You are on page 1of 7

Kerogen

From AAPG Wiki


(Redirected from Kerogen types)

Depositional environment is the


dominant factor in determining the Exploring for Oil and Gas Traps
types of organic matter found in a
rock. Only two types of organic
matter are found in rocks: land
derived and aquatic algae derived.
Heat and pressure convert organic
matter into a substance called
humin and then into kerogen. Time
and temperature convert kerogen
into petroleum.
Series Treatise in Petroleum Geology
Part Critical elements of the petroleum system
Contents Chapter Evaluating source rocks
Author Carol A. Law
◾ 1 What is kerogen?
◾ 2 Kerogen quality Link Web page
◾ 3 Parameter differences (http://archives.datapages.com/data/specpubs/beaumont/ch06/ch06.htm)
◾ 4 Transformation ratio Store AAPG Store (http://store.aapg.org/detail.aspx?id=545)
◾ 4.1 Example
◾ 5 Kerogen types
◾ 5.1 Kerogen Type I
◾ 5.2 Kerogen Type II
◾ 5.3 Kerogen Type III
◾ 5.4 Kerogen Type IV
◾ 6 See also
◾ 7 References
◾ 8 External links

What is kerogen?
Geochemists[1][2] define kerogen as the fraction of sedimentary organic constituent of sedimentary rocks that is
insoluble in the usual organic solvents. Kerogens are composed of a variety of organic materials, including algae,
pollen, wood, vitrinite, and structureless material. The types of kerogens present in a rock largely control the type
of hydrocarbons generated in that rock. Different types of kerogen contain different amounts of hydrogen relative
to carbon and oxygen. The hydrogen content of kerogen is the controlling factor for oil vs. gas yields from the
primary hydrocarbon-generating reactions.

Structured kerogens include woody, herbaceous, vitrinite, and inertinite. Amorphous kerogens are by far the most
prevalent and include most of the algal material.
Kerogen quality
The type of kerogen present determines source rock quality.
The more oil prone a kerogen, the higher its quality. Four
basic types of kerogen are found in sedimentary rocks. A
single type or a mixture of types may be present in a source
rock. The table below lists and defines these four basic
kerogen types.

Predominant Amount Typical


Kerogen
hydrocarbon of depositional
type
potential hydrogen environment
I Oil prone Abundant Lacustrine
II Oil and gas prone Moderate Marine
III Gas prone Small Terrestrial
Neither
(primarily
IV composed of None Terrestrial(?)
vitrinite) or inert
material
Van Krevelen diagram.[3]

Parameter differences
The table below shows examples of the relationships between hydrocarbon generation zones, maturity, and
transformation ratio for standard types II and III kerogens, based on a specific burial and thermal history model.
The most significant difference is in the depth to the onset of oil generation, where 1000 m separates the top of
the oil windows of these two kerogen types.

Vitrinite Vitrinite Present-


Hydrocarbon Present-
refl., % refl., % Transformation Transformation day depth,
generation day depth,
Ro, Type Ro, Type ratio, %, Type II ratio, %, Type III m, Type
zone m, Type II
II III III
Onset oil 0.55 0.85 5 12 2200 3200
Onset peak rate
0.65 1.00 17 31 2600 3500
generation
Onset
gas/cracking 0.95 1.35 88 64 3400 4050
liquids

Transformation ratio
When we compare hydrocarbon generation curves and transformation ratio curves from 1-D models, we can
develop a relationship in a way similar to that for generation-maturity. If vitrinite reflectance data are available,
the relationship between transformation ratio and maturity can be used to predict (1) percentage of kerogen that
has generated hydrocarbons at a given depth and (2) hydrocarbon yields.
Example

Based on Figure 1, we determine that at a depth of 2.6


km the modeled well is presently in the oil generation
zone and approximately 25% of the kerogen in the
source rocks at this depth has generated
hydrocarbons. We know from the hydrocarbon
generation-maturity relationship that at 2.6 km this
well has a vitrinite reflectance ( Ro) of 0.7%. If
another well in the basin contains similar source rocks Figure 1 . Copyright: results of Genex 1-D basin modeling
and has a maturity of 0.7% Ro at 3.7 km, then we can software, courtesy Institute Français du Petrole.
predict that the section at 3.7 km is mature for liquid
generation and has generated a liquid hydrocarbons,
converting approximately 25% of its kerogen to hydrocarbons.

Kerogen types
The
following Source and Migration Processes and Evaluation Techniques
descriptions
of kerogen
types indicate
their
biological
input,
stratigraphy,
and
depositional
processes that Series Treatise Handbook
control their
Part Petroleum Generation and Migration
oil-generative
properties. Chapter Petroleum Source Rocks and Organic Facies
Kerogen Author S. R. Jacobson
types are
defined on Link Web page
H/C and O/C (http://archives.datapages.com/data/specpubs/geochem1/data/a037/a037/0001/0000/0003.htm)
values (or HI
PDF PDF file
and OI from
(http://archives.datapages.com/data/specpubs/geochem1/images/a037/a0370001/0000/00030.pdf)
Rock-Eval).
(requires access)
In thermally
immature Store AAPG Store (http://store.aapg.org/detail.aspx?id=436)
samples, the
chemically
extreme kerogen types I and IV (and therefore the equivalent organic facies A and D) contain macerals having
relatively uniform chemical properties. These end-members are dominated by the most and least hydrogen-rich
constituents. Other kerogen types (and therefore their equivalent organic facies) are frequently mixtures of
macerals. Microscopy is the method of choice for distinguishing the constituents of mixed organic matter
assemblages.
Before enumerating the criteria for discriminating kerogen types, it is important to consider the "mineral matrix
effect." Some mineral (polar clay) constituents retard the release of hydrocarbons from powdered whole rock
samples during Rock-Eval pyrolysis, under-evaluating the quantity, quality, and thermal maturation data.
Although this factor, the mineral matrix effect, is well known to organic geochemists, it is frequently overlooked
when interpreting Rock-Eval-dependent values used to determine kerogen type and organic facies. The mineral
matrix effect occurs when polar clays react with polar organic molecules during the nonhydrous Rock-Eval
procedure.[4][5][6][7][8][9][10][11]

Pioneers of pyrolysis found that some minerals inhibit


hydrocarbon expulsion during whole-rock pyrolysis
and not during kerogen pyrolysis.[4][5][7] The effect of
different matrix constituents[4][5][7][8] varies from
strongest to weakest: illite > Ca-bentonite > kaolinite
> Na-bentonite > calcium carbonate > gypsum.[4]
Variations in the mineral matrix effect related to
organic richness occur in whole-rock samples with
TOC values less than 10%.[4][5][7]

Geological thermal maturation processes differ from


those of Rock-Eval pyrolysis. Whole-rock Rock-Eval
samples are heated rapidly in an anhydrous
environment. Geological burial processes cause clays
to undergo physical and chemical alteration usually
preceding the slow and systematic thermal conversion
(generation) of kerogen to petroleum. These changes
occur in hydrous environments, which probably
reduce the reactive capabilities of clays, usually
before significant hydrocarbon generation has
occurred. Nevertheless, some degree of mineral
matrix effect probably does persist under geological Figure 2 Modified Van Krevelen diagram for organic facies
conditions. A through D. (After Jones.[12])

Kerogen Type I

Kerogen type I is predominantly composed of the most hydrogen-rich organic matter preserved in the rock
record. Often the organic matter is structureless (amorphous) alginite and, when immature, fluoresces golden
yellow in ultraviolet (UV) light. A large proportion of type I kerogen can be thermally converted to petroleum
and therefore is rarely recognizable in thermally mature or postmature rocks. Sometimes in thermally immature
rocks, morphologically distinct alginite is structurally or chemically assignable to specific algal or bacterial
genera. These organic-walled microfossils have high H/C values because they formed hydrocarbons biologically.
Some examples of pure assemblages with type I kerogen properties include the following: (1) the lacustrine alga
Botryococcus braunii, which sometimes retains its diagnostic cup-and-stalk colonial morphology and/or its
unique chemical compound, botryococcane;[13] (2) Tasmanites spp., which are low-salinity, cool water, marine
algal phyto-plankton with unique physical features;[14] and (3) the Ordovician marine organic-walled colonial
microfossil Gloeocapsomorpha prisca, with its diagnostic physical appearance and unique chemical signature.[15]
Where kerogen type I is widespread, it is mapped as organic facies A. It usually forms in stratified water columns
of lakes, estuaries, and lagoons.

Kerogen type I is concentrated in condensed sections where detrital sediment transport is low and primarily
pelagic. Condensed sections occur in offshore facies of transgressive systems tracts in marine and lacustrine
settings. Although this extension of terminology from marine to lacustrine environments may be unfamiliar at
first, lacustrine rocks are formed by the same dynamic processes that form marine rocks (i.e., sediment supply,
climate, tectonics, and subsidence), although changes in lake levels often reflect local changes in runoff,
evaporation, and sediment basin filling rather than the global and relative sea level changes postulated for marine
sediments.[16]

Kerogen Type II

Kerogen type II in its pure (monomaceral) form is characterized by the relatively hydrogen-rich maceral exinite.
Examples include spores and pollen of land plants, primarily marine phytoplankton cysts (acritarchs and
dinoflagellates), and some land plant components such as leaf and stem cuticles. As with kerogen type I, the
occurrence of kerogen type II depends on high biological productivity, ow mineralic dilution, and restricted
oxygenation. The pure exinitic kerogen type II is preserved in condensed sections and represents macerals that
are slightly less hydrogen rich than kerogen type I.

Kerogen type II can also be formed from partial degradation of type I kerogen or from varying mixtures of type I
and types II, III, and IV. For example, organic matter formed in different provenances can be combined, such as
when planktonic algal material falls into sediments containing transported woody macerals (kerogen type III).
Kerogen type II is recorded in transgressive systems tracts, sometimes landward of type I kerogen deposition.

Kerogen Type III

Kerogen type III contains sufficient hydrogen to be gas generative but not enough hydrogen to be oil prone. In its
pure form, it is composed of vitrinite, a maceral formed from land plant wood. As with other intermediate
kerogen types, however, various maceral mixtures or degradational processes can contribute to kerogen type III
formation. Coal-forming environments represent several different kerogen types. Most coals form in paralic
swamps and abandoned river channels. Vail et al. (in press) find that in regions where sediment supply is low,
incised valleys contain these sediments as estuarine or coastal plain deposits.

Kerogen Type IV

Kerogen type IV is a term not universally employed by organic geochemists because it is difficult to distinguish
type IV from type III using only Rock-Eval pyrolysis. It is an inert (does not generate hydrocarbons) end-member
on the hydrocarbon generative spectrum. Kerogen type IV is composed of hydrogen-poor constituents such as
inertinite, which is detrital organic matter oxidized directly by thermal maturation including fire (charcoal) or by
biological or sedimentological recycling.

See also
◾ Rock Eval analysis using hydrogen index (HI) and oxygen index (OI)
◾ Kerogen type and quality: visual assessment
◾ Pyrolysis gas chromatography
◾ Relationships between maturity and hydrocarbon generation
◾ Kerogen type and hydrocarbon generation
◾ Kerogen type and maturity
◾ Kerogen type and transformation ratio
◾ Open- vs closed-system generation modeling
◾ Van Krevelen diagram
◾ Type I kerogen
◾ Type II kerogen
◾ Type IIS kerogen
◾ Type III kerogen
◾ Type IV kerogen

References
1. ↑ Durand, B., 1980, Sedimentary organic matter and kerogen: definition and quantitative importance of
kerogen, in B. Durand, ed., Kerogen: Techniq, p. 13–14.
2. ↑ Tissot, B. P., and D. H. Welte, 1984, Petroleum Formation and Occurrence, 2 ed.: New York, Springer-
Verlag, 699 p. The best overall reference for petroleum geochemistry.
3. ↑ Dembicki, H., 2009, Three common source rock evaluation errors made by geologists during prospect or
play appraisals (http://archives.datapages.com/data/bulletns/2009/03mar/BLTN08076/BLTN08076.HTM):
AAPG Bulletin, vol. 93, issue 3, pp. 341-356.
4. ↑ 4.0 4.1 4.2 4.3 4.4 Espitalie, J., M. Madec, and B. Tissot, 1980, Role of mineral matrix in kerogen pyrolysis:
Influence on petroleum generation and migration (http://archives.datapages.com/data/bulletns/1980-
81/data/pg/0064/0001/0050/0059.htm): American Association of Petroleum Geologists Bulletin, v. 64, p.
59-66.
5. ↑ 5.0 5.1 5.2 5.3 Horsfield, B., and A. G. Douglas, 1980, The influence of minerals on the pyrolysis of
kerogens: Geochimica et Cosmochimica Acta, v. 44, p. 1110-1131.M
6. ↑ Orr, W. L., 1983, Comments on pyrolitic hydrocarbon yields in source-rock evaluation, in M. Bjoroy et
al., eds., Advances in Organic Geochemistry 1981, p. 775-787.
7. ↑ 7.0 7.1 7.2 7.3 Dembicki, H., B. Horsfield, and T. Y. Ho, 1983, Source rock evaluation by pyrolysis-gas
chromatography (http://archives.datapages.com/data/bulletns/1982-83/data/pg/0067/0007/1050/1094.htm):
American Association of Petroleum Geologists Bulletin, v. 67, p. 1094-1103.
8. ↑ 8.0 8.1 Katz, B. J., 1983, Limitations of Rock Eval pyrolysis for typing organic matter: Organic
Geochemistry, v. 4, p. 195-199.
9. ↑ Peters, K. E., 1986, Guidelines for evaluating petroleum source rocks using programmed pyrolysis
(http://archives.datapages.com/data/bulletns/1986-87/data/pg/0070/0003/0300/0318.htm): American
Association Petroleum Geologists Bulletin, v. 70, p. 318-329.
10. ↑ Crossey, L. J., E. S. Hagan, R. C. Surdam, and P. W. Lapointe, 1986, Correlation of organic parameters
derived from elemental analysis and programmed pyrolysis of kerogen: Society of Economic
Paleontologists and Mineralogists, p. 36-45
11. ↑ Langford, F. F., and M. M. Blanc-Valleron, 1990, Interpreting Rock-Eval data using graphs of
pyrolizable hydrocarbons vs. total organic carbon (http://archives.datapages.com/data/bulletns/1990-
91/data/pg/0074/0006/0000/0799.htm): American Association Petroleum Geologists Bulletin, v. 74, p. 799-
80
12. ↑ Jones, R. W., 1987, Organic Facies, in J. Brooks and D. H. Welte, eds., Advances in Petroleum
Geochemistry, v. 2, Academic Press, London, p. 1-90.
13. ↑ Moldowan, J. M., and W. K. Seifert, 1980, First discovery of botryococcane in petroleum: Chemical
Communications, v. 34, p. 912-914.
14. ↑ Prauss, M., and W. Riegel, 1989, Evidence from phytoplankton associations for causes of black shale
formation in epicontinental seas: Neues Jahrbuch fur Geologie und Palaontologie, Monatshefte, v. 11, p.
671-685.
15. ↑ Reed, J. D., H. A. Illich, and B. Horsfield, 1986, Biochemical evolutionary significance of Ordovician
oils and their source: Organic Geochemistry, v. 10, p. 347-358.
16. ↑ Haq, B. U., J. Hardenbohl, and P. K. Vail, 1988, Mesozoic and Cenozoic chronostratigraphy and cycles
of sea-level Change, in C. K. Wilgus et al., eds., Sea-Level Changes: An Integrated Approach, SEPM
Special Publication 42, p. 71-108.
External links
◾ Original content in Datapages find literature about
Kerogen
(http://archives.datapages.com/data/specpubs/beaumont/ch06/ch06.htm)
◾ Find the book in the AAPG Store (http://store.aapg.org/detail.aspx?id=545)
Retrieved from "http://wiki.aapg.org/index.php?title=Kerogen&oldid=25013"

Categories: Critical elements of the petroleum system Evaluating source rocks Geochemistry

◾ This page was last modified on 1 August 2016, at 13:46.


◾ This page has been accessed 61,222 times.
◾ Content is available under Attribution-ShareAlike 3.0 Unported unless otherwise noted.

You might also like