SlideShare a Scribd company logo
1 of 8
Download to read offline
Post-oil solid bitumen network in the Woodford Shale, USA — A potential
primary migration pathway
Brian J. Cardott a,
⁎, Charles R. Landis b
, Mark E. Curtis c
a
Oklahoma Geological Survey, 100 E. Boyd St., Rm. N-131, Norman, OK 73019-0628, USA
b
Minerals End Inc., The Woodlands, TX 77381, USA
c
University of Oklahoma, Norman, OK 73019-0628, USA
a b s t r a c ta r t i c l e i n f o
Article history:
Received 18 December 2013
Received in revised form 28 August 2014
Accepted 28 August 2014
Available online 6 September 2014
Keywords:
Gas shale
Shale oil
Post-oil solid bitumen network
Woodford Shale
Porosity development
Source-rock reservoir
An important if not predominant component of porosity in many gas shales has been identified in organic matter.
An organic network in shales has been described in the literature as organic matter (generalized term), kerogen
(primary), or bitumen (secondary). Recognition of the type and origin of an organic network in shales has rele-
vance in establishing the origin and timing of porosity and fracture development. The pervasive nature of the or-
ganic network adopting the shape of pores in Type II kerogen-rich Woodford Shale suggests it is the residue of
primary oil migration. We use the term “post-oil solid bitumen” to distinguish this bitumen occurrence from
“pre-oil solid bitumen” (defined as a precursor of oil). Three forms of this post-oil solid bitumen network are rec-
ognized in reflected white light at 500× magnification and confirmed in backscattered scanning electron micro-
scope images at N2500× magnification, namely speckled (~1–2 μm), wispy (~2–5 μm), and connected (N5 μm).
The post-oil solid bitumen network demonstrates the prior occurrence of oil generation and migration within
this hydrocarbon source rock, provides porosity for hydrocarbon storage sites, and forms hydrocarbon migration
pathways.
© 2014 Elsevier B.V. All rights reserved.
1. Introduction
The basic requirements of a shale resource system for oil and gas
(source-rock reservoir of Hart et al., 2013) are appropriate organic mat-
ter characteristics in the hydrocarbon source rock (e.g., organic matter
type, quantity, and thermal maturity) and brittle rock fabric (Jarvie,
2012). Recognition of organic-matter (e.g., maceral) types and distribu-
tion are essential to the evaluation of shales as oil and gas reservoirs.
Two basic types of organic matter present in hydrocarbon source
rock shales are kerogen and bitumen. Organic geochemists define bitu-
men as the hydrocarbon fraction extracted with organic solvents. By
this definition, pyrobitumen and kerogen are not bitumen because
they are insoluble in the same organic solvents (Durand, 1980; Peters
and Cassa, 1994; Tissot and Welte, 1984). In contrast, organic petrolo-
gists use the term bitumen without reference to solubility, using petro-
graphic features to infer its origin from its passive occurrence in rocks,
mainly as void fillings (i.e., faunal, dissolution, microfracture, coating
and laminar fillings; Landis and Castaño, 1994).
Bitumen and residual solid hydrocarbons are important components
when considering shale as a hydrocarbon source-rock reservoir. Their
existence in shale is visual proof that the rock reached an appropriate
level of thermal maturity to generate liquid hydrocarbons and can be
used as a proxy for other visual, chemical and mineralogical compo-
nents for maturation assessments, especially in rocks whose age pre-
cedes terrestrial flora.
Through the maturation process, the insoluble fraction of the original
bitumen increases, rendering less efficient attempts to remove all of the
originally generated bitumen (Landis and Castaño, 1994). The occurrence
of this insoluble pyrobitumen fraction impacts both geochemical and pet-
rological evaluations of the rock. The presence of these bitumens in a hy-
drocarbon source rock can affect geochemical analyses (e.g., affecting the
Tmax value) and petrographic analyses (e.g., appearing as vitrinite-like
organic matter affecting the vitrinite reflectance analysis and visual kero-
gen analysis) in isolated kerogen-concentrate pellets (Glikson et al.,
1992; Hackley et al., 2013; Snowdon, 1995).
Primary oil migration in a hydrocarbon source rock begins soon after
oil generation when there is sufficient oil saturation (Jarvie, 2012).
Ungerer et al. (1983, p. 135) concluded that oil migration “depends on
the development of a continuous network of oil-wet pores.” As a result
of primary oil migration and expulsion, a solid carbon residue is left
behind in the shale. Taylor et al. (1998, p. 253) described this secondary
organic matter as a “‘kerogen network’ (impsonite)”, using the asphaltic
pyrobitumen (insoluble) vein classification of Abraham (1960). Else-
where, Taylor et al. (1998, p. 84) recognized a network of solid bitumen
(termed impsonite) as a “product of oil that was generated in but not
expelled from the source rocks.” A key question is whether this organic
material is kerogen or bitumen. Belin (1992) recognized laminated and
International Journal of Coal Geology 139 (2015) 106–113
⁎ Corresponding author. Tel.: +1 405 325 8065
E-mail address: bcardott@ou.edu (B.J. Cardott).
http://dx.doi.org/10.1016/j.coal.2014.08.012
0166-5162/© 2014 Elsevier B.V. All rights reserved.
Contents lists available at ScienceDirect
International Journal of Coal Geology
journal homepage: www.elsevier.com/locate/ijcoalgeo
particulate kerogen networks under SEM backscattered electron mode
and concluded that conventional light microscopy is needed to properly
identify organic matter. Recent application of backscattered scanning
electron microscopy of the Barnett Shale by Loucks et al. (2009) recog-
nized the importance of organic matter in shale as an important source
of microporosity, but only described the material as “organic matter”. As
pointed out in the present article, much of this organic matter occurs as
an organic network. Loucks et al. (2012) recognized the importance of
the interconnected network of organic matter pores. Milliken et al.
(2013) recognized the importance of distinguishing kerogen versus bi-
tumen in the occurrence of organic matter porosity. Our purpose in this
paper is to describe the occurrence, origin, and significance of the or-
ganic network found in organic-rich shales. We show that this network
may be observed not only at high magnification in backscattered scan-
ning electron microscopy but also in reflected white light at 500×
magnification.
2. Terminology
Part of the problem of recognizing bitumen is that it can go by many
terms: bitumen, pyrobitumen, asphalt, asphaltite, asphaltic pyrobitumen,
solid bitumen, solid hydrocarbon, migrabitumen, reservoir bitumen, dead
oil, and exudatinite (Abraham, 1960; Curiale, 1986; Hunt, 1979; Jacob,
1989, 1993; Landis and Castaño, 1994; Taylor et al., 1998). The term
solid bitumen, following Curiale (1986), will be used in this article.
A common generic classification of solid bitumen (primarily
fracture-filling vein deposits of altered, once-liquid oil: asphaltite ver-
sus asphaltic pyrobitumen) was given by Abraham (1960). Jacob
(1989) modified this classification to include petrographic parameters
(e.g., bitumen reflectance, fluorescence, solubility in immersion oil)
and introduced the term migrabitumen (primarily used for vein
deposits) for amorphous, secondary macerals dispersed in rocks and
taking on the shape of voids. Curiale (1986, p. 559) developed a simple
genetic classification of solid bitumen: pre-oil solid bitumen (defined as
“early-generation (immature) products of rich source rocks”) and post-
oil solid bitumen (defined as “products of the alteration of a once-liquid
oil”). It is widely recognized that (pre-oil solid) bitumen forms from
kerogen (Barker, 1979, p. 41; Bernard et al., 2012a; Tissot and Welte,
1984, p.176). Based on hydrous-pyrolysis experiments, Lewan (1983)
demonstrated that pre-oil solid bitumen is the intermediate product be-
tween kerogen and oil.
The terms solid bitumen and migrabitumen were first used for non-
disseminated organic matter occurring as vein deposits known as
asphaltite (soluble) and asphaltic pyrobitumen (insoluble). These
fracture-fillings occur as a once-liquid oil altered to a solid from near-
surface, low-temperature alteration of crude oil by limited biodegrada-
tion, water-washing, and devolatilization (Curiale, 1983). As discussed
below, post-oil solid bitumen can also be any alteration of a once-liquid
oil into a solid, including the dispersed solid residue of oil migration.
Therefore, the genetic solid bitumen classification of Curiale (1986) will
be used here to distinguish bitumen as a precursor of oil (pre-oil solid bi-
tumen) from bitumen formed as an alteration of a once-liquid oil (post-
oil solid bitumen). Thompson-Rizer (1987) described this post-oil solid
bitumen network as an amorphous kerogen in strewn slides.
Pre-oil solid bitumen and post-oil solid bitumen can appear similar to
vitrinite in reflected white light at 500× magnification. Hackley et al.
(2013) concluded that vitrinite reflectance measurements of early mature
Devonian shale may erroneously include solid bitumen lower reflectance
values. Landis and Castaño (1994) identified three types of solid bitumen
(homogenous, granular, and coked). In addition to distinguishing
vitrinite-like bitumen to exclude from the vitrinite-reflectance analysis,
homogenous solid bitumen reflectance values may be used to calculate
a vitrinite reflectance equivalent (VRE; Landis and Castaño, 1994). The
VRE value may be used as a thermal maturity indicator when vitrinite is
not present or to verify the vitrinite-reflectance value. Distinguishing
solid bitumen from vitrinite is more easily accomplished in whole-rock
particulate pellets than in kerogen-concentrate pellets.
3. Methods
Type II kerogen-rich (oil generative organic matter) marine Woodford
Shale (Late Devonian–Early Mississippian) samples from Oklahoma, USA
covering a wide range of thermal maturities (0.50–6.36% vitrinite reflec-
tance, VRo; random reflectance measured in non-polarized light follow-
ing ASTM, 2011) have been examined in reflected white light (200×
and 500×; whole-rock particulate pellets; oil immersion) using a Vickers
M17 Research Microscope system equipped with a Smith illuminator
(Cardott, 2012). The samples (Table 1) are from the Organic Petrography
Laboratory (OPL) of the Oklahoma Geological Survey. A coal stringer sam-
ple (OPL 1387) from a Woodford Shale core was used to determine the
thermal maturity whereas the shale sample in the same core was used
to visualize the post-oil solid bitumen network. An outcrop sample (OPL
1300) did not show signs of weathering (Lo and Cardott, 1994). The pre-
dominant form of silica in the quartz-rich Woodford Shale is biogenic sil-
ica formed from radiolarians (Cardott and Chaplin, 1993). Core, outcrop,
and well cuttings samples that were recognized in reflected white light
to contain post-oil solid bitumen forms or amorphous organic matter
groundmass were selected for examination by scanning electron micros-
copy (SEM). Core, outcrop, and larger chips from cuttings were initially
prepared by mechanical polishing. The samples were then ion milled in
a Fischione Model 1060 ion mill using Argon gas. Milling was performed
with the sample rotating under two crossing ion beams at 5 kV accelerat-
ing voltage for 3 h. The elevation of the ion beams was 2° above the sam-
ple horizontal. Ion beam milling provides a very flat surface for imaging
and preserved the microstructure of the samples with minimal artifacts.
Previous work has shown that even with higher energy focused ion
Table 1
Woodford Shale samples used in this study arranged by increasing vitrinite reflectance.
OPL numbera
Sample type County Geologic province Depth (m) Latitudeb
Longitudeb
VRo (%)c
n VRo range (%)
1300 Grab Murray Arbuckle Mountains Surface 34.444411 −97.130651 0.53 31 0.45–0.67
1333 Core Pottawatomie Cherokee Platform 1395 34.99476 −97.05175 0.59 50 0.50–0.70
601 Core Marshall Ardmore Basin 926 34.05082 −96.64148 0.62 24 0.48–0.77
1371 Cuttings Coal Arkoma Basin 1939 34.64914 −96.39634 0.76 30 0.66–0.88
1366 Cuttings Coal Arkoma Basin 1900 34.650283 −96.356129 0.85 35 0.76–1.01
1372 Cuttings Coal Arkoma Basin 2125 34.67924 −96.36197 0.89 27 0.74–1.07
1398 Core Washington Cherokee Platform 512 36.43725 −95.92348 0.90 29 0.82–1.02
1397 Cuttings Johnston Ardmore Basin 2525 34.173551 −96.792293 0.98 26 0.85–1.10
1076 Core Okfuskee Cherokee Platform 1126 35.33143 −96.08535 1.23 48 1.10–1.47
1402 Cuttings Carter Ardmore Basin 3812 34.287887 −97.072195 1.31 24 1.19–1.45
1387 Core of coal stringer Canadian Anadarko Basin 3809 35.66829 −98.301331 1.62 50 1.51–1.74
1373 Cuttings Coal Arkoma Basin 2556 34.562946 −96.230753 1.67 23 1.38–1.99
a
Oklahoma Geological Survey Organic Petrography Laboratory sample number.
b
NAD 83.
c
Random vitrinite reflectance in non-polarized light with fixed stage.
107B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
beam milling, no significant alteration of the organic matter microstruc-
ture is observed as a result of the ion milling (Curtis et al., 2010). In addi-
tion, pores have been observed in organic matter that has not undergone
ion milling suggesting that pores exist naturally in the organic matter of
some samples. Samples were imaged in a FEI Helios Dual Beam FIB/SEM
using backscattered electrons (BSE) for atomic number contrast. The
accelerating voltage was 1 kV and the beam current was 0.40 nA.
4. Discussion
4.1. Occurrence of post-oil solid bitumen network
Pre-oil solid bitumens are common in immature to mature (oil win-
dow) hydrocarbon source rocks. They are recognized in reflected white
light in whole-rock pellets by their texture (e.g., homogenous, granular,
and coked), semi-translucent character with internal reflections from
imbedded pyrite, and by the pyrite that occurs on their edges (Fig. 1;
ASTM, 2011).
In addition to amorphous blobs of pre-oil solid bitumen and
fracture-filling post-oil solid bitumen, small pieces of organic matter
filling voids are recognized under reflected white light at 500× magni-
fication. Others have recognized this network, although not using the
terminology we use here. Mahlstedt and Horsfield (2012) referred to
post-oil solid bitumen as a carbon-rich pyrobitumen (“pore-occluding
petroleum”) that can undergo secondary cracking to gas and conden-
sate at N1.1% VRo. Landis and Castaño (1994) described residual solid
hydrocarbons occurring as intergranular pore fillings (b10 μm). The
void-filling occurrence of this material suggests that it is the solid resi-
due of primary oil migration (e.g., post-oil solid bitumen). The term
post-oil solid bitumen, as we use it here, has no inference for solubility.
Three post-oil solid bitumen forms are recognized in reflected white
light at 500× in whole rock particulate pellets: speckled (~1–2 μm;
Fig. 2a); wispy (~2–5 μm; Fig. 2b); and connected (N5 μm; Fig. 2c).
The sizes are relative and should not be considered specific.
Speckled and wispy are mostly isolated occurrences in the rock rath-
er than as a connected network. The speckled post-oil solid bitumen
network, where it occurs, is near the magnification limit of the light mi-
croscope (~1–2 μm) and is used mostly as a presence or absence indica-
tor of oil generation and migration. Speckled post-oil solid bitumen
must be carefully distinguished from clay minerals. Speckled and
wispy types are not exclusive — that is, they do not occur in shale as ei-
ther one or the other. Grains that contain the wispy post-oil solid bitu-
men network also will contain areas of speckled post-oil solid
bitumen network (Fig. 3). Speckled and wispy forms are easy to see
but difficult to photograph in reflected white light (500×) because of
focus issues and interference from pyrite. The post-oil solid bitumen
shapes and sizes that form a network are confirmed and best viewed
in BSE images. BSE imaging is sensitive to the atomic number of the
sample material. This results in the low atomic number organic matter
(mostly carbon) appearing dark gray whereas higher atomic number
minerals such as quartz, carbonates, and pyrite exhibit progressively
higher grayscale values (Figs. 4–5). Belin (1994) noted that even
though organic matter types cannot be identified in BSE images, rela-
tionships of organic matter and minerals are revealed with fine resolu-
tion. Although they did not use the solid bitumen terminology described
here, Bernard et al. (2012b, p. 7) recognized the speckled and wispy
post-oil solid bitumen network by using scanning transmission X-ray
microscopy and transmission electron microscopy of the Barnett
Shale. Similar to what we see in the Woodford Shale, their work indicat-
ed that “Organic matter appears as micron sized angular organic grains
irregularly distributed within the mineral matrix or as organic masses
filling intergranular porosity and exhibiting smoothly curved concave
surfaces.”
4.2. Origin of post-oil solid bitumen network
Primary oil migration occurs within the hydrocarbon source rock
(Cordell, 1972). McAuliffe (1979) proposed that primary oil migration
occurs in a 3D kerogen network. However, the occurrence of secondary,
amorphous dispersed organic matter filling voids suggests that this
organic matter was once a liquid, and thus not kerogen. The pervasive
nature and relationship with fracture filling bitumen indicate that it is
post-oil solid bitumen. The mechanisms of primary oil migration
through a kerogen network proposed by McAuliffe (1979) hold true
for a post-oil solid bitumen network. Much of the generated oil does
not migrate out of the rock. Meyer (2012, p. 72) indicated that “for
every barrel of crude oil in conventional reservoirs … there are 8 bbl
of potentially producible oil equivalents remaining in the source rock”
and “Speculative estimates of just how much generated oil remains in
shale source rocks range between 45% and 95% depending on the geol-
ogy of the formation and the quality of the estimate.” Some of the shale-
hosted oil will result in a carbon residue (possibly the same as residual
oil of Fan et al., 2012). Hunt (1996, p. 597–598, see references within on
p. 598) recognized a “refractory bitumen” or “pyrobitumen residue”
retained in the source rock.
4.3. Lowest thermal maturity with post-oil solid bitumen network
The most common organic-matter type in low thermal maturity
Type II kerogen-rich shales and boghead coals is amorphous organic
matter (AOM) (Mastalerz et al., 2012; Thompson-Rizer, 1993). This pri-
mary maceral, AOM, is derived from degraded, unidentifiable precursor
organisms (Pacton et al., 2011). AOM is equivalent to the term
bituminite (ASTM, 2011). Lewan (1987) reported that amorphous
Fig. 1. (A) Homogenous texture in semi-translucent pre-oil solid bitumen (dark gray material in center of photomicrograph) showing internal reflections from pyrite (reflected white light,
500×; whole-rock particulate pellet; Woodford Shale; OPL 1333; 0.59% VRo). (B) Granular texture in pre-oil solid bitumen (reflected white light, 500×; whole-rock particulate pellet;
Woodford Shale; OPL 1076; 1.23% VRo). Solid bitumen classification is modified from Curiale (1986) and Landis and Castaño (1994).
108 B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
Type II kerogen comprised N80 vol% of isolated kerogen from Woodford
Shale samples in Oklahoma. There are several classifications of AOM.
Thompson and Dembicki (1986) recognized four types of AOM related
to hydrocarbon-generating potential (Types A–D). Taylor et al. (1998,
p. 250) also recognized four types of unstructured organic matter
(bituminite). Senftle et al. (1993) indicated that fluorescing AOM
(fluoramorphinite) can be distinguished from nonfluorescing AOM
(hebamorphinite) up to 1.1% VRo in an estimate of oil and gas potential.
Only the fluorescing type of AOM (fluoramorphinite and types A and
D) is considered a source of pre-oil solid bitumen and oil. AOM is best
recognized using strewn slides in transmitted white light and reflected
fluorescent light at 500× magnification. The distribution of AOM in
shale forms an organic network (Figs. 6, 7). The AOM network could
be misidentified as the post-oil solid bitumen network.
Post-oil solid bitumen ultimately forms from oil generated in the oil
window. The lowest thermal maturity containing a post-oil solid
bitumen network is uncertain because it may be confused with AOM
(a primary maceral). The post-oil solid bitumen network could have de-
veloped preferentially along the AOM network. Lewan (1987) described
primary oil migration occurring along a continuous bitumen network
formed from kerogen and impregnating AOM. He described the devel-
opment of an opaque pyrobitumen groundmass (e.g., post-oil solid bitu-
men network) carbonized from bitumen and retained oil.
The lowest thermal maturity where the post-oil solid bitumen net-
work is observed in the Woodford Shale is 0.76% VRo near the middle
of the oil window (Fig. 8). Below this thermal maturity, the organic net-
work could be AOM.
4.4. Significance of post-oil solid bitumen network
The presence of post-oil solid bitumen demonstrates that oil was
generated in or migrated through the rock even though the rock is cur-
rently at a higher thermal maturity than the oil window (Thompson-
Rizer, 1987). Recent applications of BSE images have not only revealed
the occurrence of a network of organic matter, but also the development
Fig. 2. (A) Speckled (~1-2 μm) post-oil solid bitumen network (white arrow; reflected
white light, 500×; Woodford Shale; OPL 1402; 1.31% VRo). (B) Wispy (~2–5 μm) post-
oil solid bitumen network (white arrow; reflected white light, 500×; Woodford Shale;
OPL 1372; 0.89% VRo). (C) Connected (N 5 μm) post-oil solid bitumen network (white
arrow; reflected white light, 500×; Woodford Shale; OPL 1366; 0.85% VRo). P = pyrite.
Fig. 3. Speckled (~1–2 μm) and wispy (~2–5 μm) post-oil solid bitumen network in the
same grain (reflected white light, 500×; Woodford Shale; OPL 1387; 1.62% VRo).
Fig. 4. Speckled (~1–2 μm; white arrow) and wispy (~2–5 μm; black arrow) post-oil solid
bitumen network in backscattered SEM (5000×; Woodford Shale; OPL 1397e; 0.98% VRo).
109B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
of secondary nanoporosity (i.e., pores several nanometers in size;
Loucks et al., 2009; Ruppert et al., 2013; Fig. 9b).
Nanoporosity in organics has been described in the literature as de-
veloping at N0.6% to ~0.9% VRo primarily in post-oil solid bitumen.
Curtis et al. (2012a) reported that the development of secondary
nanoporosity is related to both thermal maturity (beginning about
0.9% VRo; Fig. 10) and organic-matter type (e.g., post-oil solid bitumen).
In contrast, Reed et al. (2012) reported nanopore development in or-
ganic matter beginning at about 0.8% VRo. Loucks et al. (2012) and
Zhang et al. (2012) reported nanopore development in organic matter
N0.60% VRo. Romero-Sarmiento et al. (2013) attributed nanoporosity
development to the maturation of kerogen in the Barnett Shale begin-
ning at about 0.7% VRo. Bernard et al. (2012a) reported nanoporous
pyrobitumen (e.g., post-oil solid bitumen) in a Posidonia Shale sample
at 1.45% VRo, but no organic nanoporosity in samples at 0.5% and
0.85% VRo. Milliken et al. (2012) demonstrated that secondary porosity
develops primarily in intergranular organic matter (e.g., post-oil solid
bitumen) instead of within particulate organic matter (e.g., kerogen)
and that porosity increases with increasing total organic carbon content.
Hao et al. (2013, p. 1342) concluded that “gas sorption in organic-rich
shales is mainly associated with micropores” (b2 nm).
In addition to the biogenic-silica-rich Woodford Shale, a post-oil
solid bitumen network has also been observed in other shales, including
the Barnett, Haynesville, and Horn River shales (Curtis et al., 2012a).
Although not using the post-oil solid bitumen network terminology
used here, others have observed the network in other formations.
Bernard et al. (2012a) recognized aliphatic-rich bitumens (e.g., pre-oil
solid bitumen) and aromatic-rich pyrobitumens (e.g., post-oil solid
bitumen) in an overmature (1.45% VRo) sample of the Posidonia
Shale. Bernard et al. (2012b) recognized pre-oil solid bitumen (derived
from thermally degraded kerogen) and post-oil solid bitumen
(nanoporous pyrobitumen resulting from the secondary thermal crack-
ing of retained oil) in the Barnett Shale. Milliken et al. (2012) recognized
secondary porosity in “organic particulate debris and solid bitumen”
using field-emission scanning electron microscope images of Ar ion-
milled surfaces of the Barnett Shale (Mississippian). The predominant
pore-filling organic matter, interpreted as solid bitumen, was recog-
nized as originating as a liquid hydrocarbon. Hackley (2012) recognized
an interconnected (post-oil) solid bitumen network in whole-rock par-
ticulate pellets of argillaceous lime wackestones and mudstones of the
Lower Cretaceous Pearsall Formation. Uffmann et al. (2012) described
a (post-oil) solid bitumen network in whole-rock pellets of high ther-
mal maturity Mississippian and Pennsylvanian black shales from
Germany and Belgium. Fishman et al. (2012) did not recognize bitumen
or pyrobitumen in the Kimmeridge Clay Formation and concluded that
petroleum storage potential was attributed to inorganic pores. In con-
trast, Fishman et al. (2013) recognized nanoporosity in high maturity
(~1.2% VRo) migrated bitumen from Eagle Ford Shale core samples
equivalent to the post-oil solid bitumen terminology used here.
Kosakowski and Krajewski (2014, their Fig. 11E) recognized a post-oil
solid bitumen network in carbonates in Poland.
Organic pores are not only sites of methane storage by adsorption to
the pore walls (Hackley, 2012; Zhang et al., 2012), but also provide mi-
gration pathways for production of natural gas. 3-D reconstructions of
Focused Ion Beam/SEM tomography samples illustrate the distribution
and connection of nanopores in the post-oil solid bitumen network
(Curtis et al., 2012b, their Fig. 8). Microfractures that connect to the
post-oil solid bitumen network can be seen in BSE images (Fig. 9a) dem-
onstrating the preferred fracture pattern following zones of weakness
through the bitumen. The microfractures (formed either naturally or in-
duced by hydrofracturing) contribute to the rock permeability. Zagorski
et al. (2013, p. 172) recognized that “The observed intraorganic porosity
displays a high degree of connectivity and is responsible for a significant
portion of the Marcellus Shale's productivity and gas in place.”
Applied to gas shales, Blood (2011, p. 56) recognized that “Organic
particles are the sites of adsorbed gas, and amorphous organic matter
and bitumen represent the dominant sites of porosity development
within the Marcellus.” As recognized by Belin (1992) for a kerogen
Fig. 5. Connected (N 5 μm) post-oil solid bitumen network (white arrow) in backscattered
SEM (2540×; Woodford Shale; OPL 1402; 1.31% VRo).
Fig. 6. Amorphous organic matter (AOM; also referred to as bituminite) groundmass (dark
gray material) in Woodford Shale marine boghead coal in backscattered SEM (800×; OPL
1300e; 0.53% VRo).
Fig. 7. Amorphous organic matter (AOM) matrix in backscattered SEM (7500×; Woodford
Shale; OPL 601; 0.62% VRo).
110 B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
network, the post-oil solid bitumen network may be discontinuous
(e.g., speckled and wispy) or continuous (e.g., connected) based on
total organic carbon content and available porosity. Lewan (1987) ob-
served that the (post-oil solid) bitumen network occurs in amorphous
Type II kerogen-rich shales. We are in agreement with Lewan (1987,
p. 128) that “Impregnation of the groundmass with [pre-oil solid] bitu-
men to form a continuous network appears to be a prerequisite for the
expulsion of generated oil.” The pervasive post-oil solid bitumen resi-
due left behind during primary oil migration provides nanoporosity
sites for hydrocarbon storage and microfracture permeability and path-
ways for hydrocarbon production.
5. Summary and conclusions
Primary oil migration in shales leaves behind a solid carbon residue in
available porosity that we describe as a post-oil solid bitumen network.
Development of the network is dependent on kerogen-type and total-
organic-carbon content. This study reports the development of a post-
oil solid bitumen in one of the most recognized conventional hydrocar-
bon source rocks in North America. The Woodford Shale lithofacies are
broadly characterized as Type II kerogen assemblages. The initial devel-
opment of unconventional reservoirs focused on Paleozoic rocks but fur-
ther work on a wider range of kerogen types is needed to assess the
proposed concept of post-oil solid bitumen networks more broadly.
The network is recognized in our Type II kerogen-rich Woodford Shale
samples in reflected white light at 500× magnification as speckled
(~1–2 μm), wispy (~2–5 μm), and connected (N5 μm) forms. (The sizes
are relative and should not be considered specific.) Speckled and wispy
forms are mostly isolated occurrences in the rock rather than as a con-
nected network. The speckled post-oil solid bitumen network is near
the magnification limit of the light microscope (~1–2 μm) and is used
mostly as a presence or absence indicator of oil generation and primary
oil migration. The small size should not be confused with clay minerals.
Speckled and wispy networks often both occur in the same rock. The
post-oil solid bitumen shapes and sizes that form a network are con-
firmed and best viewed in backscattered scanning electron microscope
images.
The post-oil solid bitumen networks in the Woodford Shale demon-
strate: (1) that the rock has generated oil; (2) that oil has migrated
through the rock; (3) that secondary nanoporosity, developing
Fig. 8. Lowest thermal maturity (0.76% VRo) Woodford Shale sample that contains post-oil solid bitumen network (connected; N 5 μm) in (A) reflected white light (500×) with pre-oil solid
bitumen homogenous form, and (B) backscattered SEM (dark gray; 15,000×; OPL 1371). The organic matter does not contain nanopores.
Fig. 9. (A) Microfracture development (possibly caused by sample handling) along post-oil solid bitumen network from Woodford Shale cuttings sample of highest thermal maturity
(1.67% VRo) condensate well in backscattered SEM (5000×; OPL 1373) demonstrates preferred zones of weakness within bitumen for fracture formation; (B) Nanoporosity in wispy
(~2–5 μm) post-oil solid bitumen network (dark gray) in backscattered SEM (20,000×; OPL 1373).
111B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
beginning at ~0.9% VRo, provides storage sites for hydrocarbons;
(4) that these sites are zones of weakness for the formation of
microfractures; and (5) that they also form migration pathways for
hydrocarbons.
Acknowledgments
The authors gratefully acknowledge reviews by Joseph A. Curiale and
an anonymous reviewer that improved the manuscript.
References
Abraham, H., 1960. sixth edition. Asphalts and Allied Substances 5 vol. Van Nostrand
Company, Inc., New York.
ASTM, 2011. Standard test method for microscopical determination of the reflectance of
vitrinite dispersed in sedimentary rocks: West Conshohocken, PA, American Society
for Testing and Materials International. Annual book of ASTM standards: Petroleum
products, lubricants, and fossil fuels; gaseous fuels; coal and coke, sec. 5, v. 5.06,
D7708-11, pp. 823–830 http://dx.doi.org/10.1520/D7708-11.
Barker, C., 1979. Organic geochemistry in petroleum exploration. American Association of
Petroleum Geologists Education Course Note 10, (159 pp.).
Belin, S., 1992. Application of backscattered electron imaging to the study of source rocks
microtextures. Org. Geochem. 18, 333–346.
Belin, S., 1994. Backscattered electron imaging applied to source rock sedimentology: a
comparision with conventional methods in organic petrology. Bulletin Des Centres
De Recherches Exploration-Production, v. 18, Special Publication, pp. 165–187.
Bernard, S., Horsfield, B., Schulz, H.-M., Wirth, R., Schreiber, A., Sherwood, N., 2012a.
Geochemical evolution of organic-rich shales with increasing maturity: a STXM and
TEM study of the Posidonia Shale (Lower Toarcian, northern Germany). Mar. Pet.
Geol. 31, 70–89.
Bernard, S., Wirth, R., Schreiber, A., Schulz, H.-M., Horsfield, B., 2012b. Formation of
nanoporous pyrobitumen residues during maturation of the Barnett Shale (Fort
Worth Basin). Int. J. Coal Geol. 103, 3–11.
Blood, D.R., 2011. Sequence stratigraphy crucial to lateral placement in Marcellus Shale
play, part two. Am. Oil Gas Report. 54 (8), 52–60.
Cardott, B.J., 2012. Thermal maturity of Woodford Shale gas and oil plays, Oklahoma, USA.
Int. J. Coal Geol. 103, 109–119.
Cardott, B.J., Chaplin, J.R., 1993. Guidebook for selected stops in the western Arbuckle
Mountains, southern Oklahoma. Okla. Geol. Surv. Spec. Publ. 93-3 (55 pp.).
Cordell, R.J., 1972. Depths of oil origin and primary migration: a review and critique. AAPG
Bull. 56, 2029–2067.
Curiale, J.A., 1983. Petroleum occurrences and source-rock potential of the Ouachita
Mountains, southeastern Oklahoma. Bull. Okla. Geol. Surv. 135, 65.
Curiale, J.A., 1986. Origin of solid bitumens, with emphasis on biological marker results.
Org. Geochem. 10, 559–580.
Curtis, M.E., Ambrose, R.J., Sondergeld, C.H., Rai, C.S., 2010. Structural characterization of
gas shales on the micro- and nano-scales. Canadian Unconventional Resources and
International Petroleum Conference, Calgary, Alberta, Canada, SPE 137693.
Curtis, M.E., Cardott, B.J., Sondergeld, C.H., Rai, C.S., 2012a. Development of organic poros-
ity in the Woodford Shale with increasing thermal maturity. Int. J. Coal Geol. 103,
26–31.
Curtis, M.E., Sondergeld, C.H., Ambrose, R.J., Rai, C.S., 2012b. Microstructural investigation
of gas shales in two and three dimensions using nanometer-scale resolution imaging.
AAPG Bull. 96, 665–677.
Durand, B., 1980. Sedimentary organic matter and kerogen. Definition and quantitative
importance of kerogen. In: Durand, B. (Ed.), Kerogen. Insoluble Organic Matter
from Sedimentary Rocks. Editions Technip, Paris, pp. 13–34.
Fan, B., Pang, X., Zhang, X., Zhang, J., 2012. Distinction between oil expulsion history and
gas expulsion history. Nat. Resour. Res. 21, 233–243.
Fishman, N.S., Hackley, P.C., Lowers, H.A., Hill, R.J., Evenhoff, S.O., Eberl, D.D., Blum, A.E.,
2012. The nature of porosity in organic-rich mudstones of the Upper Jurassic
Kimmeridge Clay Formation, North Sea, offshore United Kingdom. Int. J. Coal Geol.
103, 32–50.
Fishman, N.S., Guthrie, J., Honarpour, M., 2013. The stratigraphic distribution of hydrocar-
bon storage and its effect on producible hydrocarbons in the Eagle Ford Formation,
south Texas. Unconventional Resources Technology Conference, SPE-AAPG-SEG, Den-
ver, CO, Paper 1579007 (6 pp.).
Glikson, M., Taylor, D., Morris, D., 1992. Lower Paleozoic and Precambrian petroleum
source rocks and the coalification path of alginite. Org. Geochem. 18, 881–897.
Hackley, P.C., 2012. Geological and geochemical characterization of the Lower Cretaceous
Pearsall Formation, Maverick Basin, south Texas: a future shale gas resource? AAPG
Bull. 96, 1449–1482.
Hackley, P.C., Ryder, R.T., Trippi, M.H., Alimi, H., 2013. Thermal maturity of northern
Appalachian Basin Devonian shales: insights from sterane and terpane biomarkers.
Fuel 106, 455–462.
Hao, F., Zou, H., Lu, Y., 2013. Mechanisms of shale gas storage: implications for shale gas
exploration in China. AAPG Bull. 97, 1325–1346.
Hart, B.S., Macquaker, J.H.S., Taylor, K.G., 2013. Mudstone (“shale”) depositional and dia-
genetic processes: implications for seismic analyses of source-rock reservoirs.
Interpretation 1, B7–B26.
Hunt, J.M., 1979. Petroleum Geochemistry and Geology. W.H. Freeman and Company, San
Francisco (617 pp.).
Hunt, J.M., 1996. Petroleum Geochemistry and Geology, second edition. W.H. Freeman
and Company, New York (743 pp.).
Jacob, H., 1989. Classification, structure, genesis and practical importance of natural solid
oil bitumen (“migrabitumen”). Int. J. Coal Geol. 11, 65–79.
Jacob, H., 1993. Nomenclature, classification, characterization, and genesis of natural solid
bitumen (migrabitumen). In: Parnell, J., Kucha, H., Landais, P. (Eds.), Bitumens in Ore
Deposits. Springer-Verlag, New York, pp. 11–27.
Jarvie, D.M., 2012. Shale resource systems for oil and gas: part 2—shale-oil resource
systems. In: Breyer, J.A. (Ed.), Shale Reservoirs—Giant Resources for the 21st Century.
AAPG Memoir 97, pp. 89–119.
Kosakowski, P., Krajewski, M., 2014. Hydrocarbon potential of the Zechstein Main
Dolomite in the western part of the Wielkopolska platform, SW Poland: new sedi-
mentological and geochemical data. Mar. Pet. Geol. 49, 99–120.
Landis, C.R., Castaño, J.R., 1994. Maturation and bulk chemical properties of a suite of solid
hydrocarbons. Org. Geochem. 22, 137–149.
Lewan, M.D., 1983. Effects of thermal maturation on stable organic carbon isotopes as
determined by hydrous pyrolysis of Woodford Shale. Geochim. Cosmochim. Acta
47, 1471–1479.
Lewan, M.D., 1987. Petrographic study of primary petroleum migration in the Woodford
Shale and related rock units. Migration of Hydrocarbons in Sedimentary Basins: Paris,
Collection Colloques et Séminaires, Editions Technip, pp. 113–130.
Lo, H.B., Cardott, B.J., 1994. Detection of natural weathering of Upper McAlester coal and
Woodford Shale, Oklahoma, U.S.A. Org. Geochem. 22, 73–83.
Loucks, R.G., Reed, R.M., Ruppel, S.C., Jarvie, D.M., 2009. Morphology, genesis, and distri-
bution of nanometer-scale pores in siliceous mudstones of the Mississippian Barnett
Shale. J. Sediment. Res. 79, 848–861.
Loucks, R.G., Reed, R.M., Ruppel, S.C., Hammes, U., 2012. Spectrum of pore types and
networks in mudrocks and a descriptive classification for matrix-related mudrock
pores. AAPG Bull. 96, 1071–1098.
Mahlstedt, N., Horsfield, B., 2012. Metagenetic methane generation in gas shales I. Screen-
ing protocols using immature samples. Mar. Pet. Geol. 31, 27–42.
Mastalerz, M., Schimmelmann, A., Lis, G.P., Drobniak, A., Stankiewicz, A., 2012. Influence
of maceral composition on geochemical characteristics of immature shale kerogen:
insight from density fraction analysis. Int. J. Coal Geol. 103, 60–69.
McAuliffe, C.D., 1979. Oil and gas migration—chemical and physical constraints. AAPG
Bull. 63, 761–781.
Meyer, P.K., 2012. Shale source rocks a game-changer due to 8-to-1 resource potential. Oil
Gas J. 110 (5), 72–74.
Milliken, K.L., Esch, W.L., Reed, R.M., Zhang, T., 2012. Grain assemblages and strong diage-
netic overprinting in siliceous mudrocks, Barnett Shale (Mississippian), Fort Worth
Basin, Texas. AAPG Bull. 96, 1553–1578.
Milliken, K.L., Rudnicki, M., Awwiller, D.N., Zhang, T., 2013. Organic matter-hosted pore
system, Marcellus Formation (Devonian), Pennsylvania. AAPG Bull. 97, 177–200.
Pacton, M., Gorin, G.E., Vasconcelos, C., 2011. Amorphous organic matter—experimental
data on formation and the role of microbes. Rev. Palaeobot. Palynol. 166, 253–267.
Peters, K.E., Cassa, M.R., 1994. Applied source rock geochemistry. In: Magoon, L.B., Dow,
W.G. (Eds.), The Petroleum System—From Source to Trap. AAPG Memoir 60, pp.
93–120.
Reed, R.M., Loucks, R., Milliken, K.L., 2012. Heterogeneity of shape and microscale spatial
distribution in organic-matter-hosted pores of gas shales. AAPG Annual Convention
and Exhibition, Long Beach, CA, Abstract 1236631.
Romero-Sarmiento, M.-F., Ducros, M., Carpentier, B., Lorant, F., Cacas, M.-C., Pegaz-Fiornet,
S., Wolf, S., Rohais, S., Moretti, I., 2013. Quantitative evaluation of TOC, organic poros-
ity and gas retention distribution in a gas shale play using petroleum system model-
ing: application to the Mississippian Barnett Shale. Mar. Pet. Geol. 45, 315–330.
Fig. 10. Nanoporosity development at ~0.90% VRo in wispy post-oil solid bitumen network
shown in backscattered SEM (25,000×; Woodford Shale; OPL 1398).
112 B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
Ruppert, L.F., Sakurovs, R., Blach, T.P., He, L., Melnichenko, Y.B., Mildner, D.F.R., Alcantar-
Lopez, L., 2013. A USANS/SANS study of the accessibility of pores in the Barnett
Shale to methane and water. Energy Fuel 27, 772–779.
Senftle, J.T., Landis, C.R., McLaughlin, R.L., 1993. Organic petrographic approach to kerogen
characterization. In: Engel, M.H., Macko, S.A. (Eds.), Organic Geochemistry. Plenum
Press, New York, pp. 355–374.
Snowdon, L.R., 1995. Rock-eval tmax suppression: documentation and amelioration.
AAPG Bull. 79, 1337–1348.
Taylor, G.H., Teichmüller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998. Organic
Petrology. Gebrüder Borntraeger, Berlin & Stuttgart (704 pp.).
Thompson, C.L., Dembicki Jr., H., 1986. Optical characteristics of amorphous kerogens and
the hydrocarbon-generating potential of source rocks. Int. J. Coal Geol. 6, 229–249.
Thompson-Rizer, C.L., 1987. Some optical characteristics of solid bitumen in visual
kerogen preparations. Org. Geochem. 11, 385–392.
Thompson-Rizer, C.L., 1993. Optical description of amorphous kerogen in both thin
sections and isolated kerogen preparations of Precambrian to Eocene shale samples.
Precambrian Res. 61, 181–190.
Tissot, B., Welte, D.H., 1984. Petroleum Formation and Occurrence, 2nd ed. Springer-
Verlag, New York (699 pp.).
Uffmann, A.K., Littke, R., Rippen, D., 2012. Mineralogy and geochemistry of Mississippian
and Lower Pennsylvanian black shales at the northern margin of the Variscan Moun-
tain belt (Germany and Belgium). Int. J. Coal Geol. 103, 92–108.
Ungerer, P., Behar, E., Discamps, D., 1983. Tentative calculation of the overall volume
expansion of organic matter during hydrocarbon genesis from geochemistry data.
Implications for primary migration. Adv. Org. Geochem. 1981, 129–135.
Zagorski, W.A., Emery, M., Bowman, D.C., 2013. Factors control Marcellus productivity.
Am. Oil Gas Report. 54 (8), 172–180.
Zhang, T., Ellis, G.S., Ruppel, S.C., Milliken, K., Yang, R., 2012. Effect of organic-matter type
and thermal maturity on methane adsorption in shale-gas systems. Org. Geochem.
47, 120–131.
113B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113

More Related Content

What's hot

Rare earth elements and their properties and their applications in steels
Rare earth elements and their properties and their applications in steelsRare earth elements and their properties and their applications in steels
Rare earth elements and their properties and their applications in steelsHitesh Basitti
 
Concept of Source Rock Characterisation
Concept of Source Rock CharacterisationConcept of Source Rock Characterisation
Concept of Source Rock CharacterisationEmmanuelTubonemi
 
Pseudo-Source Rock Characterization
Pseudo-Source Rock CharacterizationPseudo-Source Rock Characterization
Pseudo-Source Rock CharacterizationIOSR Journals
 
Ranking Light to Heavy Rare Earth Deposits Worldwide
Ranking Light to Heavy Rare Earth Deposits WorldwideRanking Light to Heavy Rare Earth Deposits Worldwide
Ranking Light to Heavy Rare Earth Deposits WorldwideRare Earths / Rare Metals
 
P chapter 7 properties and structure of bitumens usa 1978
P chapter 7 properties and structure of bitumens usa 1978P chapter 7 properties and structure of bitumens usa 1978
P chapter 7 properties and structure of bitumens usa 1978Vainicat Rpo
 
Dolomite; Petrography & Geochemistry
Dolomite; Petrography & GeochemistryDolomite; Petrography & Geochemistry
Dolomite; Petrography & GeochemistryOmar Radwan
 
Preventing Clay Swelling
Preventing Clay Swelling Preventing Clay Swelling
Preventing Clay Swelling Omar Radwan
 
Chapter 7 metamorphic rocks
Chapter 7   metamorphic rocksChapter 7   metamorphic rocks
Chapter 7 metamorphic rocksjjones0227
 
Diversification of magma
Diversification of magmaDiversification of magma
Diversification of magmaPramoda Raj
 
A terrestrial magmatic hibonite grossite-vanadium assemblage desilication an...
 A terrestrial magmatic hibonite grossite-vanadium assemblage desilication an... A terrestrial magmatic hibonite grossite-vanadium assemblage desilication an...
A terrestrial magmatic hibonite grossite-vanadium assemblage desilication an...James AH Campbell
 
Cation exchange and it’s role on soil behaviour
Cation exchange and it’s role on soil behaviourCation exchange and it’s role on soil behaviour
Cation exchange and it’s role on soil behaviourShahram Maghami
 
Ch 11 diversification
Ch 11 diversificationCh 11 diversification
Ch 11 diversificationRaghav Gadgil
 

What's hot (19)

Rare earth elements and their properties and their applications in steels
Rare earth elements and their properties and their applications in steelsRare earth elements and their properties and their applications in steels
Rare earth elements and their properties and their applications in steels
 
Concept of Source Rock Characterisation
Concept of Source Rock CharacterisationConcept of Source Rock Characterisation
Concept of Source Rock Characterisation
 
Pseudo-Source Rock Characterization
Pseudo-Source Rock CharacterizationPseudo-Source Rock Characterization
Pseudo-Source Rock Characterization
 
Ranking Light to Heavy Rare Earth Deposits Worldwide
Ranking Light to Heavy Rare Earth Deposits WorldwideRanking Light to Heavy Rare Earth Deposits Worldwide
Ranking Light to Heavy Rare Earth Deposits Worldwide
 
P chapter 7 properties and structure of bitumens usa 1978
P chapter 7 properties and structure of bitumens usa 1978P chapter 7 properties and structure of bitumens usa 1978
P chapter 7 properties and structure of bitumens usa 1978
 
Rare earth elements summary
Rare earth elements summaryRare earth elements summary
Rare earth elements summary
 
Gtag
GtagGtag
Gtag
 
Gas hydrates
Gas hydrates Gas hydrates
Gas hydrates
 
Dolomite; Petrography & Geochemistry
Dolomite; Petrography & GeochemistryDolomite; Petrography & Geochemistry
Dolomite; Petrography & Geochemistry
 
Petroleum
PetroleumPetroleum
Petroleum
 
Preventing Clay Swelling
Preventing Clay Swelling Preventing Clay Swelling
Preventing Clay Swelling
 
Undergrad defense powerpoint
Undergrad defense powerpointUndergrad defense powerpoint
Undergrad defense powerpoint
 
Rare Earth Review - Libertas Partners LLP
Rare Earth Review - Libertas Partners LLPRare Earth Review - Libertas Partners LLP
Rare Earth Review - Libertas Partners LLP
 
Chapter 7 metamorphic rocks
Chapter 7   metamorphic rocksChapter 7   metamorphic rocks
Chapter 7 metamorphic rocks
 
Diversification of magma
Diversification of magmaDiversification of magma
Diversification of magma
 
A terrestrial magmatic hibonite grossite-vanadium assemblage desilication an...
 A terrestrial magmatic hibonite grossite-vanadium assemblage desilication an... A terrestrial magmatic hibonite grossite-vanadium assemblage desilication an...
A terrestrial magmatic hibonite grossite-vanadium assemblage desilication an...
 
Cation exchange and it’s role on soil behaviour
Cation exchange and it’s role on soil behaviourCation exchange and it’s role on soil behaviour
Cation exchange and it’s role on soil behaviour
 
Ch 11 diversification
Ch 11 diversificationCh 11 diversification
Ch 11 diversification
 
Soil colloidal chemistry
Soil colloidal chemistrySoil colloidal chemistry
Soil colloidal chemistry
 

Viewers also liked

Shale: What You May Not Know
Shale: What You May Not KnowShale: What You May Not Know
Shale: What You May Not Knowadvwealth
 
Magnum Hunter Resources Investor Presentation - July 2014
Magnum Hunter Resources Investor Presentation - July 2014Magnum Hunter Resources Investor Presentation - July 2014
Magnum Hunter Resources Investor Presentation - July 2014Marcellus Drilling News
 
Canaan Resource Partners' History and Strategy 2016
Canaan Resource Partners' History and Strategy 2016Canaan Resource Partners' History and Strategy 2016
Canaan Resource Partners' History and Strategy 2016Freddie Barela
 
Atlas Energy Barnett Shale Acquisition Presentation
Atlas Energy Barnett Shale Acquisition PresentationAtlas Energy Barnett Shale Acquisition Presentation
Atlas Energy Barnett Shale Acquisition PresentationCompany Spotlight
 
Challenger August Investor Presentation 290811
Challenger August Investor Presentation 290811Challenger August Investor Presentation 290811
Challenger August Investor Presentation 290811princeslea79
 
Energent Group: Frac Sand Trends
Energent Group: Frac Sand TrendsEnergent Group: Frac Sand Trends
Energent Group: Frac Sand TrendsTodd Bush
 
Constellation Energy Partners - Q4 2013
Constellation Energy Partners - Q4 2013Constellation Energy Partners - Q4 2013
Constellation Energy Partners - Q4 2013BRS Resources Ltd.
 
Unconventional Wells not yet completed.ppt
Unconventional Wells not yet completed.pptUnconventional Wells not yet completed.ppt
Unconventional Wells not yet completed.pptJerry Beets
 
Permian Delaware and Midland basins play.ppt
Permian Delaware and Midland basins play.pptPermian Delaware and Midland basins play.ppt
Permian Delaware and Midland basins play.pptJerry Beets
 
MNCC - 2013-10-03 - Open World Forum
MNCC - 2013-10-03 - Open World ForumMNCC - 2013-10-03 - Open World Forum
MNCC - 2013-10-03 - Open World ForumCyrille Savelief
 
Bloomberg Training Program Certification
Bloomberg Training Program CertificationBloomberg Training Program Certification
Bloomberg Training Program CertificationAdam Anson
 
Mission san luis presentation hum
Mission san luis presentation humMission san luis presentation hum
Mission san luis presentation humErin McNamara
 
Coalition Formation and Price of Anarchy in Cournot Oligopolies
Coalition Formation and Price of Anarchy in Cournot OligopoliesCoalition Formation and Price of Anarchy in Cournot Oligopolies
Coalition Formation and Price of Anarchy in Cournot OligopoliesSur Samtani
 

Viewers also liked (20)

Organic matter
Organic matterOrganic matter
Organic matter
 
Shale: What You May Not Know
Shale: What You May Not KnowShale: What You May Not Know
Shale: What You May Not Know
 
Magnum Hunter Resources Investor Presentation - July 2014
Magnum Hunter Resources Investor Presentation - July 2014Magnum Hunter Resources Investor Presentation - July 2014
Magnum Hunter Resources Investor Presentation - July 2014
 
H063771
H063771H063771
H063771
 
Es4011724
Es4011724Es4011724
Es4011724
 
Canaan Resource Partners' History and Strategy 2016
Canaan Resource Partners' History and Strategy 2016Canaan Resource Partners' History and Strategy 2016
Canaan Resource Partners' History and Strategy 2016
 
Atlas Energy Barnett Shale Acquisition Presentation
Atlas Energy Barnett Shale Acquisition PresentationAtlas Energy Barnett Shale Acquisition Presentation
Atlas Energy Barnett Shale Acquisition Presentation
 
Challenger August Investor Presentation 290811
Challenger August Investor Presentation 290811Challenger August Investor Presentation 290811
Challenger August Investor Presentation 290811
 
andrew quarles
andrew quarlesandrew quarles
andrew quarles
 
Energent Group: Frac Sand Trends
Energent Group: Frac Sand TrendsEnergent Group: Frac Sand Trends
Energent Group: Frac Sand Trends
 
Constellation Energy Partners - Q4 2013
Constellation Energy Partners - Q4 2013Constellation Energy Partners - Q4 2013
Constellation Energy Partners - Q4 2013
 
Shale gas presentation
Shale gas presentationShale gas presentation
Shale gas presentation
 
Unconventional Wells not yet completed.ppt
Unconventional Wells not yet completed.pptUnconventional Wells not yet completed.ppt
Unconventional Wells not yet completed.ppt
 
Permian Delaware and Midland basins play.ppt
Permian Delaware and Midland basins play.pptPermian Delaware and Midland basins play.ppt
Permian Delaware and Midland basins play.ppt
 
MNCC - 2013-10-03 - Open World Forum
MNCC - 2013-10-03 - Open World ForumMNCC - 2013-10-03 - Open World Forum
MNCC - 2013-10-03 - Open World Forum
 
Bloomberg Training Program Certification
Bloomberg Training Program CertificationBloomberg Training Program Certification
Bloomberg Training Program Certification
 
Americandominios
Americandominios Americandominios
Americandominios
 
Mission san luis presentation hum
Mission san luis presentation humMission san luis presentation hum
Mission san luis presentation hum
 
Project Assessment
Project AssessmentProject Assessment
Project Assessment
 
Coalition Formation and Price of Anarchy in Cournot Oligopolies
Coalition Formation and Price of Anarchy in Cournot OligopoliesCoalition Formation and Price of Anarchy in Cournot Oligopolies
Coalition Formation and Price of Anarchy in Cournot Oligopolies
 

Similar to Cardott Landis and Curtis 2015 IJCG 2015

Use of Rock-Eval pyrolysis
Use of Rock-Eval pyrolysis Use of Rock-Eval pyrolysis
Use of Rock-Eval pyrolysis selim Haouari
 
Tight carbonate reservoir characterization
Tight carbonate reservoir characterizationTight carbonate reservoir characterization
Tight carbonate reservoir characterizationKhalid Al-Khidir
 
COAL AND ITS RELATION TO OIL AND GAS
COAL AND ITS RELATION TO OIL AND GASCOAL AND ITS RELATION TO OIL AND GAS
COAL AND ITS RELATION TO OIL AND GASSayo Oladele
 
CH-8 Reservoir Rocks (1).pptx
CH-8 Reservoir Rocks (1).pptxCH-8 Reservoir Rocks (1).pptx
CH-8 Reservoir Rocks (1).pptxHamais2
 
petroleum system elements .pptx
petroleum system elements .pptxpetroleum system elements .pptx
petroleum system elements .pptxSaadTaman
 
Introduction_to_Petroleum source_rock.pptx
Introduction_to_Petroleum source_rock.pptxIntroduction_to_Petroleum source_rock.pptx
Introduction_to_Petroleum source_rock.pptxSaiNaingLinaung
 
Davis et al 2007 IJCG Petroleum Potential Of Tertiary Coals From W Indonesia
Davis et al 2007 IJCG Petroleum Potential Of Tertiary Coals From W IndonesiaDavis et al 2007 IJCG Petroleum Potential Of Tertiary Coals From W Indonesia
Davis et al 2007 IJCG Petroleum Potential Of Tertiary Coals From W Indonesiawoprd9
 
Element of Petroleum System
Element of Petroleum SystemElement of Petroleum System
Element of Petroleum SystemEric HAGENIMANA
 
Organic matter and hydeocarbon potentialpptx
Organic matter and hydeocarbon potentialpptxOrganic matter and hydeocarbon potentialpptx
Organic matter and hydeocarbon potentialpptxSubhashree Mishra
 
Petroleum geologi of south australia
Petroleum geologi of south australiaPetroleum geologi of south australia
Petroleum geologi of south australiaIntan_AP
 
Delineation of Hydrocarbon Bearing Reservoirs from Surface Seismic and Well L...
Delineation of Hydrocarbon Bearing Reservoirs from Surface Seismic and Well L...Delineation of Hydrocarbon Bearing Reservoirs from Surface Seismic and Well L...
Delineation of Hydrocarbon Bearing Reservoirs from Surface Seismic and Well L...IOSR Journals
 
unconventional-oil-and-gas-potential_compress.pptx
unconventional-oil-and-gas-potential_compress.pptxunconventional-oil-and-gas-potential_compress.pptx
unconventional-oil-and-gas-potential_compress.pptxNaserDZulfiqar
 
Tarek Saati Total
Tarek Saati TotalTarek Saati Total
Tarek Saati Totaltareksaati
 
Distribution of Possible Fatty Acids and Alkanones in some Thermally Immature...
Distribution of Possible Fatty Acids and Alkanones in some Thermally Immature...Distribution of Possible Fatty Acids and Alkanones in some Thermally Immature...
Distribution of Possible Fatty Acids and Alkanones in some Thermally Immature...Premier Publishers
 

Similar to Cardott Landis and Curtis 2015 IJCG 2015 (20)

Use of Rock-Eval pyrolysis
Use of Rock-Eval pyrolysis Use of Rock-Eval pyrolysis
Use of Rock-Eval pyrolysis
 
Tight carbonate reservoir characterization
Tight carbonate reservoir characterizationTight carbonate reservoir characterization
Tight carbonate reservoir characterization
 
COAL AND ITS RELATION TO OIL AND GAS
COAL AND ITS RELATION TO OIL AND GASCOAL AND ITS RELATION TO OIL AND GAS
COAL AND ITS RELATION TO OIL AND GAS
 
CH-8 Reservoir Rocks (1).pptx
CH-8 Reservoir Rocks (1).pptxCH-8 Reservoir Rocks (1).pptx
CH-8 Reservoir Rocks (1).pptx
 
lec 1-Introduction.pdf
lec 1-Introduction.pdflec 1-Introduction.pdf
lec 1-Introduction.pdf
 
Pubblication #1
Pubblication #1Pubblication #1
Pubblication #1
 
Oil shale..New fossil fuel for century
Oil shale..New fossil fuel for centuryOil shale..New fossil fuel for century
Oil shale..New fossil fuel for century
 
petroleum system elements .pptx
petroleum system elements .pptxpetroleum system elements .pptx
petroleum system elements .pptx
 
naturalgas_S.Karakitsiou
naturalgas_S.Karakitsiounaturalgas_S.Karakitsiou
naturalgas_S.Karakitsiou
 
Petroleum
PetroleumPetroleum
Petroleum
 
Introduction_to_Petroleum source_rock.pptx
Introduction_to_Petroleum source_rock.pptxIntroduction_to_Petroleum source_rock.pptx
Introduction_to_Petroleum source_rock.pptx
 
Davis et al 2007 IJCG Petroleum Potential Of Tertiary Coals From W Indonesia
Davis et al 2007 IJCG Petroleum Potential Of Tertiary Coals From W IndonesiaDavis et al 2007 IJCG Petroleum Potential Of Tertiary Coals From W Indonesia
Davis et al 2007 IJCG Petroleum Potential Of Tertiary Coals From W Indonesia
 
Element of Petroleum System
Element of Petroleum SystemElement of Petroleum System
Element of Petroleum System
 
Organic matter and hydeocarbon potentialpptx
Organic matter and hydeocarbon potentialpptxOrganic matter and hydeocarbon potentialpptx
Organic matter and hydeocarbon potentialpptx
 
Petroleum geologi of south australia
Petroleum geologi of south australiaPetroleum geologi of south australia
Petroleum geologi of south australia
 
Delineation of Hydrocarbon Bearing Reservoirs from Surface Seismic and Well L...
Delineation of Hydrocarbon Bearing Reservoirs from Surface Seismic and Well L...Delineation of Hydrocarbon Bearing Reservoirs from Surface Seismic and Well L...
Delineation of Hydrocarbon Bearing Reservoirs from Surface Seismic and Well L...
 
unconventional-oil-and-gas-potential_compress.pptx
unconventional-oil-and-gas-potential_compress.pptxunconventional-oil-and-gas-potential_compress.pptx
unconventional-oil-and-gas-potential_compress.pptx
 
Tarek Saati Total
Tarek Saati TotalTarek Saati Total
Tarek Saati Total
 
Distribution of Possible Fatty Acids and Alkanones in some Thermally Immature...
Distribution of Possible Fatty Acids and Alkanones in some Thermally Immature...Distribution of Possible Fatty Acids and Alkanones in some Thermally Immature...
Distribution of Possible Fatty Acids and Alkanones in some Thermally Immature...
 
NW - B.Sc. Thesis
NW - B.Sc. ThesisNW - B.Sc. Thesis
NW - B.Sc. Thesis
 

Cardott Landis and Curtis 2015 IJCG 2015

  • 1. Post-oil solid bitumen network in the Woodford Shale, USA — A potential primary migration pathway Brian J. Cardott a, ⁎, Charles R. Landis b , Mark E. Curtis c a Oklahoma Geological Survey, 100 E. Boyd St., Rm. N-131, Norman, OK 73019-0628, USA b Minerals End Inc., The Woodlands, TX 77381, USA c University of Oklahoma, Norman, OK 73019-0628, USA a b s t r a c ta r t i c l e i n f o Article history: Received 18 December 2013 Received in revised form 28 August 2014 Accepted 28 August 2014 Available online 6 September 2014 Keywords: Gas shale Shale oil Post-oil solid bitumen network Woodford Shale Porosity development Source-rock reservoir An important if not predominant component of porosity in many gas shales has been identified in organic matter. An organic network in shales has been described in the literature as organic matter (generalized term), kerogen (primary), or bitumen (secondary). Recognition of the type and origin of an organic network in shales has rele- vance in establishing the origin and timing of porosity and fracture development. The pervasive nature of the or- ganic network adopting the shape of pores in Type II kerogen-rich Woodford Shale suggests it is the residue of primary oil migration. We use the term “post-oil solid bitumen” to distinguish this bitumen occurrence from “pre-oil solid bitumen” (defined as a precursor of oil). Three forms of this post-oil solid bitumen network are rec- ognized in reflected white light at 500× magnification and confirmed in backscattered scanning electron micro- scope images at N2500× magnification, namely speckled (~1–2 μm), wispy (~2–5 μm), and connected (N5 μm). The post-oil solid bitumen network demonstrates the prior occurrence of oil generation and migration within this hydrocarbon source rock, provides porosity for hydrocarbon storage sites, and forms hydrocarbon migration pathways. © 2014 Elsevier B.V. All rights reserved. 1. Introduction The basic requirements of a shale resource system for oil and gas (source-rock reservoir of Hart et al., 2013) are appropriate organic mat- ter characteristics in the hydrocarbon source rock (e.g., organic matter type, quantity, and thermal maturity) and brittle rock fabric (Jarvie, 2012). Recognition of organic-matter (e.g., maceral) types and distribu- tion are essential to the evaluation of shales as oil and gas reservoirs. Two basic types of organic matter present in hydrocarbon source rock shales are kerogen and bitumen. Organic geochemists define bitu- men as the hydrocarbon fraction extracted with organic solvents. By this definition, pyrobitumen and kerogen are not bitumen because they are insoluble in the same organic solvents (Durand, 1980; Peters and Cassa, 1994; Tissot and Welte, 1984). In contrast, organic petrolo- gists use the term bitumen without reference to solubility, using petro- graphic features to infer its origin from its passive occurrence in rocks, mainly as void fillings (i.e., faunal, dissolution, microfracture, coating and laminar fillings; Landis and Castaño, 1994). Bitumen and residual solid hydrocarbons are important components when considering shale as a hydrocarbon source-rock reservoir. Their existence in shale is visual proof that the rock reached an appropriate level of thermal maturity to generate liquid hydrocarbons and can be used as a proxy for other visual, chemical and mineralogical compo- nents for maturation assessments, especially in rocks whose age pre- cedes terrestrial flora. Through the maturation process, the insoluble fraction of the original bitumen increases, rendering less efficient attempts to remove all of the originally generated bitumen (Landis and Castaño, 1994). The occurrence of this insoluble pyrobitumen fraction impacts both geochemical and pet- rological evaluations of the rock. The presence of these bitumens in a hy- drocarbon source rock can affect geochemical analyses (e.g., affecting the Tmax value) and petrographic analyses (e.g., appearing as vitrinite-like organic matter affecting the vitrinite reflectance analysis and visual kero- gen analysis) in isolated kerogen-concentrate pellets (Glikson et al., 1992; Hackley et al., 2013; Snowdon, 1995). Primary oil migration in a hydrocarbon source rock begins soon after oil generation when there is sufficient oil saturation (Jarvie, 2012). Ungerer et al. (1983, p. 135) concluded that oil migration “depends on the development of a continuous network of oil-wet pores.” As a result of primary oil migration and expulsion, a solid carbon residue is left behind in the shale. Taylor et al. (1998, p. 253) described this secondary organic matter as a “‘kerogen network’ (impsonite)”, using the asphaltic pyrobitumen (insoluble) vein classification of Abraham (1960). Else- where, Taylor et al. (1998, p. 84) recognized a network of solid bitumen (termed impsonite) as a “product of oil that was generated in but not expelled from the source rocks.” A key question is whether this organic material is kerogen or bitumen. Belin (1992) recognized laminated and International Journal of Coal Geology 139 (2015) 106–113 ⁎ Corresponding author. Tel.: +1 405 325 8065 E-mail address: bcardott@ou.edu (B.J. Cardott). http://dx.doi.org/10.1016/j.coal.2014.08.012 0166-5162/© 2014 Elsevier B.V. All rights reserved. Contents lists available at ScienceDirect International Journal of Coal Geology journal homepage: www.elsevier.com/locate/ijcoalgeo
  • 2. particulate kerogen networks under SEM backscattered electron mode and concluded that conventional light microscopy is needed to properly identify organic matter. Recent application of backscattered scanning electron microscopy of the Barnett Shale by Loucks et al. (2009) recog- nized the importance of organic matter in shale as an important source of microporosity, but only described the material as “organic matter”. As pointed out in the present article, much of this organic matter occurs as an organic network. Loucks et al. (2012) recognized the importance of the interconnected network of organic matter pores. Milliken et al. (2013) recognized the importance of distinguishing kerogen versus bi- tumen in the occurrence of organic matter porosity. Our purpose in this paper is to describe the occurrence, origin, and significance of the or- ganic network found in organic-rich shales. We show that this network may be observed not only at high magnification in backscattered scan- ning electron microscopy but also in reflected white light at 500× magnification. 2. Terminology Part of the problem of recognizing bitumen is that it can go by many terms: bitumen, pyrobitumen, asphalt, asphaltite, asphaltic pyrobitumen, solid bitumen, solid hydrocarbon, migrabitumen, reservoir bitumen, dead oil, and exudatinite (Abraham, 1960; Curiale, 1986; Hunt, 1979; Jacob, 1989, 1993; Landis and Castaño, 1994; Taylor et al., 1998). The term solid bitumen, following Curiale (1986), will be used in this article. A common generic classification of solid bitumen (primarily fracture-filling vein deposits of altered, once-liquid oil: asphaltite ver- sus asphaltic pyrobitumen) was given by Abraham (1960). Jacob (1989) modified this classification to include petrographic parameters (e.g., bitumen reflectance, fluorescence, solubility in immersion oil) and introduced the term migrabitumen (primarily used for vein deposits) for amorphous, secondary macerals dispersed in rocks and taking on the shape of voids. Curiale (1986, p. 559) developed a simple genetic classification of solid bitumen: pre-oil solid bitumen (defined as “early-generation (immature) products of rich source rocks”) and post- oil solid bitumen (defined as “products of the alteration of a once-liquid oil”). It is widely recognized that (pre-oil solid) bitumen forms from kerogen (Barker, 1979, p. 41; Bernard et al., 2012a; Tissot and Welte, 1984, p.176). Based on hydrous-pyrolysis experiments, Lewan (1983) demonstrated that pre-oil solid bitumen is the intermediate product be- tween kerogen and oil. The terms solid bitumen and migrabitumen were first used for non- disseminated organic matter occurring as vein deposits known as asphaltite (soluble) and asphaltic pyrobitumen (insoluble). These fracture-fillings occur as a once-liquid oil altered to a solid from near- surface, low-temperature alteration of crude oil by limited biodegrada- tion, water-washing, and devolatilization (Curiale, 1983). As discussed below, post-oil solid bitumen can also be any alteration of a once-liquid oil into a solid, including the dispersed solid residue of oil migration. Therefore, the genetic solid bitumen classification of Curiale (1986) will be used here to distinguish bitumen as a precursor of oil (pre-oil solid bi- tumen) from bitumen formed as an alteration of a once-liquid oil (post- oil solid bitumen). Thompson-Rizer (1987) described this post-oil solid bitumen network as an amorphous kerogen in strewn slides. Pre-oil solid bitumen and post-oil solid bitumen can appear similar to vitrinite in reflected white light at 500× magnification. Hackley et al. (2013) concluded that vitrinite reflectance measurements of early mature Devonian shale may erroneously include solid bitumen lower reflectance values. Landis and Castaño (1994) identified three types of solid bitumen (homogenous, granular, and coked). In addition to distinguishing vitrinite-like bitumen to exclude from the vitrinite-reflectance analysis, homogenous solid bitumen reflectance values may be used to calculate a vitrinite reflectance equivalent (VRE; Landis and Castaño, 1994). The VRE value may be used as a thermal maturity indicator when vitrinite is not present or to verify the vitrinite-reflectance value. Distinguishing solid bitumen from vitrinite is more easily accomplished in whole-rock particulate pellets than in kerogen-concentrate pellets. 3. Methods Type II kerogen-rich (oil generative organic matter) marine Woodford Shale (Late Devonian–Early Mississippian) samples from Oklahoma, USA covering a wide range of thermal maturities (0.50–6.36% vitrinite reflec- tance, VRo; random reflectance measured in non-polarized light follow- ing ASTM, 2011) have been examined in reflected white light (200× and 500×; whole-rock particulate pellets; oil immersion) using a Vickers M17 Research Microscope system equipped with a Smith illuminator (Cardott, 2012). The samples (Table 1) are from the Organic Petrography Laboratory (OPL) of the Oklahoma Geological Survey. A coal stringer sam- ple (OPL 1387) from a Woodford Shale core was used to determine the thermal maturity whereas the shale sample in the same core was used to visualize the post-oil solid bitumen network. An outcrop sample (OPL 1300) did not show signs of weathering (Lo and Cardott, 1994). The pre- dominant form of silica in the quartz-rich Woodford Shale is biogenic sil- ica formed from radiolarians (Cardott and Chaplin, 1993). Core, outcrop, and well cuttings samples that were recognized in reflected white light to contain post-oil solid bitumen forms or amorphous organic matter groundmass were selected for examination by scanning electron micros- copy (SEM). Core, outcrop, and larger chips from cuttings were initially prepared by mechanical polishing. The samples were then ion milled in a Fischione Model 1060 ion mill using Argon gas. Milling was performed with the sample rotating under two crossing ion beams at 5 kV accelerat- ing voltage for 3 h. The elevation of the ion beams was 2° above the sam- ple horizontal. Ion beam milling provides a very flat surface for imaging and preserved the microstructure of the samples with minimal artifacts. Previous work has shown that even with higher energy focused ion Table 1 Woodford Shale samples used in this study arranged by increasing vitrinite reflectance. OPL numbera Sample type County Geologic province Depth (m) Latitudeb Longitudeb VRo (%)c n VRo range (%) 1300 Grab Murray Arbuckle Mountains Surface 34.444411 −97.130651 0.53 31 0.45–0.67 1333 Core Pottawatomie Cherokee Platform 1395 34.99476 −97.05175 0.59 50 0.50–0.70 601 Core Marshall Ardmore Basin 926 34.05082 −96.64148 0.62 24 0.48–0.77 1371 Cuttings Coal Arkoma Basin 1939 34.64914 −96.39634 0.76 30 0.66–0.88 1366 Cuttings Coal Arkoma Basin 1900 34.650283 −96.356129 0.85 35 0.76–1.01 1372 Cuttings Coal Arkoma Basin 2125 34.67924 −96.36197 0.89 27 0.74–1.07 1398 Core Washington Cherokee Platform 512 36.43725 −95.92348 0.90 29 0.82–1.02 1397 Cuttings Johnston Ardmore Basin 2525 34.173551 −96.792293 0.98 26 0.85–1.10 1076 Core Okfuskee Cherokee Platform 1126 35.33143 −96.08535 1.23 48 1.10–1.47 1402 Cuttings Carter Ardmore Basin 3812 34.287887 −97.072195 1.31 24 1.19–1.45 1387 Core of coal stringer Canadian Anadarko Basin 3809 35.66829 −98.301331 1.62 50 1.51–1.74 1373 Cuttings Coal Arkoma Basin 2556 34.562946 −96.230753 1.67 23 1.38–1.99 a Oklahoma Geological Survey Organic Petrography Laboratory sample number. b NAD 83. c Random vitrinite reflectance in non-polarized light with fixed stage. 107B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
  • 3. beam milling, no significant alteration of the organic matter microstruc- ture is observed as a result of the ion milling (Curtis et al., 2010). In addi- tion, pores have been observed in organic matter that has not undergone ion milling suggesting that pores exist naturally in the organic matter of some samples. Samples were imaged in a FEI Helios Dual Beam FIB/SEM using backscattered electrons (BSE) for atomic number contrast. The accelerating voltage was 1 kV and the beam current was 0.40 nA. 4. Discussion 4.1. Occurrence of post-oil solid bitumen network Pre-oil solid bitumens are common in immature to mature (oil win- dow) hydrocarbon source rocks. They are recognized in reflected white light in whole-rock pellets by their texture (e.g., homogenous, granular, and coked), semi-translucent character with internal reflections from imbedded pyrite, and by the pyrite that occurs on their edges (Fig. 1; ASTM, 2011). In addition to amorphous blobs of pre-oil solid bitumen and fracture-filling post-oil solid bitumen, small pieces of organic matter filling voids are recognized under reflected white light at 500× magni- fication. Others have recognized this network, although not using the terminology we use here. Mahlstedt and Horsfield (2012) referred to post-oil solid bitumen as a carbon-rich pyrobitumen (“pore-occluding petroleum”) that can undergo secondary cracking to gas and conden- sate at N1.1% VRo. Landis and Castaño (1994) described residual solid hydrocarbons occurring as intergranular pore fillings (b10 μm). The void-filling occurrence of this material suggests that it is the solid resi- due of primary oil migration (e.g., post-oil solid bitumen). The term post-oil solid bitumen, as we use it here, has no inference for solubility. Three post-oil solid bitumen forms are recognized in reflected white light at 500× in whole rock particulate pellets: speckled (~1–2 μm; Fig. 2a); wispy (~2–5 μm; Fig. 2b); and connected (N5 μm; Fig. 2c). The sizes are relative and should not be considered specific. Speckled and wispy are mostly isolated occurrences in the rock rath- er than as a connected network. The speckled post-oil solid bitumen network, where it occurs, is near the magnification limit of the light mi- croscope (~1–2 μm) and is used mostly as a presence or absence indica- tor of oil generation and migration. Speckled post-oil solid bitumen must be carefully distinguished from clay minerals. Speckled and wispy types are not exclusive — that is, they do not occur in shale as ei- ther one or the other. Grains that contain the wispy post-oil solid bitu- men network also will contain areas of speckled post-oil solid bitumen network (Fig. 3). Speckled and wispy forms are easy to see but difficult to photograph in reflected white light (500×) because of focus issues and interference from pyrite. The post-oil solid bitumen shapes and sizes that form a network are confirmed and best viewed in BSE images. BSE imaging is sensitive to the atomic number of the sample material. This results in the low atomic number organic matter (mostly carbon) appearing dark gray whereas higher atomic number minerals such as quartz, carbonates, and pyrite exhibit progressively higher grayscale values (Figs. 4–5). Belin (1994) noted that even though organic matter types cannot be identified in BSE images, rela- tionships of organic matter and minerals are revealed with fine resolu- tion. Although they did not use the solid bitumen terminology described here, Bernard et al. (2012b, p. 7) recognized the speckled and wispy post-oil solid bitumen network by using scanning transmission X-ray microscopy and transmission electron microscopy of the Barnett Shale. Similar to what we see in the Woodford Shale, their work indicat- ed that “Organic matter appears as micron sized angular organic grains irregularly distributed within the mineral matrix or as organic masses filling intergranular porosity and exhibiting smoothly curved concave surfaces.” 4.2. Origin of post-oil solid bitumen network Primary oil migration occurs within the hydrocarbon source rock (Cordell, 1972). McAuliffe (1979) proposed that primary oil migration occurs in a 3D kerogen network. However, the occurrence of secondary, amorphous dispersed organic matter filling voids suggests that this organic matter was once a liquid, and thus not kerogen. The pervasive nature and relationship with fracture filling bitumen indicate that it is post-oil solid bitumen. The mechanisms of primary oil migration through a kerogen network proposed by McAuliffe (1979) hold true for a post-oil solid bitumen network. Much of the generated oil does not migrate out of the rock. Meyer (2012, p. 72) indicated that “for every barrel of crude oil in conventional reservoirs … there are 8 bbl of potentially producible oil equivalents remaining in the source rock” and “Speculative estimates of just how much generated oil remains in shale source rocks range between 45% and 95% depending on the geol- ogy of the formation and the quality of the estimate.” Some of the shale- hosted oil will result in a carbon residue (possibly the same as residual oil of Fan et al., 2012). Hunt (1996, p. 597–598, see references within on p. 598) recognized a “refractory bitumen” or “pyrobitumen residue” retained in the source rock. 4.3. Lowest thermal maturity with post-oil solid bitumen network The most common organic-matter type in low thermal maturity Type II kerogen-rich shales and boghead coals is amorphous organic matter (AOM) (Mastalerz et al., 2012; Thompson-Rizer, 1993). This pri- mary maceral, AOM, is derived from degraded, unidentifiable precursor organisms (Pacton et al., 2011). AOM is equivalent to the term bituminite (ASTM, 2011). Lewan (1987) reported that amorphous Fig. 1. (A) Homogenous texture in semi-translucent pre-oil solid bitumen (dark gray material in center of photomicrograph) showing internal reflections from pyrite (reflected white light, 500×; whole-rock particulate pellet; Woodford Shale; OPL 1333; 0.59% VRo). (B) Granular texture in pre-oil solid bitumen (reflected white light, 500×; whole-rock particulate pellet; Woodford Shale; OPL 1076; 1.23% VRo). Solid bitumen classification is modified from Curiale (1986) and Landis and Castaño (1994). 108 B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
  • 4. Type II kerogen comprised N80 vol% of isolated kerogen from Woodford Shale samples in Oklahoma. There are several classifications of AOM. Thompson and Dembicki (1986) recognized four types of AOM related to hydrocarbon-generating potential (Types A–D). Taylor et al. (1998, p. 250) also recognized four types of unstructured organic matter (bituminite). Senftle et al. (1993) indicated that fluorescing AOM (fluoramorphinite) can be distinguished from nonfluorescing AOM (hebamorphinite) up to 1.1% VRo in an estimate of oil and gas potential. Only the fluorescing type of AOM (fluoramorphinite and types A and D) is considered a source of pre-oil solid bitumen and oil. AOM is best recognized using strewn slides in transmitted white light and reflected fluorescent light at 500× magnification. The distribution of AOM in shale forms an organic network (Figs. 6, 7). The AOM network could be misidentified as the post-oil solid bitumen network. Post-oil solid bitumen ultimately forms from oil generated in the oil window. The lowest thermal maturity containing a post-oil solid bitumen network is uncertain because it may be confused with AOM (a primary maceral). The post-oil solid bitumen network could have de- veloped preferentially along the AOM network. Lewan (1987) described primary oil migration occurring along a continuous bitumen network formed from kerogen and impregnating AOM. He described the devel- opment of an opaque pyrobitumen groundmass (e.g., post-oil solid bitu- men network) carbonized from bitumen and retained oil. The lowest thermal maturity where the post-oil solid bitumen net- work is observed in the Woodford Shale is 0.76% VRo near the middle of the oil window (Fig. 8). Below this thermal maturity, the organic net- work could be AOM. 4.4. Significance of post-oil solid bitumen network The presence of post-oil solid bitumen demonstrates that oil was generated in or migrated through the rock even though the rock is cur- rently at a higher thermal maturity than the oil window (Thompson- Rizer, 1987). Recent applications of BSE images have not only revealed the occurrence of a network of organic matter, but also the development Fig. 2. (A) Speckled (~1-2 μm) post-oil solid bitumen network (white arrow; reflected white light, 500×; Woodford Shale; OPL 1402; 1.31% VRo). (B) Wispy (~2–5 μm) post- oil solid bitumen network (white arrow; reflected white light, 500×; Woodford Shale; OPL 1372; 0.89% VRo). (C) Connected (N 5 μm) post-oil solid bitumen network (white arrow; reflected white light, 500×; Woodford Shale; OPL 1366; 0.85% VRo). P = pyrite. Fig. 3. Speckled (~1–2 μm) and wispy (~2–5 μm) post-oil solid bitumen network in the same grain (reflected white light, 500×; Woodford Shale; OPL 1387; 1.62% VRo). Fig. 4. Speckled (~1–2 μm; white arrow) and wispy (~2–5 μm; black arrow) post-oil solid bitumen network in backscattered SEM (5000×; Woodford Shale; OPL 1397e; 0.98% VRo). 109B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
  • 5. of secondary nanoporosity (i.e., pores several nanometers in size; Loucks et al., 2009; Ruppert et al., 2013; Fig. 9b). Nanoporosity in organics has been described in the literature as de- veloping at N0.6% to ~0.9% VRo primarily in post-oil solid bitumen. Curtis et al. (2012a) reported that the development of secondary nanoporosity is related to both thermal maturity (beginning about 0.9% VRo; Fig. 10) and organic-matter type (e.g., post-oil solid bitumen). In contrast, Reed et al. (2012) reported nanopore development in or- ganic matter beginning at about 0.8% VRo. Loucks et al. (2012) and Zhang et al. (2012) reported nanopore development in organic matter N0.60% VRo. Romero-Sarmiento et al. (2013) attributed nanoporosity development to the maturation of kerogen in the Barnett Shale begin- ning at about 0.7% VRo. Bernard et al. (2012a) reported nanoporous pyrobitumen (e.g., post-oil solid bitumen) in a Posidonia Shale sample at 1.45% VRo, but no organic nanoporosity in samples at 0.5% and 0.85% VRo. Milliken et al. (2012) demonstrated that secondary porosity develops primarily in intergranular organic matter (e.g., post-oil solid bitumen) instead of within particulate organic matter (e.g., kerogen) and that porosity increases with increasing total organic carbon content. Hao et al. (2013, p. 1342) concluded that “gas sorption in organic-rich shales is mainly associated with micropores” (b2 nm). In addition to the biogenic-silica-rich Woodford Shale, a post-oil solid bitumen network has also been observed in other shales, including the Barnett, Haynesville, and Horn River shales (Curtis et al., 2012a). Although not using the post-oil solid bitumen network terminology used here, others have observed the network in other formations. Bernard et al. (2012a) recognized aliphatic-rich bitumens (e.g., pre-oil solid bitumen) and aromatic-rich pyrobitumens (e.g., post-oil solid bitumen) in an overmature (1.45% VRo) sample of the Posidonia Shale. Bernard et al. (2012b) recognized pre-oil solid bitumen (derived from thermally degraded kerogen) and post-oil solid bitumen (nanoporous pyrobitumen resulting from the secondary thermal crack- ing of retained oil) in the Barnett Shale. Milliken et al. (2012) recognized secondary porosity in “organic particulate debris and solid bitumen” using field-emission scanning electron microscope images of Ar ion- milled surfaces of the Barnett Shale (Mississippian). The predominant pore-filling organic matter, interpreted as solid bitumen, was recog- nized as originating as a liquid hydrocarbon. Hackley (2012) recognized an interconnected (post-oil) solid bitumen network in whole-rock par- ticulate pellets of argillaceous lime wackestones and mudstones of the Lower Cretaceous Pearsall Formation. Uffmann et al. (2012) described a (post-oil) solid bitumen network in whole-rock pellets of high ther- mal maturity Mississippian and Pennsylvanian black shales from Germany and Belgium. Fishman et al. (2012) did not recognize bitumen or pyrobitumen in the Kimmeridge Clay Formation and concluded that petroleum storage potential was attributed to inorganic pores. In con- trast, Fishman et al. (2013) recognized nanoporosity in high maturity (~1.2% VRo) migrated bitumen from Eagle Ford Shale core samples equivalent to the post-oil solid bitumen terminology used here. Kosakowski and Krajewski (2014, their Fig. 11E) recognized a post-oil solid bitumen network in carbonates in Poland. Organic pores are not only sites of methane storage by adsorption to the pore walls (Hackley, 2012; Zhang et al., 2012), but also provide mi- gration pathways for production of natural gas. 3-D reconstructions of Focused Ion Beam/SEM tomography samples illustrate the distribution and connection of nanopores in the post-oil solid bitumen network (Curtis et al., 2012b, their Fig. 8). Microfractures that connect to the post-oil solid bitumen network can be seen in BSE images (Fig. 9a) dem- onstrating the preferred fracture pattern following zones of weakness through the bitumen. The microfractures (formed either naturally or in- duced by hydrofracturing) contribute to the rock permeability. Zagorski et al. (2013, p. 172) recognized that “The observed intraorganic porosity displays a high degree of connectivity and is responsible for a significant portion of the Marcellus Shale's productivity and gas in place.” Applied to gas shales, Blood (2011, p. 56) recognized that “Organic particles are the sites of adsorbed gas, and amorphous organic matter and bitumen represent the dominant sites of porosity development within the Marcellus.” As recognized by Belin (1992) for a kerogen Fig. 5. Connected (N 5 μm) post-oil solid bitumen network (white arrow) in backscattered SEM (2540×; Woodford Shale; OPL 1402; 1.31% VRo). Fig. 6. Amorphous organic matter (AOM; also referred to as bituminite) groundmass (dark gray material) in Woodford Shale marine boghead coal in backscattered SEM (800×; OPL 1300e; 0.53% VRo). Fig. 7. Amorphous organic matter (AOM) matrix in backscattered SEM (7500×; Woodford Shale; OPL 601; 0.62% VRo). 110 B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
  • 6. network, the post-oil solid bitumen network may be discontinuous (e.g., speckled and wispy) or continuous (e.g., connected) based on total organic carbon content and available porosity. Lewan (1987) ob- served that the (post-oil solid) bitumen network occurs in amorphous Type II kerogen-rich shales. We are in agreement with Lewan (1987, p. 128) that “Impregnation of the groundmass with [pre-oil solid] bitu- men to form a continuous network appears to be a prerequisite for the expulsion of generated oil.” The pervasive post-oil solid bitumen resi- due left behind during primary oil migration provides nanoporosity sites for hydrocarbon storage and microfracture permeability and path- ways for hydrocarbon production. 5. Summary and conclusions Primary oil migration in shales leaves behind a solid carbon residue in available porosity that we describe as a post-oil solid bitumen network. Development of the network is dependent on kerogen-type and total- organic-carbon content. This study reports the development of a post- oil solid bitumen in one of the most recognized conventional hydrocar- bon source rocks in North America. The Woodford Shale lithofacies are broadly characterized as Type II kerogen assemblages. The initial devel- opment of unconventional reservoirs focused on Paleozoic rocks but fur- ther work on a wider range of kerogen types is needed to assess the proposed concept of post-oil solid bitumen networks more broadly. The network is recognized in our Type II kerogen-rich Woodford Shale samples in reflected white light at 500× magnification as speckled (~1–2 μm), wispy (~2–5 μm), and connected (N5 μm) forms. (The sizes are relative and should not be considered specific.) Speckled and wispy forms are mostly isolated occurrences in the rock rather than as a con- nected network. The speckled post-oil solid bitumen network is near the magnification limit of the light microscope (~1–2 μm) and is used mostly as a presence or absence indicator of oil generation and primary oil migration. The small size should not be confused with clay minerals. Speckled and wispy networks often both occur in the same rock. The post-oil solid bitumen shapes and sizes that form a network are con- firmed and best viewed in backscattered scanning electron microscope images. The post-oil solid bitumen networks in the Woodford Shale demon- strate: (1) that the rock has generated oil; (2) that oil has migrated through the rock; (3) that secondary nanoporosity, developing Fig. 8. Lowest thermal maturity (0.76% VRo) Woodford Shale sample that contains post-oil solid bitumen network (connected; N 5 μm) in (A) reflected white light (500×) with pre-oil solid bitumen homogenous form, and (B) backscattered SEM (dark gray; 15,000×; OPL 1371). The organic matter does not contain nanopores. Fig. 9. (A) Microfracture development (possibly caused by sample handling) along post-oil solid bitumen network from Woodford Shale cuttings sample of highest thermal maturity (1.67% VRo) condensate well in backscattered SEM (5000×; OPL 1373) demonstrates preferred zones of weakness within bitumen for fracture formation; (B) Nanoporosity in wispy (~2–5 μm) post-oil solid bitumen network (dark gray) in backscattered SEM (20,000×; OPL 1373). 111B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
  • 7. beginning at ~0.9% VRo, provides storage sites for hydrocarbons; (4) that these sites are zones of weakness for the formation of microfractures; and (5) that they also form migration pathways for hydrocarbons. Acknowledgments The authors gratefully acknowledge reviews by Joseph A. Curiale and an anonymous reviewer that improved the manuscript. References Abraham, H., 1960. sixth edition. Asphalts and Allied Substances 5 vol. Van Nostrand Company, Inc., New York. ASTM, 2011. Standard test method for microscopical determination of the reflectance of vitrinite dispersed in sedimentary rocks: West Conshohocken, PA, American Society for Testing and Materials International. Annual book of ASTM standards: Petroleum products, lubricants, and fossil fuels; gaseous fuels; coal and coke, sec. 5, v. 5.06, D7708-11, pp. 823–830 http://dx.doi.org/10.1520/D7708-11. Barker, C., 1979. Organic geochemistry in petroleum exploration. American Association of Petroleum Geologists Education Course Note 10, (159 pp.). Belin, S., 1992. Application of backscattered electron imaging to the study of source rocks microtextures. Org. Geochem. 18, 333–346. Belin, S., 1994. Backscattered electron imaging applied to source rock sedimentology: a comparision with conventional methods in organic petrology. Bulletin Des Centres De Recherches Exploration-Production, v. 18, Special Publication, pp. 165–187. Bernard, S., Horsfield, B., Schulz, H.-M., Wirth, R., Schreiber, A., Sherwood, N., 2012a. Geochemical evolution of organic-rich shales with increasing maturity: a STXM and TEM study of the Posidonia Shale (Lower Toarcian, northern Germany). Mar. Pet. Geol. 31, 70–89. Bernard, S., Wirth, R., Schreiber, A., Schulz, H.-M., Horsfield, B., 2012b. Formation of nanoporous pyrobitumen residues during maturation of the Barnett Shale (Fort Worth Basin). Int. J. Coal Geol. 103, 3–11. Blood, D.R., 2011. Sequence stratigraphy crucial to lateral placement in Marcellus Shale play, part two. Am. Oil Gas Report. 54 (8), 52–60. Cardott, B.J., 2012. Thermal maturity of Woodford Shale gas and oil plays, Oklahoma, USA. Int. J. Coal Geol. 103, 109–119. Cardott, B.J., Chaplin, J.R., 1993. Guidebook for selected stops in the western Arbuckle Mountains, southern Oklahoma. Okla. Geol. Surv. Spec. Publ. 93-3 (55 pp.). Cordell, R.J., 1972. Depths of oil origin and primary migration: a review and critique. AAPG Bull. 56, 2029–2067. Curiale, J.A., 1983. Petroleum occurrences and source-rock potential of the Ouachita Mountains, southeastern Oklahoma. Bull. Okla. Geol. Surv. 135, 65. Curiale, J.A., 1986. Origin of solid bitumens, with emphasis on biological marker results. Org. Geochem. 10, 559–580. Curtis, M.E., Ambrose, R.J., Sondergeld, C.H., Rai, C.S., 2010. Structural characterization of gas shales on the micro- and nano-scales. Canadian Unconventional Resources and International Petroleum Conference, Calgary, Alberta, Canada, SPE 137693. Curtis, M.E., Cardott, B.J., Sondergeld, C.H., Rai, C.S., 2012a. Development of organic poros- ity in the Woodford Shale with increasing thermal maturity. Int. J. Coal Geol. 103, 26–31. Curtis, M.E., Sondergeld, C.H., Ambrose, R.J., Rai, C.S., 2012b. Microstructural investigation of gas shales in two and three dimensions using nanometer-scale resolution imaging. AAPG Bull. 96, 665–677. Durand, B., 1980. Sedimentary organic matter and kerogen. Definition and quantitative importance of kerogen. In: Durand, B. (Ed.), Kerogen. Insoluble Organic Matter from Sedimentary Rocks. Editions Technip, Paris, pp. 13–34. Fan, B., Pang, X., Zhang, X., Zhang, J., 2012. Distinction between oil expulsion history and gas expulsion history. Nat. Resour. Res. 21, 233–243. Fishman, N.S., Hackley, P.C., Lowers, H.A., Hill, R.J., Evenhoff, S.O., Eberl, D.D., Blum, A.E., 2012. The nature of porosity in organic-rich mudstones of the Upper Jurassic Kimmeridge Clay Formation, North Sea, offshore United Kingdom. Int. J. Coal Geol. 103, 32–50. Fishman, N.S., Guthrie, J., Honarpour, M., 2013. The stratigraphic distribution of hydrocar- bon storage and its effect on producible hydrocarbons in the Eagle Ford Formation, south Texas. Unconventional Resources Technology Conference, SPE-AAPG-SEG, Den- ver, CO, Paper 1579007 (6 pp.). Glikson, M., Taylor, D., Morris, D., 1992. Lower Paleozoic and Precambrian petroleum source rocks and the coalification path of alginite. Org. Geochem. 18, 881–897. Hackley, P.C., 2012. Geological and geochemical characterization of the Lower Cretaceous Pearsall Formation, Maverick Basin, south Texas: a future shale gas resource? AAPG Bull. 96, 1449–1482. Hackley, P.C., Ryder, R.T., Trippi, M.H., Alimi, H., 2013. Thermal maturity of northern Appalachian Basin Devonian shales: insights from sterane and terpane biomarkers. Fuel 106, 455–462. Hao, F., Zou, H., Lu, Y., 2013. Mechanisms of shale gas storage: implications for shale gas exploration in China. AAPG Bull. 97, 1325–1346. Hart, B.S., Macquaker, J.H.S., Taylor, K.G., 2013. Mudstone (“shale”) depositional and dia- genetic processes: implications for seismic analyses of source-rock reservoirs. Interpretation 1, B7–B26. Hunt, J.M., 1979. Petroleum Geochemistry and Geology. W.H. Freeman and Company, San Francisco (617 pp.). Hunt, J.M., 1996. Petroleum Geochemistry and Geology, second edition. W.H. Freeman and Company, New York (743 pp.). Jacob, H., 1989. Classification, structure, genesis and practical importance of natural solid oil bitumen (“migrabitumen”). Int. J. Coal Geol. 11, 65–79. Jacob, H., 1993. Nomenclature, classification, characterization, and genesis of natural solid bitumen (migrabitumen). In: Parnell, J., Kucha, H., Landais, P. (Eds.), Bitumens in Ore Deposits. Springer-Verlag, New York, pp. 11–27. Jarvie, D.M., 2012. Shale resource systems for oil and gas: part 2—shale-oil resource systems. In: Breyer, J.A. (Ed.), Shale Reservoirs—Giant Resources for the 21st Century. AAPG Memoir 97, pp. 89–119. Kosakowski, P., Krajewski, M., 2014. Hydrocarbon potential of the Zechstein Main Dolomite in the western part of the Wielkopolska platform, SW Poland: new sedi- mentological and geochemical data. Mar. Pet. Geol. 49, 99–120. Landis, C.R., Castaño, J.R., 1994. Maturation and bulk chemical properties of a suite of solid hydrocarbons. Org. Geochem. 22, 137–149. Lewan, M.D., 1983. Effects of thermal maturation on stable organic carbon isotopes as determined by hydrous pyrolysis of Woodford Shale. Geochim. Cosmochim. Acta 47, 1471–1479. Lewan, M.D., 1987. Petrographic study of primary petroleum migration in the Woodford Shale and related rock units. Migration of Hydrocarbons in Sedimentary Basins: Paris, Collection Colloques et Séminaires, Editions Technip, pp. 113–130. Lo, H.B., Cardott, B.J., 1994. Detection of natural weathering of Upper McAlester coal and Woodford Shale, Oklahoma, U.S.A. Org. Geochem. 22, 73–83. Loucks, R.G., Reed, R.M., Ruppel, S.C., Jarvie, D.M., 2009. Morphology, genesis, and distri- bution of nanometer-scale pores in siliceous mudstones of the Mississippian Barnett Shale. J. Sediment. Res. 79, 848–861. Loucks, R.G., Reed, R.M., Ruppel, S.C., Hammes, U., 2012. Spectrum of pore types and networks in mudrocks and a descriptive classification for matrix-related mudrock pores. AAPG Bull. 96, 1071–1098. Mahlstedt, N., Horsfield, B., 2012. Metagenetic methane generation in gas shales I. Screen- ing protocols using immature samples. Mar. Pet. Geol. 31, 27–42. Mastalerz, M., Schimmelmann, A., Lis, G.P., Drobniak, A., Stankiewicz, A., 2012. Influence of maceral composition on geochemical characteristics of immature shale kerogen: insight from density fraction analysis. Int. J. Coal Geol. 103, 60–69. McAuliffe, C.D., 1979. Oil and gas migration—chemical and physical constraints. AAPG Bull. 63, 761–781. Meyer, P.K., 2012. Shale source rocks a game-changer due to 8-to-1 resource potential. Oil Gas J. 110 (5), 72–74. Milliken, K.L., Esch, W.L., Reed, R.M., Zhang, T., 2012. Grain assemblages and strong diage- netic overprinting in siliceous mudrocks, Barnett Shale (Mississippian), Fort Worth Basin, Texas. AAPG Bull. 96, 1553–1578. Milliken, K.L., Rudnicki, M., Awwiller, D.N., Zhang, T., 2013. Organic matter-hosted pore system, Marcellus Formation (Devonian), Pennsylvania. AAPG Bull. 97, 177–200. Pacton, M., Gorin, G.E., Vasconcelos, C., 2011. Amorphous organic matter—experimental data on formation and the role of microbes. Rev. Palaeobot. Palynol. 166, 253–267. Peters, K.E., Cassa, M.R., 1994. Applied source rock geochemistry. In: Magoon, L.B., Dow, W.G. (Eds.), The Petroleum System—From Source to Trap. AAPG Memoir 60, pp. 93–120. Reed, R.M., Loucks, R., Milliken, K.L., 2012. Heterogeneity of shape and microscale spatial distribution in organic-matter-hosted pores of gas shales. AAPG Annual Convention and Exhibition, Long Beach, CA, Abstract 1236631. Romero-Sarmiento, M.-F., Ducros, M., Carpentier, B., Lorant, F., Cacas, M.-C., Pegaz-Fiornet, S., Wolf, S., Rohais, S., Moretti, I., 2013. Quantitative evaluation of TOC, organic poros- ity and gas retention distribution in a gas shale play using petroleum system model- ing: application to the Mississippian Barnett Shale. Mar. Pet. Geol. 45, 315–330. Fig. 10. Nanoporosity development at ~0.90% VRo in wispy post-oil solid bitumen network shown in backscattered SEM (25,000×; Woodford Shale; OPL 1398). 112 B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113
  • 8. Ruppert, L.F., Sakurovs, R., Blach, T.P., He, L., Melnichenko, Y.B., Mildner, D.F.R., Alcantar- Lopez, L., 2013. A USANS/SANS study of the accessibility of pores in the Barnett Shale to methane and water. Energy Fuel 27, 772–779. Senftle, J.T., Landis, C.R., McLaughlin, R.L., 1993. Organic petrographic approach to kerogen characterization. In: Engel, M.H., Macko, S.A. (Eds.), Organic Geochemistry. Plenum Press, New York, pp. 355–374. Snowdon, L.R., 1995. Rock-eval tmax suppression: documentation and amelioration. AAPG Bull. 79, 1337–1348. Taylor, G.H., Teichmüller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998. Organic Petrology. Gebrüder Borntraeger, Berlin & Stuttgart (704 pp.). Thompson, C.L., Dembicki Jr., H., 1986. Optical characteristics of amorphous kerogens and the hydrocarbon-generating potential of source rocks. Int. J. Coal Geol. 6, 229–249. Thompson-Rizer, C.L., 1987. Some optical characteristics of solid bitumen in visual kerogen preparations. Org. Geochem. 11, 385–392. Thompson-Rizer, C.L., 1993. Optical description of amorphous kerogen in both thin sections and isolated kerogen preparations of Precambrian to Eocene shale samples. Precambrian Res. 61, 181–190. Tissot, B., Welte, D.H., 1984. Petroleum Formation and Occurrence, 2nd ed. Springer- Verlag, New York (699 pp.). Uffmann, A.K., Littke, R., Rippen, D., 2012. Mineralogy and geochemistry of Mississippian and Lower Pennsylvanian black shales at the northern margin of the Variscan Moun- tain belt (Germany and Belgium). Int. J. Coal Geol. 103, 92–108. Ungerer, P., Behar, E., Discamps, D., 1983. Tentative calculation of the overall volume expansion of organic matter during hydrocarbon genesis from geochemistry data. Implications for primary migration. Adv. Org. Geochem. 1981, 129–135. Zagorski, W.A., Emery, M., Bowman, D.C., 2013. Factors control Marcellus productivity. Am. Oil Gas Report. 54 (8), 172–180. Zhang, T., Ellis, G.S., Ruppel, S.C., Milliken, K., Yang, R., 2012. Effect of organic-matter type and thermal maturity on methane adsorption in shale-gas systems. Org. Geochem. 47, 120–131. 113B.J. Cardott et al. / International Journal of Coal Geology 139 (2015) 106–113