the loss of porosity as sediment particles respond to increasing effective stress by reorienting into more mechanically stable arrangements and pore fluid is expulsed (Hedberg, 1936; Hamilton, 1976; Magara, 1978; Goulty, 2004). This elasto-plastic reduction in porosity of clayey sediment under hydrostatic conditions generally is expressed as some form of the exponential decay function first proposed by Athy (1930),

 

f=foe-cz

 

where z is depth in meters, fo is the initial porosity at z = 0, and c is the compaction coefficient. Indeed, the majority of porosity-depth algorithms created from empirical data (e.g., Sclater and Christie, 1980; Huang and Gradstein, 1990; Hansen, 1996; among others) define by a rapid reduction in porosity at shallow depth, followed by a reduced rate of porosity occlusion in progressively older and more deeply buried sediment.  

Under certain conditions, notably when fluid expulsion during burial is restricted due to low permeability and/or rapid sedimentation, mechanical compaction fails to keep pace with increasing vertical effective stress such that the pore pressure is greater than hydrostatic (Swarbrick et al., 2002). This phenomenon, termed disequilibrium compaction

This paper seeks to demonstrate that Upper Devonian rocks of the Catskill Delta Complex of western New York state were overpressured by disequilibrium compaction relatively early in their burial history. We will use compaction strain measurements from around early formed carbonate concretions in the Upper Devonian Rhinestreet black shale to calculate the final porosity achieved by gravitational mechanical compaction. These results, interpreted in the context of the Upper Devonian-Mississippian stratigraphy of this region of the Appalachian Basin, will be used to estimate the depth at which the Rhinestreet shale became overpressured, its fluid retention depth. The approach documented in this paper may find application in studies of other shale-rich basinal sequences.

 

Field and Petrographic Data 

Figures 2-1 – 5-1 

Figure 2-1. Rhinestreet shale exposed along Eighteenmile Creek.

Figure 2-2. Cashaqua shale-Rhinestreet shale contact along Eighteenmile Creek (note heavily jointed nature of the Rhinstreet shale and carbonate nodule horizons in the Cashaqua shale).

Figure 3-1. Middle concretion horizon in the Rhinestreet shale along Eighteenmile Creek.

Figure 3-2. Close-up of middle concretion horizon in the Rhinestreet shale; note differential compaction of host shale around the concretions (stick = 1.9 m).

Figure 3-3. Small carbonate concretion in the lower concretion horizon showing differential compaction of host black shale around the concretion and laminated internal structure of the concretion.

Figure 3-4. A small, tilted concretion in the lower concretion horizon (scale = 1 m).

Figure 3-5. Differential compaction of host black shale around a concretion; note laminated internal structure of the concretion (scale = 13 cm).

Figure 4-1. Secondary electron image of a Rhinestreet shale collected from a concretion strain shadow showing relatively open or porous microfabric composed of randomly arranged clay flake domains in edge-to-edge and edge-to-face domain contacts.

Figure 4-2. Secondary electron image showing the strongly oriented microfabric in sample recovered 24 cm from a concretion strain shadow.

Figure 4-3. Secondary electron image showing the open framework of an etched Rhinestreet carbonate concretion sample showing a framboid (F) and randomly oriented clay grain domains (c), and microscaprite (ms).

Figure 4-4. Secondary electron image showing a moderately planar clay-grain microfabric in an etched carbonate concretion sample. Many of the clay grains are oriented subparallel to the page. A more open microfabric can be seen in the center of the image.

Figure 5-1. Thicknesses of correlative layers inside a concretion (Td) and in encapsulating shale (Tc).

 

The Upper Devonian succession of western New York State, which includes the Rhinestreet shale, grades upward from a base of marine shales and scattered turbiditesiltstones into shallow marine or brackish-water deposits thus recording progradation of the Catskill Delta across the Acadian foreland basin (Faill, 1985; Ettensohn, 1992). Marine deposits of the Catskill Delta Complex in the northern Appalachian Plateau are arranged in several cycles, each one defined by a basal unit of uniformly laminated fissile black shale that passes upward through a transition zone of alternating black and gray shale beds into strata dominated by poorly bedded (poorly fissile) gray shale and occasional turbidite siltstone and thin black shale beds.  

The Rhinestreet shale, the thickest of the black shale units of the Lake Erie District, western New York state, comprises at least 54 m of heavily jointed, dark-gray to black fissile and thinly laminated pyritic shale, thin gray shale intervals, sparse thin siltstone beds and several intervals of carbonate concretions (Buehler and Tesmer, 1963; Lash and Blood, 2006). The Rhinestreet shale is underlain by the Cashaqua gray shale, the contact being sharp and easily recognized in the field, and passes upward through a zone of interbedded black and gray shale into the Angola shale (Buehler and Tesmer 1963).

The majority of carbonate concretions of the Rhinestreet shale are found in three stratigraphically confined but laterally persistent horizons (see lithologic log). Most concretions are oblate ellipsoids with maximum diameters and thicknesses ranging up to 2.7 m and 1.1 m, respectively. Field observations, including randomly tilted concretions and differential compaction of host sediment laminae around concretions, are consistent with early diagenetic growth in unconsolidated sediment. Further, estimates of pre-cementation host sediment porosity based on the volume percentage of calcium carbonate cement (74 to 93%) and, perhaps most importantly, the preservation of a cardhouse clay fabric observed within concretion samples studied with the scanning electron microscope, suggest that concretionary growth occurred rapidly within perhaps a meter of the seafloor (Lash and Blood, 2004a,b).  

Concretions offer a unique opportunity to quantify the effects of gravitational compaction of the Rhinestreet shale. However, to ensure that our calculations yield finite compaction strain of the host shale, we must be certain that the Rhinestreet concretions formed rapidly and, most importantly, close to the sediment-water interface. Field observations, including the wrapping of shale around concretions and the lack of center-to-edge deviation in laminae thickness, demonstrate that concretions formed rapidly at shallow depth, perhaps a meter or so below the seafloor (Lash and Blood, 2004a, b). Lash and Blood (2004a) maintain that Rhinestreet concretions formed by the passive infilling of host sediment porosity by carbonate cement (e.g., Raiswell and Fisher, 2000). Accordingly, the volume percent of carbonate cement in the concretion matrix is a proxy for the porosity of the host sediment at the time of concretion growth (Raiswell 1971; Gautier, 1982). Volume percent of 21 Rhinestreet concretion samples collected from four concretions varies from 74 to 93% (mean = 83%), a range that encompasses the high end of porosity of modern marine clay deposits close to or at the water-sediment interface (e.g., Müller, 1967; Velde, 1996) further suggesting a very shallow depth of origin. 

Scanning electron microscopic analysis of concretion and host shale samples also provides evidence for shallow concretionary growth (Lash and Blood, 2004a, b). Specifically, SEM images of mudstone samples collected from concretion strain shadows reveal a porous fabric of randomly oriented platy particles, which higher magnification proves to be face-to-face clay flake stacks or domains. Domains typically are arranged in a low-density network or cardhouse fabric of edge-to-edge and edge-to-face contacts marked by large voids relative to the thickness of clay flakes and domains (Lash and Blood, 2004b). Secondary electron images of shale samples collected only 20 to 30 cm from strain shadows, however, show a generally low-porosity microfabric defined by a moderately to strongly preferred orientation of clay flake domains (Lash and Blood, 2004b). The almost negligible degree of compaction observed in strain shadow samples demonstrates that gravitational compaction of the Rhinestreet shale was minimal before carbonate concretions had become incompressible, pointing to a shallow diagenetic origin of the concretions (Lash and Blood, 2004b). Moreover, SEM observations of concretion samples evince a generally open arrangement of detrital clay grains typical of newly deposited flocculated clayey sediment preserved by diagenetic carbonate precipitation (Lash and Blood, 2004a). However, the locally moderate planar clay grain microfabric observed in some concretion samples suggests that the sediment had started to compact, at least locally, as concretions began to form.

Return to top.

 

Interpretation 

Figures 5-2 – 6-2 

Figure 5-2. Porosity versus depth of burial relationships for shale and clayey sediments (after Rieke and Chilingarian, 1974) showing the estimated maximum PFRD depth range based on the mean and the upper and lower standard deviation ranges for the calculated paleoporosity of the Rhinestreet shale. The dashed normal compaction trend, which was used to define the maximum depth for the Rhinestreet shale PFRD, is from Ham (1966); all other trends predict a shallower PFRD. MBD = modeled maximum burial depth of the Rhinestreet shale.

Figure 6-1. Calculated PFRD values for the Rhinestreet shale (fp = 37.8%) using empirical algorithms. Shaded interval defines the PFRD range based on comparison of the Rhinestreet fp with published normal compaction curves for shale. BB = Baldwin and Butler  (1985); FD = Falvey and Deighton (1982); GL = Gallagher and Lambeck (1989); HA = Hansen (1996); HGY = Hegarty et al. (1988); H = Hermanrud et al. (1998); HG = Huang and Gradstein (1990); LR = Liu and Roaldset (1994); SC = Sclater and Christie (1980); T = Tingay et al. (2000); V = Velde (1996); MBD = modeled maximum burial depth of the Rhinestreet shale. The majority of calculated PFRDs fall within the depth range estimated from comparison with published compaction trends.

Figure 6-2. Upper Devonian-Mississippian stratigraphic relations of western New York state (Lake Erie District) and northwest and north-central Pennsylvania. Stratigraphic columns, including thicknesses, are modified from Lindberg (1985). Devonian time scale (including the Devonian-Mississippian boundary) is from Kaufmann (2006); Mississippian time scale is from Gradstein et al. (2004). Note the marked increase in sedimentation rate halfway through the Famennian stage, perhaps glacio-eustatic in origin (e.g., Veevers et al., 1987); that may have been responsible for the onset of disequilibrium in the Upper Devonian Rhinestreet shale.

 

The most obvious measure of gravitational compaction strain sustained by a volume of sediment following accumulation on the seafloor is the change in layer thickness from the concretion into correlative layers of the encapsulating shale. We measured the thickness of bedding or a bedding interval inside concretions (Ti) (Td in Figure 2-12) (and the thickness of that same interval in the shale (To) (Tc in Figure 2-12), a presumed proxy for the original seafloor thickness of the host sediment. Gravitational compaction strain of

the shale outside the strain shadow of the concretion, ε3, is calculated by the following expression,

                                                                                                           

e3=(Ti-To)/Ti

 

The mean e3 of the Rhinestreet black shale based on the analysis of 118 concretions and encapsulating shale throughout the unit, expressed as a negative value, is –0.518 or 51.8% (± 4.9%). This value is noteworthy because normally compacted marine shales typically compact more than 65% upon burial to depths comparable to the maximum burial depth of the Rhinestreet shale.

 

Compaction strain of the Rhinestreet shale can be used as a measure of the porosity achieved at the termination of gravitational mechanical compaction if we assume that all volume loss was caused by vertical shortening, a reasonable assumption based on the lack of layer-parallel shortening caused by Alleghanian tectonics in rocks of the Lake Erie District. Compaction strain is converted to paleoporosity, fp, by the following equation derived by Jacob (1949).

 

fp =(fo +100e3)/(e3 +1)

 

in which fp is expressed as volume percent.

 

However, in order to calculate the fp of the Rhinestreet shale, we first must obtain a reasonable value for the porosity of the sediment at the onset of normal compaction, fo. Textural evidence described in more detail by Lash and Blood (2004a) relates the formation of Rhinestreet concretions to the passive precipitation of diagenetic carbonate in void spaces of the organic-rich host sediment. Thus, the estimated Rhinestreet sediment porosity at the time of concretion growth, based on the volume percent carbonate cement (e.g., Raiswell, 1976; Gautier, 1982), varied from 74 to 93%. However, some authors have suggested that the porosity of newly deposited clayey sediment decreases from as much as 90% to perhaps 60-65% within a decimeter or so of the seafloor (Weller, 1959; Von Engelhardt, 1977; Magara, 1978; Luo et al., 1993). Moreover, Kawamura and Ogawa (2004) demonstrated an especially rapid reduction in void ratio of pelagic clay equal to a 5% drop in porosity down to a depth of 10 cm below the seafloor. Luo et al. (1993) postulated that such marked losses of porosity within several meters to a few tens of meters of the seafloor should be considered a continuation of the depositional process rather than the initial phase of normal load-induced mechanical compaction. Thus, we interpret the range in CaCO3 volume in analyzed Rhinestreet concretions (74 – 93%) to reflect the rapid occlusion of porosity as the water-rich carbonaceous clay passed into and through the zone of anaerobic methane oxidation where concretionary growth occurred (Lash and Blood, 2004a). Indeed, the strong tendancy of organic matter to absorb water thereby favoring a very open depositional microfabric results the rapid collapse of the clay grains into a preferred orientation very early (and at very shallow depth) in the diagenetic history of these types of deposits (e.g, Meade, 1966; Keller, 1982). Thus, we arbitrarily set fo = 70% for our calculation of the Rhinestreet shale fp. Our calculated Rhinestreet shale fp using Jacob’s equation is 37.8% (±7.1%), a value markedly higher than that expected for shale normally compacted to the modeled 3.1 km maximum burial depth of the Rhinestreet shale (Lash and Blood, 2006).

 Return to top.

Conclusions 

Strain analysis of overburden-induced differential mechanical compaction of shale around early (and shallow) formed carbonate concretions in the Rhinestreet shale indicates that the host shale was mechanically compacted ~ 58%, less than that expected for shale buried to 3.1 km, the modeled maximum burial depth of the Rhinestreet shale. Using a reasonable assumption regarding the porosity of the organic-rich sediment at the onset of normal compaction, the calculated compaction strain was translated to a paleoporosity of 37.8%, well in excess of the actual porosity of the Rhinestreet shale as determined by mercury capillary injection pressure measurements. Normal compaction of the Rhinestreet shale was halted well before it entered the oil window as a consequence of the pre-catagenic elevation of pore pressure above hydrostatic. The depth at which the Rhinestreet shale was overpressured, the paleo-fluid retention depth, was estimated by (1) comparison of the paleoporosity with published normal compaction curves and (2) use of several empirically derived porosity-depth algorithms describing the normal compaction of shale. The onset of overpressuring of the Rhinestreet shale appears to have taken place between 850 and 1,380 m below the seafloor, not even half way to its maximum burial depth. The most likely explanation for the early and relatively shallow onset of overpressure in the Rhinestreet shale is disequilibrium compaction. The marked increase in sedimentation rate from the Frasnian and early Famennian (30 m Ma-1) to the latter half of the Famennian (118 m Ma-1) followed by a sharp decrease in sedimentation rate in the Mississppian suggests that disequilibrium compaction was induced toward the end of the Famennian in response to an acceleration of the rate of progradation of the Catskill Delta Complex perhaps induced by a late Famennian glacio-eustatic event. The presence of ~1.1 km of Devonian strata on the base of the Rhinestreet shale suggests that the PFRD must have been at a depth of ~ 1,100 m, well within the estimated range of the PFRD based on published compaction curves and porosity-depth algorithms.

 

References 

Athy, L.F., 1930, Density, porosity and compaction of sedimentary rocks: AAPG Bulletin, v. 14, p. 1-24.

Baldwin, B., and Butler, C.O., 1985, Compaction curves: AAPG Bulletin, v. 69, p. 622-626.

Buehler, E.J. and Tesmer, I.H., 1963,Geology of Erie County, New York: Buffalo Society of Natural Sciences, 21 p.

Burland, J.B., 1990, On the compressibility and shear strength of natural clays: Geotechnique, v. 40, p. 329-378.

Ettensohn, F.R., 1992, Controls on the origin of the Devonian-Mississippian     oil and gas shales, east-central United States: Fuel, v. 71, p. 1487-1492.

Falvey, D.A., and Deighton, I., 1982, Recent advances in burial and thermal geohistory analysis: Journal of the Australian Petroleum Exploration Association, v. 22, p. 65-81.

Faill, R.T., 1985, The Acadian orogeny and the Catskill Delta, in Woodrow, D.L., and Sevon, W.D., editors, The Catskill Delta: Geological Society of America Special Paper 201, p. 15-37.

Gallagher, K., and Lambeck, K., 1989, Subsidence, sedimentation and sea-level changes in the Eromanga Basin, Australia: Basin Research, v. 2, p. 115-131.

Gautier, D.L., 1982, Siderite concretions: indicators of early diagenesis in the Gammon shale (Cretaceous): Journal of Sedimentary Research, v. 52, p. 859-871.

Goulty, N.R., 2004, Mechanical compaction behaviour of natural clays and implications for pore pressure calculation: Petroleum Geoscience, v. 10, p. 73-79.

Gradstein, F.M., OGG, J.G. and Smith, A.G., 2004, A geologic time scale 2004: Cambridge University Press, New York, 610 p.

Ham, H.H., 1966, New charts help estimate formation pressures: Oil and Gas Journal, v. 64, p. 58-63.

Hamilton, E.L., 1976, Variations of density and porosity with depth in deep-sea sediments: Journal of Sedimentary Petrology, v. 46, p. 280-300.

Hansen, S., 1996, A compaction trend for Cretaceous and Tertiary shales on Norwegian Shelf based on sonic transit times: Petroleum Geoscience, v. 2, p. 159-166.

Harrold, T.W.D., Swarbrick, R.E., and Goulty, N.R., 2000, Pore pressure estimation from mudrock porosities in Tertiary basins, southeast Asia, in Swarbrick, R.E., ed., Overpressure 2000 – workshop proceedings: CD volume, paper OP2000_9, 6 p.

Hedberg, H.D., 1936, Gravitational compaction of clays and shales: American Journal of Science, v. 31, p. 241-287.

Hegarty, K.A., Weissel, J.K., and Mutter, J.C., 1988, Subsidence history of Australia’s southern margin: constraints on basin models: AAPG Bulletin, v. 72, p. 615-633.

Hermanrud, C., Wensaas, L., Teige, G.M.G., Vik, E., Nordgard Bolas, H.M., and Hansen, S., 1998, Shale porosities from well logs on Haltenbanken (offshore mid-Norway) show no influence of overpressuring, in Law, B.E., Ulmishek, G.F., and Slavin, V.I., eds., Abnormal pressures in hydrocarbon environments: AAPG Memoir 70, p. 65-85.

Huang, Z., and Gradstein, F., 1990, Depth-porosity relationship from deep sea sediments: Scientific Drilling, v. 1, p. 157-162.

Jacob, C.E., 1949, Flow of Ground Water, in Rouse, H., ed., Engineering hydraulics: John Wiley and Sons, Inc., New York, p. 321-386.

Kaufmann, B., 2006, Calibrating the Devonian time scale: a synthesis of U-Pb ID-TIMS ages and conodont stratigraphy: Earth-Science Reviews, v. 75, p. 175-190.

Kawamura, K., and Ogawa, Y., 2004, Progressive change of pelagic clay microstructure during burial process: examples from piston cores and ODP cores: Marine Geology, v. 207, p. 131-144.

Keller, G.H., 1982, Organic matter and the geotechnical properties of submarine sediments: Geo-Marine Letters, v. 2, p. 191-198.

Lash, G.G., and Blood, D.R., 2004a, Geochemical and textural evidence for early diagenetic growth of stratigraphically confined carbonate concretions, Upper Devonian Rhinestreet black shale, western New York: Chemical Geology, v. 206, p. 407-424.

Lash, G.G., and Blood, D.R., 2004b, Depositional clay fabric preserved in early diagenetic carbonate concretion pressure shadows, Upper Devonian (Frasnian) Rhinestreet shale, western New York: Journal of Sedimentary Research, v. 74, p. 110-116.

Lash, G.G., and Blood, D.R., 2006, The Upper Devonian Rhinestreet black shale of western New York state – evolution of a hydrocarbon system. New York State Geological Association, 78th Annual Meeting Guidebook, p. 223-289.

Lindberg, F.A., 1985, Northern Appalachian Region:  COSUNA Project: AAPG Bookstore, Tulsa Oklahoma.

Liu, G., and Roaldset, E., 1994, A new decompaction model and its application to the northern North Sea: First Break, v. 12, p. 81-89.

Luo, X., Brigaud, F., and Vasseur, G., 1993, Compaction coefficients of argillaceous sediments: their implications, significance and determination, in Dore, A.G., et al., eds., Basin modeling: advances and applications: Norwegian Petroleum Society Memoir 3, p. 321-332.

Magara, K., 1978, Compaction and fluid migration, practical petroleum geology: Elsevier, Amsterdam.

Meade, R.H., 1966, Factors influencing the early stages of compaction of clays and sands – review: Journal of Sedimentary Petrology, v. 36, p. 1085-1101.

Müller, G., 1967, Diagenesis in argillaceous sediments. In Diagenesis in Sediments, in Larson, G., and Chilinger, G.V., eds., Developments in Sedimentology 8, p. 127-177.

Raiswell, R., 1971, The growth of Cambrian and Liassic concretions: Sedimentology, v. 17, p. 147-171.

Raiswell, R., 1976, The microbiological formation of carbonate concretions in the Upper Lias of NE England: Chemical Geology, v. 18, p. 227-244.

Raiswell, R., and Fisher, Q., 2000, Mudrock-hosted carbonate concretions: a review of growth mechanisms and their influence on chemical and isotopic composition: Geological Society of London Journal, v. 157, p. 239-251.

Rieke, H.H., III, and Chilingarian, C.V., 1974, Compaction of Argillaceous Sediments: Elsevier, New York, 424 p.

Sclater, J.G., and Christie, P.A.F., 1980, Continental stretching: an explanation of the post-mid-Cretaceous subsidence of the central North Sea basin: Journal of Geophysical Research, v. 85, p. 3711-3739.

Swarbrick, R.E., Osborne, M.J., and Yardley, G.S., 2002, Comparison of overpressure magnitude resulting from the main generating mechanisms, in Huffman, A.R., and Bowers, G.L., eds., Pressure regimes in sedimentary basins and their prediction: AAPG Memoir 76, p. 1-12.

Tingay, M.R.P., Hillis, R.R., Swarbrick, R.E., Mildren, S.D., Morley, C.K., and Okpere, E.C., 2000, The sonic and density log expression of overpressure in Brunei Darussalam, in Swarbrick, R.E., ed., Overpressure 2000 – workshop proceedings: CD volume, paper OP2000_21, 8 p.

Veevers, J.J. and Powell, C. McA., 1987, Late Paleozoic glacial episodes in Gondwanaland reflected in transgressive-regressive depositional sequences in Euramerica: Geological Society of America Bulletin, v. 98, p. 475-487.

Velde, B., 1996, Compaction trends of clay-rich deep sea sediments: Marine Geology, v. 133, p. 193-201.

Von Engelhardt, W., 1977, The origin of sediments and sedimentary rocks: John Wiley & Sons, New York.

Weller, J.M., 1959, Compaction of sediments: AAPG Bulletin, v. 43, p. 273-310.

Return to top.