You are on page 1of 87

International Journal of Coal Geology, 12 (1989) 1-87 1

Elsevier Science Publishers B.V., Amsterdam - - Printed in The Netherlands

The genesis of coal from the viewpoint of coal


petrology

MARLIES TEICHM(~LLER
Geologisches Landesamt Nordrhein- Westfalen, de Greiff Str. 195, 4150 Krefeld, F.R. Germany
(Received June 15, 1988; revised and accepted September 26, 1988)

ABSTRACT

Teichmfiller, M., 1989. The genesis of coal from the viewpoint of coal petrology. In: P.C. Lyons
and B. Alperu (Editors), Peat and Coal: Origin, Facies, and Depositional Models. Int. J. Coal
Geol., 12: 1-87.

The genesis of the microscopic constituents of coal (macerals) and of maceral associations
(microlithotypes) representing various coal facies is discussed in the light of recent advances that
have been made in the study of coals and modern peats. Peat microscopy especially promoted the
understanding of humification and biological gelification, both of which are decisive for the gen-
esis of huminites/vitrinites in brown coals and hard coals, respectively. Microbial activity during
peatification has been documented by chemists, microbiologists and petrologists and, recently, by
electron-microscopical observations in coals. This activity leads to the incorporation of lipid sub-
stances into the humic matter and, later, influences the chemical, physical and technological prop-
erties of vitrinites. Geochemical gelification transforms the huminites of brown coals into the
vitrinites of.bituminous coals and is probably caused by the onset of bitumen generation in coal.
In this process, the hydrogel brown coal is transformed to the bitumogel bituminous coal. Coal as
a source rock for oil has become an important research subject for coal petrologists and oil explo-
ration. The formation of more or less liquid bitumen and its later cracking to gaseous hydrocarbons
leads to the "coalification jumps" of the macerals, especially liptinites and vitrinites and to the
generation of secondary macerals like exsudatinite (bitumen) and micrinite (dead carbon). Bi-
tuminization in coal has been revealed by the use of fluorescence microscopy which also found the
presence of new liptinite macerals (bituminite, fluorinite and exsudatinite) and suggests a higher
proportion of algal material in coals than was thought earlier. The genesis of various types of
inertinite macerals will probably become an increasingly important subject of research, especially
with regard to "rank-inertinites" that seem to attain their high inertinitic reflectance initially
during coalification.
Besides the origin of macerals, the origin of various coal facies in terms of maceral associations
is considered in relation to the palaeogeographic depositional milieu, the peat-forming vegetation,
palaeoclimate, water and nutrient supply, acidity, marine and calcareous influence and fire, with
examples from Euramerican Carboniferous coals, Gondwana coals, and Tertiary brown coals. Veg-
etation, water and oxygen supply (Eh-conditions) are most important for the generation of coal
facies.
Vegetation has been reconstructed with the help of palaeobotany, including palynology, cutic-

The contents is shown on p. 87

0166-5162189/$03.50 © 1989 Elsevier Science Publishers B.V.


ular analysis and wood anatomy (in brown coals ) and with the study of petrified (permineralized)
peats preserved in coal balls. The microscopical investigations of botanically recognizable plant
remains isolated from coals or accompanying rocks has been especially informative. However, the
results of palynological and cuticular analyses should be evaluated with caution, considering the
varying resistance of palynomorphs and cuticles and, in the case of palynomorphs, their possible
extrapaludal origin.
Microscopical studies of modern peats that were derived from different types of vegetation
growing in different hydrological environments have been very helpful in interpreting coal facies,
especially in brown coals. The much disputed origin of brown-coal lithotypes is considered in
detail. For hard coals there seems to be agreement that coal layers rich in vitrites and liptinite-
poor clarite were deposited in wet-forest swamps, that microlayered liptinite-rich clarites are sub-
aquatic humic muds, and that inertinite-rich durites represent oxidized peats. The origin of lip-
tinite-rich durites is controversial (either subaquatic muds or oxidized peats, the latter perhaps
from raised bogs ).
A recent trend in the literature is the assumption that many if not all mineable coal seams have
been deposited in raised bogs under ombrogenous, oligotrophic conditions. Although raised bogs
with a dense arborescent vegetation are known from Borneo, it seems not probable that most
seams (which were deposited in foredeeps or steadily sinking graben zones) are of ombrogenous
origin although a succession from relatively wet and eutrophic to relatively dry and oligotrophic
conditions may be true for occasional seams. But, as in modern peats, striking changes of coal
facies are commonly caused by hydrological events, e.g., inundations as a consequence of crevasses
or of fires that burned the peat.
The origin of the inertinite-rich Gondwana coals is discussed in relation to their specific cli-
matic conditions and a comparison with sub-arctic peats is recommended.
Coals influenced by marine and calcareous conditions are characterized by high contents of
sulfur and pyrite, as well as by hydrogen-rich vitrinites with a relatively low reflectance and high
fluorescence, all due to a strong bacterial activity in peats of low acidity.
Proposals for further research are made in the conclusions.

INTRODUCTION

Coal is a rock derived mainly from plant remains that suffered peatification
and coalification. Due to a relatively good preservation and the possibility of
comparing plant remains with modern plants, it is easier to reconstruct coal-
forming vegetation in peats and brown coals than in bituminous or hard coals.
To understand the genesis of coal constituents, the application of combined
botanical and coal petrological methods has been most effective. Such studies
have been carried out in the past on hard coals, among others, by Thiessen and
White (1913), Stopes (1919), Thiessen (1920), Seyler (1928), Hickling and
Marshall (1932, 1933), Thiessen and Sprunk (1941), Teichmiiller (1952),
Hoehne (1954), Teichmiiller and Schonefeld (1955), Kosanke and Harrison
(1957) and, recently, by Phillips et al. (1985) and Winston (1986). These
authors studied the microscopic appearance of botanically identified, coalified
plant remains, occurring either in clastic rocks or in coal seams, or they traced
plant fossils from petrified peats in coal balls into the coal seam.
Another relatively small group applied "cuticular analysis" (Jurasky, 1934/
1935 ) to identify the epiderms of leaves that contributed to peat formation in
Tertiary brown coals (Weyland, 1956; Benda, 1960; Kilpper, 1960; Schneider,
1980). Many authors used palynological methods to draw conclusions from
spores and pollen occurring in coals to the peat-forming vegetation (e.g.,
Thomson, 1950/1951; Hunger, 1955; Smith, 1962, 1968; Hacquebard and Don-
aldson, 1969; Von der Brelie and Wolf, 1981a,b; Sontag and Schneider, 1982;
Bartram, 1987). Most important and effective were microscopic studies of
modern peats in relation to coal genesis (Koch, 1966, 1969b, 1970a; Cohen,
1968, 1973, 1974; Cohen and Spackman, 1977, 1980; Cohen et al., 1987; Styan
and Bustin 1983a,b). Besides botany and peat microscopy, the study of mineral
occurrence (e.g., pyrite, clay partings ) in coal seams, chemical analyses of iso-
lated coal constituents as well as the comparison of "coal facies" with the sed-
imentology of accompanying rocks helped to clarify problems of coal genesis.
From the viewpoint of coal petrology, the genesis of coal is primarily the
genesis of macerals, microlithotypes, and lithotypes. Macerals are the most
uniform microscopical constituents of coal and are comparable with minerals
in other rocks. Microlithotypes are typical maceral associations that can be
identified under the microscope, and lithotypes are layers of coal seams which
can be distinguished with the naked eye. After a short survey of the genesis of
macerals, including most important changes during peatification and coalifi-
cation, the genesis of hard-coal microlithotypes and lithotypes and the much
discussed origin of brown-coal lithotypes will be considered, thereby mainly
introducing results that are not yet included in the Textbook of Coal Petrology
(Stach et al., 1982 ).

1 ORIGINOF MACERALS

Besides the parent plant material and the initial decomposition before and
during the peat stage, the degree of coalification (=rank) is decisive for the
microscopic appearance of macerals. Morphology and reflectance under inci-
dent light are the main properties to distinguish macerals and maceral groups
under the microscope (Table 1 ). In low-rank coals, the relatively hydrogen-
rich liptinites show the lowest reflectance, the relatively oxygen-rich vitrinites
a medium reflectance, and the relatively carbon-rich inertinites the highest
reflectance. In the stage of low-volatile bituminous coals, vitrinite reflectance
is surpassed by liptinite reflectance (Fig. 1 ) and at the meta-anthracite stage
vitrinite and (former) liptinite reflectances surpass the reflectance of inertin-
ites (Alpern and Lemos de Sousa, 1970).
Whereas liptinite and inertinite macerals have the same names in brown
coals and hard coals, the precursors of vitrinites in the brown-coal stage are
the huminites, which are much more differentiated than the vitrinites (Table
2).
4

TABLE 1

Maceral groups and macerals

Group Maceral Characteristics


telinite cell walls
telo- tissue
collinite amorphous (gel or gelified
detro- detritus)
Vitrinite corpocollinite cell fillings
vitrodetrinite detritus

sporinite spores, pollen


cutinite cuticles
suberinite suberinized cell walls (cork)
fluorinite plant oils
Liptinite resinite resins, waxes, latex
alginite algae
bituminite amorphous (bacterial, algal, faunal)
chlorophyllinite chlorophyll
liptodetrinite detritus
exsudatinite secondary exudates

fusinite cell walls (charred, oxydized)


semifusinite cell walls (partly charred, oxydized)
sclerotinite fungal cell walls
Inertinite* macrinite amorphous gel (oxydized, metabolic)
inertodetrinite detritus
micrinite secondary relics ofoil generation (mainly)
*A small part of inertinite is derived from melanin-rich plant and animal material ("primary
inertinite"). A greater part does not attain its inertinitic properties until during the early coalifi-
cation process ("rank inertinite").

1.1 Genesis of huminites/vitrinites

Huminites/vitrinites originate mainly from lignin and cellulose of the plant


cell wall and partly also from tannins that impregnate the cell walls and fill
cell lumens. Proteins and lipid substances also take part in the formation of
huminites/vitrinites. Before considering the genesis of some individual ma-
cerals, the formation of huminites and vitrinites during peatification will be
described.

1.1.1 Humification and biochemical gelification


Humification is the main process during the peat stage. It is strongest at the
peat surface and is caused by a slight oxidation and by microbial activity. Cel-
lulose is more easily attacked by fungi and bacteria than lignin. Casagrande et
30

25

20

X
0O
El5

10

05

0 05 10 15 20
% Rmax vitrinite
volatile yield (wt. %)
60 50 40 30 20 12
I I, I i I I I I
70 80 85 90
carbon ( w t %)
Fig. 1. Coalification tracks of the three maceral groups. Modified after Smith and Cook (1980).
Rmax = mean maximum reflectance; wt. = weight.

al. (1985) found in peats of the Okefenokee Swamp that cellulose decreases
from 16% (of the total organic matter) at the peat surface to less than 1% at
a depth of 2 m, and that the microbial activity is high at the peat surface but
by 15 to 20 cm depth is already very low. With increasing depth, fungal and
aerobic bacterial activity is replaced by anaerobic bacterial activity. According
to Belyaev et al. (1981), quoted in Given and Miller (1985), for example, the
bacterium Clostridium sp. ferments cellulose under anaerobic conditions and
converts it to lower fatty acids. Nevertheless, small parts of both cellulose and
lignin are still present in soft brown coals (Teichmiiller and Thomson 1958)~
Hatcher et al. (1981) even found 75% lignin and 25% cellulose in a piece of
wood embedded in a Miocene brown coal. Bacterial substances including pro-
teins and fats, as well as the metabolic products of bacteria and fungi, enter
the humification process in addition to lignin, cellulose and tannins. They rep-
resent the lipid parent material of huminites (and vitrinites) (cf. Taylor and
Liu, 1987). The same is valid for algae that have become structurally wholly
TABLE 2

Correlation of the huminite macerals of brown coal with the vitrinite macerals of hard coals (after Alpern and Teichmfiller, 1971, and International
Handbook of Coal Petrography 1971, 1975)

Brown Coal Hard Coal

Maceral Maceral Maceral Maceral Type Maceral Maceral Type Maceral Maceral
Group Subgroup Variety group
Textinite A (dark)
B (bright)
A
Humotelinite Texto-Ulminite Telinite 1
B
Ulminite Telinite
A
Eu-Ulminite Telinite 2
B

Huminite Attrinite Vitrodetrinite Vitrinite


Humodetrinite
Densinite Desmocollinite
Levi- Detrogelinite
Gelinite gelinite Telogelinite Telocollinite
Collinite
Eugelinite Gelocollinite
Humocollinite Porigelinite
Corpo- Phlobaphinite
huminite Pseudo-Phlobaphinite Corpocollinite
destroyed and which commonly are relatively rich in fatty and proteinaceous
substances.
Humification is not only a biological but also a chemical process caused by
slight oxidation in the "peatigenic layer" at and near the peat surface. It re-
sults, inter alia, in the formation of humic substances such as fulvic and humic
acids, humins, and humates. Figure 2 shows a schematic representation of the
chemical reactions taking place during the formation of humic substances ac-
cording to Flaig (1965). As seen under the microscope, the appearance and
increasing intensification of a brown colour in thin sections, the disappearance
of cellulose, indicated by loss of anisotropy, and change or loss of fluorescence
in cell walls are characteristic of humification.
The recent controversy as to whether cellulose or lignin causes the fluores-
cence of cell walls in peats and soft brown coals (Stout and Bensley, 1987) has
not been resolved. Most brown-coal petrologists attribute fluorescence to cel-
lulose (among others, Jurasky, 1940; Teichmiiller, 1950; Russel, 1984; Russel
and Barton, 1984) because cellulose anisotropy is usually combined with a
striking fluorescence of the cell wall. But it is also known that cellulose-con-
taining cell walls in brown coals are usually lignified, probably because an in-
timate mixture with lignin protected the cellulose from decay. Botanists like
Treiber (1957, p. 82) in his book on "The chemistry of the plant cell wall"

Scheme of the formation of humic substances.

I ign in o
i
C6- C3- compounds
I deco m . " ~
position

I
C6- Cl - c om p ounds I polymerisates

oliphotic
Y i polymeri - ;
1
decomposition.
compounds i sates I
I

1
energy metabo-
I
i
I
I
decompo -~
:
i
I

lism of m i c r o - i sition i
\ ; organisms f t
÷ N - c o n l o i n i n g compounds derived f r o m the
\ decomposition of proteins /
V
humic substances
humic acids

I
decomposition

Fig. 2. Scheme of the chemical reactions during humification. From Flaig (1965).
reports that the native, non-lignified cell wall fluoresces blue (under UV-light
and that, after lignification, the colour can change to greenish or yellowish.
However, after solution of all other substances, the pure lignin lacks fluores-
cence. On the other hand, Frey-Wyssling ( 1959, p. 280) in his book "The plant
cell wall" states that lignin fluoresces weakly with a blue colour, whereas "cel-
lulose and chitin transmit ultraviolet light and, therefore, are not fluorescent"
(the latter statement is suspect since incident light was not used for excitation ).
Koch (1969b) distinguished in thin sections of peat five stages of cell-wall
humification, the latest stages (IV and V) representing the beginning and
completion, respectively, of so-called biochemical gelification. Stage V is char-
acterized by a dark brown colour, total loss of cell structure, and homogeniza-
tion to a pure gel with dessication cracks, all typical of telogelinites.
In a recent paper, Stout and Spackman (1987), likewise, described micro-
scopic changes of ligno-cellulosic cell walls in peats, the changes not being
influenced by increasing depth of burial. They differentiated the four stages:
(1) "alteration"; (2) "degradation"; (3) "humification"; and (4) "gelifica-
tion". Their "humification" (as stage 3) is understood as a stage of structural
decomposition. Therefore, it is not comparable with the chemical and micro~
bial process that normally is named humification. The latest stage (4) of Stout
and Spackman (1987) is characterized by a slight swelling of cell walls and the
beginning of porigelinite formation. It corresponds to the very beginning of
what normally is called "biochemical gelification".
The chemical genesis of huminitic cell walls has been studied recently by
Stout et al. (1987) using Curie-point pyrolysis-mass spectrometry and pyrol-
ysis-gas chromatography/mass spectrometry methods. The main processes are:
( 1 ) a rapid decrease of hemicellulose; (2) the introduction of fungal carbohy-
drates; (3) biochemical depolymerization of cellulose; (4) biological demeth-
ylation, demethoxylation and further defunctionalization of lignin; and (5)
"additional geochemical transformation of lignin leading to a predominantly
aromatic hydrocarbon network". The authors assume that the loss of hemi-
cellulose and cellulose during peatification promotes an early gelification
(known as "geochemical gelification") due to a disruption of the crystalline
framework and the resulting homogenization of the cell wall.

Biochemical gelification of humified substances includes partial or total loss


of cell structure, peptidization, softening or plasticity, compaction and homog-
enization. It may take place partly in the peat stage, but mainly in the stage of
soft brown coal. Recently, Russell ( 1984 ) and Russell and Barron (1984) stud-
ied this process in Australian brown coals with petrological as well as with
chemical methods. They found that during gelification of woody tissues cellu-
lose is progressively removed resulting in an increase of reflectance and a de-
crease of fluorescence intensity. Lignin reacts through a progressive loss of
methoxyl groups (see also Hatcher et al., 1981 ) and shows evidence of minor
oxidation. The atomic ratios H:C and O:C both decrease during biochemical
gelification, and aromaticity (fa) increases. Thus, gelified woods in soft brown
coals ("ulminites") are, so to speak, pre-coalified (Plate I-l). Biochemical
gelification is obviously promoted under water (Cohen et al., 1987) and, there-
fore, is especially common in subaquatic peat and brown-coal facies (Teich-
miiller, 1950). Chaffee et al. (1984) state that the lignin polymers must have
been attacked first by fungi under aerobic conditions. In a totally anaerobic
milieu, lignin remains preserved and is not gelified.

1.1.2 Macerals of the huminite/vitrinite group


A morphological distinction is made between cell tissues ( = telinites), de-
trims ( = detrinites), and gels ( = gelinites/collinites).

The humotelinites are subdivided into textinites (not gelified) and the + gel-
ified ulminites (Plate I-1 ) of brown coals. Textinite A is distinguished by a
very low reflectance which, for its part, may be caused by relics of cellulose (cf.
Plate III-6) and/or by resinous impregnations in the cell wall. According to
Jurasky (1940, p. 68) impregnations of cell walls with resin are restricted to
coniferous woods. Russel and Barron (1984) concluded that "the principal
cellulose-bearing submaceral appears to be textinite". Most textinites and tel-
inites in brown coals are derived from coniferous wood, whereas angiosper-
mous wood and the non-lignified tissues of herbaceous plants are structurally
more or less decomposed. Schneider (1984) gave an excellent survey of the
various humotelinites in Tertiary brown coals of Lusatia with many photo-
micrographs, according to their origin from wood ("xylo-textite"), bark ("per-
idermo-textite"), leaves ("phyllo-textite") and roots ("rhizo-textite"). In-
grown roots are especially well preserved in peats and coals because they were
protected from oxidation at the peat surface and often reached beneath the
peatigenic layer (cf. Plate II-1 and 7).
The successors of humotelinites after the geochemical gelification are the
telinites and telocoUinites of hard coals. Telinite is from cell walls and distin-
guished by its visibility within a cell structure, whereas telocollinite is struc-
tureless in incident light but may show the cell walls in thin sections or after
etching in incident light.
According to Raistrick and Marshall (1939, p. 182) the following cell struc-
tures could be identified in Carboniferous vitrinites on the basis of compari-
sons with botanically identified compressed plant remains overlying coal seams:
Bothrodendron bark, Lepidodendron bark, Sigillaria bark and gymnospermous
wood related to cordaites. In Mesozoic coals Cycadophyte wood was identified.

Humodetrinites consist of an intimate mixture of cell fragments and more


or less amorphous, humic colloidal particles (cf. Plate III-3 and 4) whose
amount increases with rising degree of biochemical gelification, beginning with
IU

attrinite, passing t h r o u g h densinite, a n d e n d i n g with detrogelinite as a m a c e r a l


type of the h u m o c o l l i n i t e subgroup (see T a b l e 2).
As seen u n d e r the e l e c t r o n microscope, K o c h a n d S c h e u e r m a n n (1970) de-
scribed the colloidal particles in G e r m a n b r o w n coals as flaky a n d filifbrm in
forest coals ( f o r m e d u n d e r relatively dry, aerobic a n d acid c o n d i t i o n s ), cloudy
in m a r s h peats ( f o r m e d u n d e r relatively moist, m o r e or less aerobic a n d basic-
n e u t r a l c o n d i t i o n s ), a n d as c o m p a c t gels in a n a e r o b i c organic muds. In the last
case, real gelinites are formed.
Very recently, T a y l o r a n d Liu (1987) used the e l e c t r o n m i c r o s c o p e to ob-
serve in h u m i n i t e s of A u s t r a l i a n soft b r o w n coals a n d s u b b i t u m i n o u s coals,
f i l a m e n t o u s s t r u c t u r e s of p r o b a b l e fungal or a c t i n o m y c e t o u s origin, sur-
r o u n d e d by o p e n - t e x t u r e d biodegraded zones (magnifications 5,000-38,000 f ).
T i n y r o u n d e d bodies with a high t r a n s m i t t a n c e for e l e c t r o n s are i n t e r p r e t e d
as lipid-rich bacterial r e m a i n s ( P l a t e I-2).
H u m o d e t r i n i t e is derived m a i n l y f r o m easily d e c o m p o s a b l e (because lignin-
p o o r a n d cellulose-rich) h e r b a c e o u s p l a n t s a n d f r o m a n g i o s p e r m o u s woods.
S o m e a u t h o r s regard h u m o d e t r i n i t e as an o x i d a t i o n p r o d u c t of forest peats,
also those of coniferous origin (Von der Brelie a n d W o l f 1981a), cf. pp. 48 - 50.

Due to the geochemical gelification (see pp. 13,14), h u m o d e t r i n i t e s are t r a n s -

PLATE I
(Width of photomicrographs in parentheses)
1. Biochemical gelification of a piece of wood found in a layer of clay in the Miocene Rhenish
brown coal. The gelification is complete at the right-hand side (eu-ulminite). Polished section,
oil immersion, 270×, (0.24 mm).
2. Dark filamentous structure (probably fungal) surrounded by open-textured biodegraded zone.
Lipid-rich bodies (L) are light. Subbituminous coal, Bass Strait, Australia. Electron (TEM)
photomicrograph, 25, 800×, (0.0025 mm). From Taylor and Liu (1987). Courtesy of G.H.
Taylor.
3. Humocollinite (vitrinite A) and desmocollinite (vitrinite B) in a low rank bituminous coal
from the Saar District. Polished section, oil immersion, 440×, {0.145 mm).
4. Phlobaphinite (corpohuminite) as cell fillings of (?) root tissues in a compressed ombrogenous
peat of Ireland. Polished section, oil immersion, 270 ×, (0.24 mm).
5,6. Cracks in vitrite and clarite of an Oligocene subbituminous coal from Upper Bavaria. As seen
under normal white light (5) the cracks seem to be empty, but under blue light irradiation (6)
they reveal to be filled with fluorescing exsudatinite. Polished section, oil immersion, 500 N,
(0.13 mm).
7. Degradofusinite as the result of fungal attack in a parenchymatousous tissue. For details see
Teichmtiller (1950, p. 440). Miocene brown coal of Konin, Poland. Polished section, oil im-
mersion, 200×, (0.32 mm).
8. Naturally charred forest peat. Charring below the peat surface. Note the high porosity and the
fragile structure of the peat-coke. Polished section, oil immersion, 80 ×, (0.8 mm).
9. Pyrofusinite from a birch-root, charred below the peat surface. Note the very fine structure of
the wood tissue (left) and the charred cortex (right). Polished surface, oil immersion, 80×,
(0.8 mm).
12

formed to desmocollinites of hard coals. Desmocollinite forms the vitrinitic


groundmass of clarite and trimacerite. On the basis of auto-radiographic stud-
ies of vitrinites, Alpern and Quesson (1956) established long ago that desmo-
collinites have a relatively high content of "inherent ash" and are heteroge-
neous in composition. Therefore, Alpern (1966) subdivided the collinites of
hard coals into homocollinites (telocollinites) and heterocollinites (desmo-
collinites), the latter with a lower reflectance and better coking potential. Brown
et al. (1964) named these components "vitrinite A" and "vitrinite B", respec-
tively (Plate I-3 ). Interestingly, the least reflecting and most strongly fluoresc-
ing types of desmocollinite occur in clarites and trimacerites with a high lip-
tinite content (Hutton and Cook, 1980; Kalkreuth, 1982; CorrSa da Silva et al.
1986). The very fine intercalations of material with a relatively high trans-
mittance for electrons (Plate I-2), detected by Liu and Taylor (1987) and
Taylor and Liu (1987) may cause the relatively low reflectance and high flu-
orescence of desmocollinites, together with absorbed bitumen that was gener-
ated during the bituminization process (see below) from these submicroscop-
ical intercalations and from liptinites embedded in the desmocollinite
groundmass. Interestingly, in the stage of low-volatile bituminous coal, Taylor
(1966) no longer observed this electron microscopical inhomogeneity of des-
mocollinite, probably as a consequence of bituminization (p. 15).
In contrast to desmocollinite, which, on account of its relatively high amount
in lipoid constituents, often belongs to the perhydrous (H-rich) vitrinites, the
pseudovitrinite of Benedict et al. (1968) is subhydrous and is distinguished by
a relatively strong reflectance and poor coking potential. The origin of pseu-
dovitrinite is not yet understood. Most authors regard pseudovitrinite as an
early oxidation product. This opinion is not supported by the low-temperature
oxidation experiments on a medium-volatile bituminous coal carried out re-
cently by Kaegi (1985). The present author assumes that pseudovitrinites rep-
resent formerly asphaltene-rich vitrinites that reached the stage of overma-
turity. Therefore, pseudovitrinites are most abundant in low-volatile
bituminous coals (R. Thompson, Bethlehem Steel Co., pers. commun. ).

The gelinites of brown coals are either biochemically totally gelified plant
tissues (telogelinites, as shown in Plate I-1 ) or gelified humic detritus (detro-
gelinite), or pure humic gels derived from colloidal solutions that entered for-
mer voids (eugelinite). Plate III-6 shows porigelinite with typical dessication
cracks between the cell walls ofMarcoduria inopinata. Biochemical gelification
seems to be promoted under water. It is characteristic of anaerobic subaquatic
facies types, such as humic gyttjae (Teichmfiller, 1950). However, oxidation
of peat and brown coal in the presence of water also leads to early gelification.
Calcium-rich coals are especially rich in gelinites, often precipitated as Ca-
humates (dopplerite). The gelinites of the brown-coal stage correspond to the
collinites of the hard-coal stage.
13

Corpohuminites are fillings of cell lumens and may be primary secretions of


the living plant or deposits formed shortly after the death of certain cells, es-
pecially in barks. Chemically these corpohuminites are oxidation or conden-
sation products derived from tannins. Soos (1963, 1966), who first studied
these cell secretions in brown coals, named them "phlobaphenites" (now spelled
phlobaphinites). Cohen (1968) and Cohen and Spackman ( 1980 ) found these
cell fillings in roots of the mangrove tree Rhizophora and of the sedge Mariscus
jamaicensis, respectively, both in peats of the Everglades in southern Florida.
Plate I-4 shows phlobaphinites in an artificially compressed Sphagnum-peat
in which the relatively strong reflectance of these cell secretions is remarkable.
Other corpohuminites are products of the biochemical gelification which en-
tered the open cell lumens.
Although corpohuminites do not take part in the process of geochemical
gelification, they are named corpocollinites in the hard-coal stage (in analogy
to other huminites/collinites). Corpohuminites/corpocollinites are very re-
sistant and, therefore, are often preserved in coal balls as non-petrified coaly
material (whereas most cell walls have been petrified by carbonates or sili-
cates). In clastic rocks, these former cell fillings are often the only relics of
former tissues, and, therefore, can serve for reflectance measurements when
other vitrinites are lacking. In oxidizing reagents like Schulze's solution, cor-
pocollinites ("resin rodlets") of Carboniferous coals are not attacked (Ko-
sanke, 1955 ). In former times, corpocollinites have been commonly mistaken
for resinite cell fillings ("melanoresinite", "resin rodlets" ) but Hoehne, as early
as 1954, recognized the humic nature of this constituent in hard coals and
named it "Sekretvitrit". He also recognized the abundance of these cell secre-
tions in Carboniferous coals and claystones where the largest rodlet-like cor-
pocollinites were found around and within coalified stalks of the pteridosperm
Alethopteris and other medullosan seed ferns. According to Kosanke (1955)
the plant genus MeduUosa produced tremendous quantities of "rodlets". The
"secretinites" of Lyons et al. (1986) do not belong to the corpocollinites. They
are obviously inertinitic cell secretions from vascular plants, mainly medullo-
san seed ferns; they belong to the "primary inertinites" (see pp.23,24 ).

1.1.3 Coalification of huminites/vitrinites


Geochemical gelification, in contrast to biochemical gelification which is fa-
cies dependent, is a coalification process by which huminites are transformed
into vitrinites. This process of"vitrinitization" takes place between the brown-
coal and the hard-coal stages*. It changes the petrographic appearance of coal

*Theterm "browncoal" includeshere the "soft browncoal" (Weichbraunkohle)and "dull brown


coal" (Mattbraunkohle) of the German classification.It correspondsto the term "lignite" of the
U.S.A.classification(AmericanSocietyfor Testingand Materials,ASTM) (seeTextbookof Coal
Petrology,Stach et al., 1982,p. 45, Table 4).
14

more than any other process during coalification: the colour of coal changes
from brown to black, its lustre from dull to bright, its hardness from soft to
hard and its microscopic picture of the cellulose-lignin-derived material from
a loosely packed agglomerate of many different huminite macerals to the densely
packed, compressed and homogenized vitrinite macerals. The causes of geo-
chemical gelification, clearly, must be connected with the rising temperature
and load pressure because, in contrast to the biochemical gelification, geo-
chemical gelification affects all huminites at a certain rank stage. Cook and
Struckmeyer (1986) suggested that vitrinitization is the result mainly of load
pressure which also causes water release. In fact, decrease of porosity and - -
as a consequence of it - - dewatering are the most important physical processes
during coalification of brown coals ("Braunkohle") and lignites, respectively.
But, as shown in Figure 3, pressure and the resulting loss of porosity do not
cause gelification, i.e., homogenization of the humic particles. Because geo-
chemical gelification ( = vitrinitization) coincides with the beginning of the
"oil window", i.e., with the formation of more or less liquid bitumen in coal,
the author suggests that this newly formed bitumen serves as the fluid com-
ponent in a new colloidal system in which the solid part is represented by the
stable aromatic groups of huminites/vitrinites. Thus, the "hydrogel" brown

Fig. 3. Effect of compression on soft brown coal as seen under the microscope. Polished sections,
oil immersion, 250 X and 125 X respectively, (width of pictures 0.17 mm and 0.32 mm, respectively ).
15

Fig. 4. "Pitch coal" from seam No. 7, Harmattan Mine, Illinois (0.42% Rr vitrinite). Note the
glass-like appearance with conchoidal fracture (width of picture 10.5 cm).

coal/lignite changes into the "bitumogel" hard coal (="bituminous coal"!)


during the stage of subbituminous coal. Coals that are especially "oil-prone"
as, for example, the marine influenced and calcareous Oligocene subbitumi-
nous coals of the northern foreland of the Alps and some low-rank Carboni-
ferous coals of Illinois develop a pitch-like appearance with a very high lustre,
and a conchoidal fracture with knife-sharp edges, as is shown in Figure 4.

B ituminization (generation of petroleum-like substances ) takes place as part


of the coalification process mainly in the stages between subbituminous and
high-volatile bituminous coal. As a result, plastic and fluid substances, partly
hydrocarbons, are expelled from the lipoid parent materials of vitrinites and
from liptinites. Microscopic indications of oil generation in coal are the occur-
rence of exsudatinite {Plate I-5,6), of "oil expulsions" from fissures in vitrin-
ites, of"smear films" on the polished surface of vitrinites, and the fluorescence
of certain vitrinites. Coalification "jumps" of liptinites (see pp. 22,23) and
vitrinites which are indicated by striking changes in fluorescence and reflec-
tance properties correspond to the beginning, the maximum and the end of oil
generation in coals. Micrinite is a relic of oil generation due to a dispropor-
tional reaction (see p. 28).
Since these first observations by the author (Teichmiiller, 1974a,b,c) "Pe-
troleum Formation in Coal" and "Coal as a Source Rock" have become modern
topics for coal petrologists (Teichmfiller, 1982b,1984,1987; Cook and Struck-
meyer, 1986, Murchison, 1987) and organic geochemists, respectively, espe-
cially Durand (Durand and Paratte, 1983; Teichmfiller and Durand, 1983).
1G

Cook and Struckmeyer (1986) assumed that the earliest oil generating in coal
stems from fatty acids, esters and alcohols which are preferentially associated
with huminites/vitrinites. This is in accordance with earlier results of Otten-
jann (1985,1988) who concluded that vitrinites generate oil earlier than spor-
inites on the basis of fluorescence alteration. In fact, Hirsch (1954) has de-
scribed the "liquid structure" of bituminous coals on the basis of X-ray studies
already more than 30 years ago, and, recently, Rouzaud (1984) was able to
demonstrate under the electron microscope small oily droplets between the
aromatic units of vitrinites in the bituminization range of coalification. Bitu-
minization has also been revealed by a striking secondary fluorescence of vi-
trinites beginning at about 0.5% random reflectance (Rr) and reaching its
maximum at about 1.1.% Rr (Teichmiiller, 1982a,b, 1984; Cook and Struck-
meyer, 1986). Combined microscopical and geochemical studies showed that
this is the range in which the yields of extract from humic coals increase con-
siderably, as do the amounts of n-alkanes and aromatic hydrocarbons in the
extracts (Radke et al., 1980 ). Bituminization causes the softening and agglom-
eration of vitrinites during carbonization and is the reason why bituminous
coals can form coke. These relationships, argued since the early seventies
(Teichmiiller 1974a,b,c), have been ascertained recently by the systematic
studies of Ottenjann et al. (1982) in Germany and of L i n e t al. (1986) in the
U.S.A. Bituminization also explains why low-rank bituminous coals (0.5-1.3%
Rr) are suitable for hydrogenation.
The liquid structure of bituminous coals (Hirsch, 1954) facilitates the ar-
rangement of the aromatic lamellae of vitrinites parallel to the bedding plane
and causes a higher anisotropy. It is not yet well known that the degree of
anisotropy at given temperature/pressure conditions depends on the hydrogen
content of vitrinites (and other macerals) (Teichmiiller, 1987, p. 143).
It is well known that the reflectance of huminites/vitrinites rises more or
less steadily during coalification (see Fig. 1 ). The degree of reflectance in-
crease is different in the different rank stages (Teichmtiller and Teichmiiller,
1982, pp. 41-51; Teichmfiller, 1982a, pp. 242-245) and is caused by increasing
aromatization of the vitrinite "molecule" which, roughly speaking, consists of
an aromatic nucleus surrounded by aliphatic groups (Van Krevelen, 1961 ).

1.2 Genesis of liptinites

Liptinites are derived from hydrogen-rich plant organs as well as from algal
and bacterial substances and decomposition products. Plant lipids, proteins,
cellulose and other carbohydrates are characteristic source materials of liptin-
ites. Exsudatinite is a secondary maceral, formed during the coalification pro-
cess at the beginning of bituminization.
17

1.2.1 Macerals of the liptinite group


Sporinite is derived from the outer cell wall of spores and pollen. It is com-
posed of the chemical substance "sporopollenin", a highly polymerized, cross-
linked material consisting of carotenoids and their esters. The botanical affin-
ities of sporinite are commonly studied by palynologists working with the "spo-
romorphae" or "palynomorphs" isolated from coal by oxidation and solution
of the humic matter. Only in few cases is it possible to recognize and identify
certain genera of sporomorphae in thin sections or polished sections of coal.
Thiessen (1947) described in thin sections the characteristically shaped "splint
spore" which is identical with microspores of the genus Densosporites belong-
ing to the Lycopsidae. This spore can also be identified in polished sections.
According to Smith (1962) it dominates the assemblages from certain Carbon-
iferous durites of English coals which he attributed to an ombrogenous type of
peat. Many attempts have been made to reconstruct the peat-forming vegeta-
tion on the basis of palynological studies of coals. However, it must be kept in
mind that palynological results do not allow precise conclusions. There is no
question that anemophilous sporomorphs are always overpresented and many
pollen exines, e.g., from Cyperaceae, are easily decomposed in peats and there-
fore are absent or underrepresented in coals (Cohen, 1968) (see also Fig. 14 of
this paper). In calcareous environments and under relatively dry conditions
spores and pollen are easily attacked by microbes, whereas they remain intact
under calcareous moist conditions in anaerobic environments.

Cutinite is derived from cuticles and cuticular layers which occur at the sur-
face of leaves, twigs and other areal parts of plants as a protection from des-
sication (Plate III-1 ). The chemical substance is called cutin and is composed
of fatty acids and waxes. In 1934/1935 the palaeobotanist and coal petrologist
Jurasky introduced "cuticular analysis" in order to identify plants whose cu-
ticles are embedded in German brown coals. This method relies, as does paly-
nology, on the great resistance of cuticles against oxidizing reagents. Cuticular
analysis has contributed much to the understanding of the genesis of brown
coal lithotypes (see pp. 50,52 ). It is, however, rarely, that the botanical origin
of cutinite can be identified directly in coal sections. The only case known to
the author is the cutinite from needles of the conifer Abietites linkii which were
first identified from the needles enriched in a Wealden claystone. This cutinite
is distinguished by a great thickness and well preserved stomata; it became a
coal petrological fossil of German Wealden coals (Teichmfiller, 1982a, Fig.
77a,b). The "leaf coal" or "paper coal" of the Pottsville Formation (basal
Pennsylvanian) in Indiana (U.S.A.) contains cutinite of Sphenopteris brad-
fordii (Guennel and Neavel, 1959 ) and of the pteridosperm Karinopteris (Nel-
son et al., 1985). The "leaf coal" of Mississippian age in the Moscow Basin
contains cutinite derived from stems of Lepidodendron and Bothrodendron
(Guennel, 1960). It is not known whether these cutinites can be identified only
18

by coal p e t r o l o g i c a l m e t h o d s . P l a t e II-5 shows highly reflecting m e t a - c u t i n i t e


in t h e a n t h r a c i t e stage of coalification.
In c o n t r a s t to sporinite, c u t i n i t e a n d alginite, t h e b o t a n i c a l origin of suber-
inite, resinite a n d fluorinite is k n o w n o n l y in a v e r y g e n e r a l way. Suberinite is
derived f r o m s u b e r i n layers of corkified cell walls, m a i n l y in tree b a r k s . T h e
c o r k l a y e r at t h e surface of t h e c o r k o a k Quercus suber c o n s i s t s of m o r e t h a n
40% suberin. S u b e r i n is a p o l y m e r c o n t a i n i n g f a t t y acids a n d glycerin e s t e r s
( T r e i b e r , 1957). It is f o r m e d n o t o n l y in b a r k s of t r e e s b u t also at t h e surface
of r o o t s a n d fruits, a c t i n g as a p r o t e c t i o n a g a i n s t dessication. A v e r y c o m m o n
o c c u r r e n c e of s u b e r i n i t e is in T e r t i a r y b r o w n coals in w h i c h l o w - r e f l e c t i n g b u t
fluorescing t h i n cell walls of s u b e r i n i t e s u r r o u n d higher reflecting, m o s t l y t a b -
u l a r - f o r m e d cell fillings of p h l o b a p h i n i t e . S u b e r i n i t e is r a r e in Mesozoic coals
a n d is u n k n o w n f r o m C a r b o n i f e r o u s coals.

Alginite, as it is d e f i n e d in the I n t e r n a t i o n a l H a n d b o o k of Coal P e t r o l o g y


(1971), is t h e c h a r a c t e r i s t i c m a c e r a l of b o g h e a d coals t h a t r e p r e s e n t f o r m e r
organic m u d s . (cf. P l a t e I I - 6 ) . T h i s classical atginite is d e r i v e d f r o m t h e fossil
algae Pila a n d Reinschia w h i c h b o t h b e l o n g to t h e Botryococcus-type of algae.

PLATE II
(Width of photomicrographs in parentheses )
1. Petrified peat in a coal ball from the Katharina seam, Ruhr District. Cross section of a stem of
Lyginopteris oldhamia with opaque, coaly cell groups in the pith. Note the ingrown stigmarian
appendices (rootlets = R) between the inner and outer cortex. Polished thin section, transmit-
ted light, 2.5× (2.6 cm).
2. One of the dark coaly bodies from the pith of Lyginopteris oldhamia in 1, under incident light
and at a much higher magnification. Compared with vitrinitic inclusions in the coal ball, the
reflectance of these cell groups is inertinitic. Example of primary fusinite. Polished thin section,
incident light, oil immersion, 110 X (0.6 mm).
3,4. Thin rims of meta-exsudatinite along cell walls of semifusinite in a high-volatile bituminous
coal from the Saar District, Germany. Vitrinite reflectance (Rr) around 1%. 3 is taken with
one polarizer, 4 is taken under crossed polarizers. Polished section, oil immersion, 270 X (0.24
mm). Photomicrographs courtesy of the late Dr. K. Hoehne.
5. Meta-cutinite (white bands of "secondary inertinite") in a Carboniferous anthracite from Ib-
benbtiren, Germany. The reflectance of the formerly dark cutinite has surpassed the vitrinite
reflectance (which is 2.8% Rmax). Polished section, oil immersion, 500X (0.13 mm).
6. Carboniferous boghead coal ( Ruhr District) in the stage of high- to medium-volatile bituminous
coal (vitrinite reflectance is 1.12% Rr in the accompanying humic coal). The concentration of
micrinite represents the solid relic of the former bituminite groundmass. The Botryococcus
alginite (grey lenses) attained a relatively high reflectance. Polished section, oil immersion,
500× {0.13 mm).
7. Dolomitized root peat in a coal ball of the Katharina seam, Lower Rhine District, Germany.
The appendices of stigmarians are nearly uncompressed. Thin section, transmitted light, 11 X
(5.8 mm).
8. Silicified calamitean peat (showing compression and microbedding) from a layer below the top
of HauptflSz seam, near Essen/Ruhr. The Calamostachyscone in the center contains sporangia
filled with spores (tiny black dots). Thin section, transmitted light, 5× (13 mm).
20

Although we must assume that many other types of algae contributed to the
formation of coal, it is surprising that up to the present time the green alga
Botryococcus braunnii, which is unusual in that it secretes fat, is the only alga
identified in coal seams. According to Liu and Taylor (1987), this alga (which
is still living at the present) is distinguished by a very high content of hydro-
carbons (up to 76% in the dry substance) in its brown resting stage which may
account for its good preservation in coals and may explain its strong yellow
fluorescence. Wolf and Wolff-Fischer ( 1984 ), Corr~a da Silva et al. (1986) and
Liu and Taylor ( 1987 ) drew attention to strongly yellow fluorescing liptinites
occurring in Carboniferous coals of the Saar Basin and in Gondwana coals of
Brazil and Australia, respectively. They call these constituents "alginite" be-
cause of their Botryococcus-like fluorescence. Part of these "alginites" resem-
ble the "lamalginite" that Hutton et al. (1980) described from oil shales. They
are thin, anastomosing filaments between inertodetrinites (Liu and Taylor,
1987 ), or they show a net-like structure, or pore canals running perpendicular
to the surface of the cell wall (Corr~a da Silva et al., 1986). But these constit-
uents have not yet been identified as algae by palaeobotanists, as is the case
for Botryococcus and the many marine algae that are known from oil shales
(M~idler, 1963; Hutton et al., 1980). On the other hand, we must assume that
a high amount of algal material participated in peat formation, even in forest
swamps and reed marshes where, at the present time, thick algal mats may
form (Plate IV-3). Obviously most algae disintegrate morphologically at the
peat surface. As a consequence of high contents of cellulose and proteins, they
are easily attacked by bacteria. The product of the microbial degradation prob-
ably goes into the amorphous liptinite maceral bituminite as well as into the
aliphatic molecular groups of vitrinites, mainly desmocollinites. According to
Liu and Taylor ( 1987 ), experiments of Verma and Martin (1976) showed that
algal decomposition products may be stabilized through complexing with humic-
acid-type phenolic polymers.

The new liptinite maceral bituminite (Teichmilller, 1974a,c) is amorphous


(i.e., without a definite shape) and occurs as finely dispersed lenses, streaks
and sometimes as a groundmass for other liptinite macerals. In brown coals it
is highly concentrated in the pale lithotypes (see pp. 56,57 ) where it forms the
groundmass (Plate III-8). In the hard-coal stage it occurs mainly in sapropelic
coals and in liptinite-rich clarites and durites. Bituminite is abundant in many
oil shales where it is known to be highly oil-prone. From these occurrences of
bituminite and from its chemical properties (hydrogen-rich, oil-prone, high
yields of extracts and of tar) it is concluded that bituminite represents a de-
compositional product of algae, animal plankton and bacteria. In this context,
it is of interest that Casagrande (1971) found that microorganisms contribute
at least 5-10% of the organic matter of peat. Masran and Pocock (1981), who
subjected tissues of specific terrestrial and marine plants to physical, chemical
21

and biological breakdown under laboratory conditions, obtained "amorphous


matter" as a product of bacterial breakdown from both terrestrial and marine
plants. In a reducing environment, under the influence of anaerobic bacteria,
the amorphous product tended to be lighter in transmitted light (i.e., darker
under incident light) than in a more oxidizing environment. In the first case,
the product obviously was bituminite; in the second case, humodetrinite/col-
linite formed. Bituminite disappears after oil generation, leaving behind solid
relics of secondary micrinite (Teichmiiller, 1974a,c; Cook and Struckmeyer,
1986) (see Plate II-6).

The maceral resinite is derived not only from resins but also from balsams,
latexes, fats and waxes. Chemically one should distinguish between terpene
resinite (from resins, balsams, copals, latexes and essential oils) and lipid re-
sinites (from fats and waxes). Terpenes are relatively stable condensation
products of isoprene molecules (C6Hs), whereas lipids of fats and waxes are
extractable mixtures of fatty acids with glycerin esters (fats) and of fatty acids
with higher alcohols (waxes), respectively. Isolated resinites from coals have
been analyzed by, among others, Murchison (1966), Mukhopadhyay and
Gormly (1984), and Lyons et al. (1984).
Botanically, resinites are secretions of cell walls, filling cell lumens and ca-
nals. Many conifer stems, when wounded, sweat out resin drops and lumps
which may form resinite in coals. Isolated resinite bodies are relics of structural
degradation. Surface corrosion and/or higher reflecting oxidation rims at the
surface indicate deposition under aerobic conditions, either at the peat surface
or under oxygenated water. Lyons et al. (1984), who described large resin rod-
lets (1-6 mm long, 0.4-1.7 mm broad) occurring as phytoclasts and in a Cre-
taceous subbituminous A coal, gave many references of fossil resins that were
found in coals and accompanying rocks.
Terpene resinites display highly variable optical and physical properties (re-
flectance, fluorescence, hardness). Recently, Teerman et al. (1987) and Crell-
ing et al. (1982) distinguished, on the basis of fluorescence colours, 4 to 5
different resinite cell fillings in individual coals of Palaeocene and Cretaceous
age, respectively. Although resinite cell fillings are known from sigillarian bark
and cordaitean wood of Carboniferous coals, Tertiary coals are known to be
especially rich in resinites, due to be abundance of conifers in Tertiary flora.
But it is well known that in the tropics many angiosperms are also rich in
resins, latexes, oils and fats, substances which may be the source material for
resinite. The Eocene brown coals of Germany, which were deposited in a trop-
ical climate, contain latex ducts referred to by the miners as "monkey hairs",
and the pale lithotypes (see p. 56) of these coals are enriched in lipid resinite.
In contrast to the resins, the waxes (lipid resinites) are deposited on the
surface of leaves and fruits as fine grains and rodlets or as crusts, particularly
in the tropics. It is difficult to recognize these fine lipid resinites in coals under
'2 '2

the microscope. Normally, they are included with the maceral liptodetrinite
( see below).
Resinites tend to be converted to bitumen (partly exsudatinite i during the
early stages of coalification.

Although probably originating from essential oils, fluorinite has been sepa-
rated from the resinite macerats because of its striking optical properties. His-
torically it was detected relatively late using the fluorescence method (Teich-
mtiller, 1974a,c ). Under short-wavelength light it shows a very striking bright-
yellow fluorescence, whereas in normal white light it cannot be distinguished
from voids or clay inclusions in coals. A typical occurrence is in the small cells
of phyllovitrinite, which, in turn, is surrounded by cutinite. Thus, at least part
of fluorinite seems to be derived from lipid cell inclusions in certain leaves.
Recently, Cook and Struckmeyer (1986) and Hageman (1987) pointed to the
possibility that fluorinite may represent secondary high-pour oil that was gen-
erated in the coal from certain cell fillings. Hageman observed a negative flu-
orescence alteration, a behaviour which normally indicates maturity, i.e., gen-
eration of oil (Teichmtiller and Ottenjann, 1977).

Liptodetrinite is a maceral that embraces fragments and fine remains from


degradation of other liptinite macerals. It may also represent small algae which
are not identifiable and/or wax grains. Liptodetrinites are abundant in sub-
aquatic coals such as sapropelic coals and certain clarites, durites and
trimacerites.

Exsudatinite, just as bituminite and fluorinite, was first detected in coals by


fluorescence studies (Teichmfiller, 1974a,b,c). It is a secondary maceral, gen-
erated during coalification at the beginning of bituminization (pp. 15,16 ) from
liptinites and perhydrous vitrinites. It fills voids like cracks (Plate I-5,6 ), bed-
ding plane joints and empty cell lumens of fusinite and sclerotinite and is com-
parable with the "migrabitumen" of Jacob (1985). It seems that the formation
of cracks is connected with the pressure that develops when bitumen is formed
in vitrinites. The chemical composition of exsudatinite is probably
asphaltenous.

1.2.2 Coalification of liptinites


In peats and low-rank coals up to the stage of subbituminous coal, liptinites
are relatively stable in their optical properties, i.e., their low reflectance and
high fluorescence. Then, an increase of reflectance, a decrease of fluorescence
intensity and a shift of fluorescence colour from lower wavelengths (green,
yellow ) to higher wavelengths (red) begins. These changes pass certain "coal-
ification jumps" which are different for the different liptinites. For sporinites,
e.g., a first "jump" occurs at a vitrinite reflectance of 0.5% Rr when petroleum-
23

like substances begin to form, a second "jump" (0.8-1.0% Rr) coincides with
the maximum of oil generation and a third "jump" (1.3% Rr) corresponds to
the "death line" of oil generation when sporinites attain the same reflectance
as vitrinite (Fig. 1) and fluorescence is more or less lost. The microscopical
changes coincide with yield and composition of extracts from liptinite-rich
coals (Radke et al., 1980). After oil generation (bituminization) some liptin-
ites "disappear" (fluorinite) or leave behind micrinite as a solid relic (from
resinite, bituminite) (see Plate II-6). Others shrink fairly considerably and
finally attain a higher reflectance than vitrinite (sporinite, cutinite ) (see Plate
II-5). The reflectance of exsudatinite surpasses that of vitrinite in the rela-
tively early stage of coking coals. Many meta-exsudatinites are distinguished
by a high degree of anisotropy (cf. Plate II-4). It is obvious that the most
striking changes of liptinite reflectance and fluorescence occur in the "oil win-
dow", i.e., when oil is generated and, later, when oily products are cracked to
gaseous hydrocarbons.

1.3 Genesis of inertinites

The properties characteristic of inertinites, such as high reflectance, little


or no fluorescence, high carbon and low hydrogen contents and strong aro-
matization are due to different causes: charring, mouldering, fungal attack,
biochemical gelification and oxidation of the same plant tissues which, when
not subject to these conditions, are humified and eventually form vitrinite.
This kind of inertinitization (which is often a rather "semi-inertinitization" )
takes place or is in progress before deposition, or it occurs in the early peat
stage near the surface. Some inertinites owe their specific optical and chemical
properties to dark coloured (i.e., relatively highly reflecting), C-rich compo-
nents of certain plants which are called melanins ("primary inertinites"). Many
inertinites, however, do not seem to attain these properties until the coalifi-
cation process ("secondary inertinites"). It is, indeed, striking that peats and
brown coals usually contain much less inertinite than hard coals.
The term "inertinite" refers to its behaviour in coals of coking rank in which
it was thought that much of the inertinite did not soften during carbonization.
However, following the early work on this subject by coal petrologists, the re-
cent studies of Diessel (1983,1985) and Brown et al. (1985) have shown that
a considerable part of the inertinite in Gondwana coals takes part in the coking
process, and that this part is distinguished by its fluorescence under green-
light irradiation.

1.3.1 Macerals of the inertinite group


Fusinites and semifusinites were partly formed by charring of plant material
or of peat, due to fires in coal swamps (see p. 69). These types are called "pyro-
fusinite" and "pyro-semifusinite", respectively, and represent the main types
:!4

not only of' fusinites but of' all inertinites in peats and brown coal {Plate I-8
and 9). Pyrofusinites are preserved in coal balls. In Illinois they contribute
2.5-7%, sometimes as much as 10% of the peat biomass and include partially
fusinized stigmarian roots {Phillips et al., 1985). The latter indicate charring
of the peat.
Permian and Carboniferous hard coals are rich in "degradofusinite" and,
especially in "degrado-semifusinite" which, in contrast to the pyrofusinites,
are distinguished by poorly preserved cell structures and are thought to be a
product of mouldering, oxidation and dehydration. It is not well known that
fungal attack of wood (e.g., by the dry-rot species Merulius lacrymans) also
alters the unused part of wood into highly reflecting humic substances {Teich-
mtiller, 1950, 1982a). Plate I-7 shows that fungal hyphae (appearing as tiny
circular cross sections) attacked a parenchymatous tissue and converted part
of its cells into highly reflecting degradofusinite. Non-attacked cells {Plate I-
7, right border) are low reflecting (humotelinite). Koch (1969b) noticed in
German peats that inertinitic tissues are often associated with fungal remains,
so he assumed a genetic connection. Recently, Cohen and Spackman (1980)
and Cohen et al. ( 1987 ) described the darkening of cell walls (in thin sections )
and an increase of reflectance around sites of fungal attack in subtropical peats
of the U.S.A. Styan and Bustin (1983a) found fungal sclerotinite together with
degrado- and pyro-fusinite in cool-climate peats of the Frazer River delta, Brit-
ish Columbia.
Pyro- and degrado-fusinites (and -semifusinites) generally are regarded as
indications of a relatively dry environment of deposition. Phillips et al. ( 1985 )
found in coal balls a rise of fusinite abundance during drier periods of the U.S.
Pennsylvanian. Likewise, Harvey and Dillon ( 1985 ) explained the increase of
inertinites in Illinois coals which were deposited in Missourian-Virgilian (Up-
per Pennsylvanian) times as compared with coals deposited in Desmoinesian
and older {Middle Pennsylvanian) times as due to a drier climate during the
Late Pennsylvanian.

"Primary [usinites" and semifusinites obviously derived from melanin-rich


plant material (Teichmiiller, 1982a, p. 277 ff. ), as it occurs, e.g., in the recent
horsetail (Equisetum arvense) (Koch, 1966). The Carboniferous seed fern Ly-
ginopteris oldhamia whose sterns are abundant in European coal balls of the
Westphalian A, contains strikingly opaque, highly reflecting clusters of scler-
enchyma cells in the pith which, according to Taylor and Millay (1981), are
one of the most diagnostic features of the Lyginopteris oldhamia stem (Plate
II-1,2 ). The high reflectance must be primary because surrounding coaly cell
material of the same stem is vitrinitic. Phillips and Peppers (1984) reported
that seed ferns and sphenopsids consistently contribute more fusain to Illinois
coals than other plants. This too may be explainable by primary causes. Phy-
tomelanines are very resistant, black pigments deposited in cell membranes of
25

plants. Chemically, they are distinguished by a high carbon content (as high
as the C-content of subbituminous coals! ) and a low hydrogen content.

"Rank fusinites" attain their high reflectance only during early coalification
(see Fig. 1 and p. 29). Possibly they stem from cell walls that were initially
impregnated with certain bituminous substances (like resins) or which still
contained cellulose in the brown-coal stage (Teichm~iller, 1982a, p. 275 ft.).
Their formation implies the generation of fluid or gaseous coalification prod-
ucts at very low rank stages and may explain the occurrence of some important
deposits of early petroleum and gas, for example, in the Canadian arctic
(Snowdon and Powell, 1982). Recently, Cook and Struckmeyer (1986) drew
attention to a possible very early bituminization in coals. From this point of
view, the Figures 3 and 4 of Plate II may be of interest because they show the
co-occurrence of semifusinite and highly anisotropic meta-exsudatinite. The
very close association of thin, highly reflecting and anisotropic rims around
fusinite cell walls has been observed by the late Dr. K Hoehne in Saar coals.
The rims may be exudates of cellulose-rich or resinous cell walls in the stage
of over-maturity. It is well known that both cellulose and resins generate bi-
tumen relatively early. An especially strong anisotropy of former bitumen at
higher-rank stages is also known, and, in this case, has been probably caused
by local magmatic heat (a sill is present in the substrata).

The inertinite maceral macrinite is a more or less highly reflecting, amor-


phous, gel-like constituent, the origin of which is still little understood. Very
intensive oxidation and dessication of plant and peat material, including, per-
haps, metabolic products of fungi and bacteria may lead to the formation of
macrinite which, however, is seldom found in peats and brown coals. A com-
pletely different hypothesis is the generation of macrinite from early gelified
material which, according to recent studies of Cohen et al. (1987), is distin-
guished by a relatively high, though still huminitic reflectance in the peat stage
and which forms preferentially in the subaquatic milieu. Like secondary fusin-
ite, most macrinites seem to attain their inertinitic reflectance during coalifi-
cation in the hard-coal stage. Part of macrinite may be derived from charred
peat (see p. 69 and Plate I-8). According to Lyons et al. (1986) angular or
irregularly broken parts of their maceral "secretinite" (i.e. highly reflecting
cell secretions of certain Carboniferous plants) may appear and, thus, be con-
fused with macrinite.

Genuine sclerotinite represents fungal mycelia that contain black melanins


already in the living stage (see primary fusinite ). Black fungal spores are known,
for example, from recent smut and rust fungi ( Ustilaginales, Puccinia) which
attack grasses, sedges and cereals. Black fungal sclerotiae (which can survive
26

unfavourable conditions, like droughts) are known from the fungi Claviceps
purpurea (which attacks Gramineae) and Rhytisma acerium, living on maple
leaves. Black fungal plectenchyme occurs as mycorrhiza, i.e., a symbiosis of
tungi and roots of higher plants. Thus, only certain, melanin-rich fungi are
predestined for the formation of sclerotinite. The old opinion that chitin (the
main substance of fungi) causes the high reflectance of sclerotinite cannot be
substantiated. On the contrary, the chitin of scorpion cuticles found in Car-
boniferous bituminous coals of Yorkshire, England (Bartram et al., 1987), and
the chitin of presumable zoo-epiderms found in Canadian subbituminous coal
(Goodarzi, 1984) is translucent and fluorescent. According to Frey-Wyssling
(1959) chitin resembles cellulose in its optical properties. Thus, most fungal
material deposited in peats will not form sclerotinite.
Today it is generally assumed that the greatest part of the so called "scler-
otinite" in Carboniferous and Permian coals stems from cell secretions (of
tannins and/or resins) which suffered oxidation or carbonization before or
shortly after deposition at the peat surface (Taylor and Cook, 1962; Koch,
1970 ). The cellular, thick-walled tissue associated with strongly reflecting rod-
lets whose cross sections have been mistaken for sclerotia in former times are
sclerenchymatous strands with an inertinitic reflectance according to Lyons
et al. (1982). They are associated with resin canals in MeduUosa and related

PLATE III
(Width of photomicrographs in parentheses )
1. Well microlayered cutinite-rich clarite, probably deposited under water, characteristic of Saar
coals. The coalified leaves consist of relatively low reflecting perhydrous phyllovitrinite with
many inclusions of resinite. Seam 1, Luisenthal Mine, Saar District, F.R. Germany. Polished
section, oil immersion, 270×, (0.24 mm).
2. Gondwana coal from the Bowen Basin, Australia. The typical microlithotype is inertite, con-
sisting of a high concentration of inertodetrinite with inclusions of mineral matter (black). In
the center a thin band of vitrinite. Polished section, oil immersion, 500 × , (0.13 mm).
3. Reed coal from a light layer of the Rhenish brown coal, consisting of more than 90% loosely
packed humodetrinite. Note the crumbly texture. In the center a fungal spore (white). Polished
section, oil immersion, 250×, (0.26 mm).
4. Taxodiaceae-forest coal from a dark layer of the Rhenish brown coal with preserved cell struc-
tures (humotelinite). Polished section, oil immersion, 250 × , (0.26 mm).
5. Pollen-rich fine-detrital gyttja, deposited under water, from the Rhenish brown coal. Polished
section, oil immersion, 250×, (0.26 mm).
6. Finely layered dark cell walls of Marcoduria inopinata, rich in cellulose. The thin lignin-rich
middle lamella (arrow) is lighter grey. Between the cell walls dense porigelinite (light grey).
Polished section, oil immersion, 440 × , (0.145 mm).
7. A dessicated face of Rhenish brown coal in the Wachtberg open mine, showing the banding of
dark brown and lighter brown layers. (ca. 25 m).
8. "Schwelkohle" (used for tar production) from a pale layer of Eocene brown coal of eastern
Germany. The dark bituminite groundmass contains particulated liptinites which, however,
remain invisible under white incident light. The light tissue in the centre (humotelinite) rep-
resents an ingrown root. Polished section, oil immersion, 250 × , (0.26 mm ).
i~ ~ • ~

--.1
seed ferns. It seems that they are primary inertinites, just like the sclerenchy-
matous cells in the pith of Lyginopteris oldhamia (see. p. 24). Later, Lyons et
al. (1986) proposed the name "secretinite" as a maceral term for these rodlets
and other highly reflecting cell secretions. Recently Cohen et al. (1987) de-
scribed secondary cell fillings from modern peats. They are relatively dark in
transmitted light and their formation seems to be related to fungal attack of'
certain tissues. They may form "secretion sclerotinite" in the hard-coat stage.

The term inertodetrinite is used for small particles of inertinite (Plate III-
2), for example, splinters of pyro-fusinite and relics of degrado-semifusinite
which may have been blown in by the wind or washed in by water. Inertode-
trinite is a characteristic maceral of subaquatic coal facies and of clastic rocks.
Smyth (1980b) assumes that the inertodetrinite of Triassic Australian coals
which were deposited in a sub-arctic climate was derived from mosses and
lichens.

Among the inertinite macerals micrinite plays a special role. Although it is


highly reflecting, micrinite may be sensitive to oxidation and heating (Stach,
1936; Nandi and Montgomery, 1967). The author observed transitions from
resinite cell fillings into micrinite in Upper Silesian low-rank bituminous coals
(Teichmfiller, 1944) and pointed to the striking abundance of micrinite in a
marine-covered Ruhr coal, which is characterized by petrographic and tech-
nological properties that suggest a relatively anoxic milieu of deposition
(Teichmiiller, 1955). Moreover, micrinite is a characteristic maceral of sap-
ropelic coals (Plate II-6). For these and other reasons the author, in a paper
on the origin of micrinite, concluded that the main part of micrinite generates
from lipoid substances during the coalification process. This is due to a dispro-
portional reaction that leads to the formation of petroleum-like substances on
the one hand and to a highly reflecting, solid relic, namely micrinite, on the
other hand (Teichmfiller, 1974a,c). Thus a great part of micrinite must be
regarded as a secondary maceral that forms from oil-prone macerals, mainly
bituminite and perhydrous vitrinites, but also from some resinites and spor-
inites during the stage of low-rank bituminous coal, which corresponds to the
"oil window" in minerogenic oil source rocks. In the meantime, this concept
has been confirmed by observations of organic petrologists in source rocks
(among others, Teichmiiller and Ottenjann, 1977; Cook, 1982; Gutjahr, 1983;
Sherwood and Cook, 1986; Cook and Struckmeyer, 1986). Cook and Struck-
meyer (1986) pointed to the abundance of micrinite in vitrinites of Australian
coals and assumed a generation from submicroscopic lipoid substances that
Taylor (1966) found in vitrinites under the electron microscope.
Observations on Florida peats (Cohen and Spackman, 1980) suggest that
part of micrinite is derived from certain parts of cell walls in some plants.
Cohen et al. (1987) observed dark-brown granular material in Salix leaves only
29

when these were shrivelled due to drying. Moreover tissues showing evidence
of microbial attack were often darkened around the area of attack. The authors
suggest that the darkening process ("inertinization") occurs under more aero-
bic conditions than those in which huminites/vitrinites are formed.

1.3.2 Coalification of inertinites


As most inertinites are already pre-coalified (so to speak) they do not change
as much as vitrinites and liptinites during coalification. Of course, semi-iner-
tinites change more than real inertinites. It is to the merit of Smith and Cook
(1980) who revealed that a great part of inertinization (in terms of rising re-
flectance) takes place no earlier than during the brown-coal and subbitumi-
nous coal stages (Fig. 1 ). This must be a disproportional reaction during which
aromatization (inertinization) occurs on one hand and expulsion of hydrocar-
bons on the other hand (Teichmfiller, 1987a, pp. 144-145). This reaction is
comparable to the formation of micrinite in the bituminization range of coal-
ification. Plate II-3 to 6 shows examples of secondary inertinites that were
formed during the coalification process. In these cases, inertinites formed from
former liptinites due to disproportional reactions during and after bitumini-
zation. In the stage of meta-anthracite, the initially relatively high reflectance
of inertinites is surpassed by the reflectance of vitrinites (Alpern and Lemos
de Sousa, 1971 ), just like during carbonization. This is due to the higher H-
content of vitrinites and their resulting earlier tendency to pre-graphitization
(Teichmiiller, 1987b, p. 146).

2 ORIGINOF COALFACIES (MICROLITHOTYPESAND LITHOTYPES)

2.1 Distinction of coal facies

The different genetic types of coal lithotypes, recognizable by eye, are char-
acterized by the types and amounts of microlithotypes and their macerals and
mineral inclusions as well as by textural properties such as the extent of mi-
crolayering {see Fig. 8). Microlithotypes refer to the naturally occurring as-
sociations of macerals both as seen under the microscope. Chemically, sulfur
and nitrogen as well as the hydrogen content of vitrinites are facies dependent.
Table 3 shows for hard coals the main microlithotypes and their maceral com-
position as well as the main lithotypes in which they typically occur.
In brown coals, lithotypes rather than microlithotypes have been used so far
to distinguish certain coal facies (see pp. 47-59 and Figs. 13,15,16). The dif-
ferent methods of lithotype classification of brown coal vary widely, and no
international agreement has yet been reached. Generally, it has been accepted
by coal petrologists that:
(1) cannels and bogheads, the sapropelic coals, are subaquatic muds deposited
in still-water ponds,
30

TABLE ;~

The main microlithotypes, their maceral composition, and the lithotypes in which they occur

Main microlithotypes Main constituents Main lithotypes


(microscopic) (microscopic) (megascopic)
Vitrite vitrinite bright coal
Clarite vitrinite, liptinite semi-bright, dull coal
(dull if much liptinite)
Durite inertinite, liptinite dull coal
Vitrinertite vitrinite, inertinite bright, dull coal
( dull if much inertinite)
Trimacerite vitrinite, liptinite, mostly dull coal
inertinite
Inertite inertinite dull coal
Liptite liptinite dull coal
Fusite, semifusite fusinite, semifusinite fusain layer, dull coal
Inertodetrite inertodetrinite dull coal
Carbargilite coal + clay minerals clay parting

(2)microlithotypes rich in huminite/vitrinite indicate relatively wet condi-


tions of peat formation, especially when combined with clay partings and
inclusions of syngenetic pyrite,
(3)microlithotypes rich in inertinites, especially in fusinite, semifusinite and
macrinite, are generated from peat types deposited under relatively dry
conditions.
Some authors regard the vitrinite/inertinite ratio as an inverse index of the
degree of oxidation to which the coal-forming peat has been exposed (Smyth
1984, Harvey and Dillon 1985, in part also Diessel 1986).
In the following discussion of the influence of various ecological and vege-
tational conditions upon coal facies, account is taken of results of recent work
which were not considered by M. and R. Teichmiiller ( 1982 ) in the "Textbook
of Coal Petrology" (Stach et al. 1982).

2.2 Influence o[ depositional (palaeogeographic) environment

Environments of peat formation are generally slowly sinking depressions


where mineral input is nil or very small, and in which the groundwater table
can keep abreast of peat formation. Such mires are called "topogenic" or "low
moors". Only in areas of very high rainfall "ombrogenous" mires or "high
moors" which include "raised bogs" and "blanket bogs", may form above the
groundwater table. The term "mire" is used here in the sense of Moore (1987)
and McCabe (1987) as a habitat in which organic material, especially peat is
accumulating.
Most large peat-forming mires occur in flat coastal plains, often protected
31

by sand bars ("overbank swamps") in lagoons and in deltaic areas. Such sites
of "paralic coals" are often distinguished from those of "limnic coals" which
were deposited farther inland along rivers ("back swamps"), around lakes (in
a lacustrine environment) or in other depressions, e.g., those of glacial origin.
A new trend of coal petrology is to search for relationships between the pal-
aeogeographical-sedimentological position of peat formation and the petrol-
ogical composition of coals (e.g., Smyth, 1979, 1980a, 1984; Styan and Bustin
1983a,b; Harvey and Dillon 1985; Diesse11986; Cohen et al. 1987). The authors
compare the facies (lithology, palaeontology) of the coal-bearing strata with
the facies of the accompanying coal. Pioneering work was done by Hacquebard
and Donaldson (1969) and by Smyth (1975, 1979, 1984). Figure 5 shows a
scheme of depositional areas of coal formation (fluvial, lacustrine, upper and
lower deltaic, lagoonal) and the main associated microlithotypes of Australian
coals. Durite + inertinite-rich coals were found to have been deposited mainly
in lakes and in the lower delta plain, whereas vitrite + clarite-rich coals oc-
,,:!:#:~?:
i, ~( ....
...'.~':'L"(: I A LACUSTRINE
CI .:Cic!T~I~,::.~ •..... B FLUVIAL
( ) " ~ ! ! ! ~ ~ I ~ " F L U V AL (b) C BRACKISHWATER
; ' ~ ~i' J {GRETACOALMEASURES}
"~'~;~"["~':;'. | [~ UPPERDELTAC
UPPER~~"~!~ !::~'~::. ~, E Lo WER DELTA,C
t ....

DELTAIC:~ ~ i

VITRITE +CLARITE
FLUV,AL

!~ AREASOF COAL /,~ - . . . ~ - ,,~,~v,~ / _.~


........ ACCUMULATION / [) i " "*'~'.~''I~'//////A~-~

10 50 90
INTERMEDIATES DURrFE+ INERTITE
Fig. 5. Depositional milieus of peat formation (a) and related microlithotypes (b) of Australian
hard coals. From Smyth (1984).
curred in upper deltaic fluvial and lagoonal environments. Intermediates be-
tween these two petrographic types are regarded as characteristic of delta plains.
Smyth (1984) regarded the ratio vitrite + clarite : durite + inertite as an indi-
cator of oxidation of the peats of the liptinite-poor Australian Gondwana coals.
She suggests (1980a) that especially thick Australian seams t > 15 m) which
contain much inertinite have been deposited in small basins on a stable craton.
In contrast to seams deposited in foredeeps, they accumulated at times of still-
stand either in lacustrine, partly fluviatile environments, or in raised bogs.
The majority of Gondwana coals of Australia, India, South Africa and Bras-
ilia (?) are rich in dull coal, often with pure inertite (a microlithotype with
more than 95% inertinite! ). These coals have mainly been deposited in depres-
sions of glacial origin (lakes, valleys). Some Gondwana coals, however, have
been deposited in foredeeps, for example, in the Permian foredeep running
north-south along the present eastern coast of Australia and are rich in vitrite
and clarite as, e.g., coals in the northern coal field of the Sydney Basin (Shi-
baoka and Smyth, 1975).
Recently, Diessel ( 1986 ) distinguished between Australian Gondwana coals
deposited in lower and upper delta plains and in piedmont plains. He intro-
duced a "Gelification Index" (GI) and a "Tissue Preservation Index (TPI) to
characterize these coals:
vitrinite + macrinite
GI = semifusinite + fusinite + inertodetrinite
telinite + telocollinite + semifusinite + fusinite
TPI-
desmocollinite + macrinite + inertodetrinite
Wet conditions of peat formation are distinguished by high GI-and high TPI-
indices, whereas dry conditions lead to low GI- and low TPI-indices (Fig. 6).
The latter is the case for seams extremely rich in inertodetrinite which, ac-
cording to Diessel, were deposited in piedmont plains where "severe oxidation
restricted the formation of telinite and telocollinite, and under conditions of
falling water table, even structured inertinite will disintegrate to form in-situ
inertodetrinite, commonly coupled with an increase in inherent ash and the
rather resistant sporinite". On the other hand, coals deposited in upper delta
plains and in fluviatile environments are rich in vitrinites (wet-forest swamps
in Fig. 6) but also in clastic clay minerals. Brackish coals deposited in delta
plains, partly as marsh peats, are distinguished by a high Gelification Index
and a low Tissue Preservation Index, as well as by high amounts of pyrite and
organic sulfur, due to marine transgressions.
According to Neavel ( 1981 ), coals deposited in lagoons carry abundant hu-
modetrinite/desmocollinite, whereas coals of landward {freshwater) environ-
ments are though to be richer in telinite, resinite and inertinite.
Figure 7 shows that the vitrinite: inertinite ratio in the Herrin coal of Illinois
33

decreases TREE DENSITY kfcrea,es


100.0 I
Ilmno.telmatlc telmatlo
50.0 I

MARSH 0

0 0

10.0 -- o o o~ ' ~ ' ~


- - ~ ' 0 0 ZI -

• u 0 '~ A

. -ere, #.~.....I~.A ~,.~-- A sw,,P ,~"

-o/
o., -
--
" F
: RES
:oZ - - -

G~ l terrestrial
. . . . w . . . . f . . . . , . . . . r . . . . , ,

0.5 1.0 1.5 2.0 2.5


TPI
Fig. 6. Coal facies in terms of the GelificationIndex (GI) and the Tissue Preservation Index
(TPI) in relation to depositionalsettings and types of mire. From Diessel ( 1986). Circles= coals
deposited in lower delta plain; triangles--coals deposited in upper delta plain; hexagons= coals
deposited in piedmont plain. Coals studied: Ba=Bayswater; Gr=Greta; Ho=Hoskisson;
Me = Melville, Wy = Wynn.

is highest (12-27) adjacent to palaeochannels (where high trees grew due to


nutrient supply) and lower (8-11) progressively away from them (Harvey and
Dillon, 1985 ). Parallel changes in the abundance of L y c o s p o r a (from lycopod
trees) have been observed (Phillips and Cecil, 1985) (cf. Fig. 11 ). High liptin-
ite contents were found in the vicinity of channels (Harvey and Dillon, 1985).
Of special interest are observations of recent mires and associated peats.
Styan and Bustin (1983a,b) described three peat-forming environments in the
Frazer River delta:
(1) An inactive portion of the distal lower delta with salt and brackish marshes
where thin, sulfur-rich peat layers are deposited and alternate with pre-
dominant layers of clastic sediments.
(2) Fluvial dominated areas between the lower and the upper delta plain where
sedge-grass marshes and the thickest peat deposits occur in abandoned
channels. The peats are rich in clay, silt and sulfur.
(3) The alluvial upper delta plain with sedge-grass peats and gyttjae, both rel-
34

: : :i : :::: :::::::!ii!iiiiiii!!ii!i!iiiiiiiiiiiii!i!!I¸~¸¸~" •

•-.-... ~ .....

Vitrinite- InertiniteR
(mineral and micrinite f ~ ~ ~

~8-11

Fig. 7. The v i t r i n i t e : inex~inite ratio in the Hex~rin No. 6 coal of Illinois in r e h t i o n to paheogeo-
graphical conditions.From Harveyand Dillon (1985).
35

atively poor in clastic minerals and sulfur. The authors point out that this
is the only environment where mineable coals would develop.
Cohen et al. (1987) consider that deltaic models for coal formation have
been greatly over-applied by geologists, because, in deltas, the influence of in-
organic sedimentation is too great to produce mineable coal seams. In deltas,
only convex peat deposits, caused by high rainfall, would prevent the influx of
mineral matter (see also Styan and Bustin, 1983 ). Recent examples of convex,
ombrogenous forest bogs in delta sites are known from Malaysia (Anderson,
1964).

2.3. Influence of climate, vegetation and water supply

The coal facies is highly dependent on the plant material that contributed
to peat formation. The vegetation in mires is influenced by the climate (which
may vary from sub-arctic to tropical), by the height of the groundwater table
(wet to dry), by nutrient supply, acidity, and marine influences. The height of
groundwater is most important not only for the kind of vegetation but also for
the redox-potential (oxic to anoxic) which determines the mode of preserva-
tion of plant remains. Peat may form in forest-swamps from various plant
associations, or in marshes with herbaceous vegetation (sedges, grasses), or in
open swamps with predominantly submerged and floating plants, or in raised
bogs with mosses or shrubs or trees. Therefore, it is usual to separate "forest
peats" from "reed peats", "subaquatic peats" (or "organic muds") and peats
of raised bogs. Plate IV (Fig. 1 to 4) shows the main types of peat-forming
vegetation and the hydrology of well known mires of the southeastern Atlantic
coastal plain of U.S.A.
Because geological age, palaeoclimate and the associated vegetation have a
great influence on the coal facies and because most coal petrological studies
have been carried out on certain coal deposits, the following sections will treat
mainly Carboniferous hard coals of Euramerica, Permian hard coals of Aus-
tralia, and Tertiary brown coals of Germany and Australia. In addition, certain
problems of coal facies which are under discussion will be dealt with separately.
Recent results on the influence of palaeoclimate will be found in section 2.4
(eutrophic-oligotrophic coals).

2.3.1 Carboniferoushard coals of Euramerica


Figure 8 shows an old scheme illustrating the genesis of various microlitho-
types for Carboniferous coals of the Ruhr District. Former organic muds are
the sapropelic coals cannel and boghead, deposited in still-water lakes or ponds,
followed landward by an "open moor" without trees, shrubs, sedges etc. but
with submerged plants and washed-in leaves and twigs, from which cutinite-
and sporinite-rich clarites and durites with a distinct microlayering were
formed. The reed-marsh zone in Figure 8 is represented by calamites standing
PLATE IV '~
37

in water as described by Scott (1980). The main and most important vegeta-
tion of Carboniferous coals were certainly lycopod forests with minor amounts
of pteridosperms and ferns as well as of cordaitean trees, the latter increasing
in the later Westphalian.
One may distinguish between a limnotelmatic forest swamp where the trees
were standing in deep water and a drier telmatic forest swamp. These forest
peats are the parent material for Carboniferous vitrites and liptinite-poor clar-
ites. If the groundwater table sinks below the peat surface, oxidation will take
place and will lead to the formation of inertinite-rich durites ("grey durains" ),
fusites and semifusites. This scheme has been used, partly with supplements,
by many coal petrologists, including Hacquebard and Donaldson (1969),
Kalkreuth (1982), Diessel (1986)and Marchioni (1981).
Hacquebard and Donaldson (1969) studied Canadian Carboniferous coals
by combining coal petrological with sedimentological and palynological meth-
ods. They distinguished the following petrographic types and interpreted them
as follows:

~grounbwa(e~
(able

forest-swamp with lepldophytes and pteridosperms cs&smlfaao reeds subaquatic deposits

fusites, dorites vitriles and sporinile- poor clarifes sporinite -rich sporinite* rich dorites, canneis and
poor in sporlnite clarites d~stlnctly mlcro-layered ciarites, bogheads
cutinife clarites

Fig. 8. The presumed moor types and the relating microlithotypes of Carboniferous coals. From
Teichmiiller (1962, 1982a).

PLATE IV
I. Margin of Taxodium (cypress) forest in the Okefenokee Swamp of Georgia (U.S.A.) with
bushes in the undergrowth and open-water plants in the foreground.
2. Reed marsh of the Everglades, southern Florida (U.S.A.), in the wet season. A tree island
("hammock") in the background. Blooming water lily (Crinurn) in the foreground.
3. Algal mat between sedges (Eleocharis) in the reed marsh of the Everglades, southern Florida
(U.S.A.). In the wet season the water table stands half a metre above the peat surface.
4. Change of moor type after a fire in the Okefenokee Swamp of Georgia, U.S.A. Charred stumps
and dried, thin stems of Taxodium trees are relics of the previous moor type (Taxodium-forest
swamp) before the fire. After the fire an open-water marsh with aquatic plants and reed grasses
and sedges developed as the present moor type.
Petrographic Type Environment, Vegetation
A = spore-rich clarite + duroclarite RM = reed moor (telmatic)
B = fusito-clarite with distinct FtM = forest terrestial moor
lenses of fusite ( relatively dry )
C = vitro-clarite + cuticle clarite FM = forest moor (telmatic)
D = duroclarite + durite + OM = open moor (timnic,
carbargilite subaquatic )

Lycospora is the typical spore for the forest swamps, whereas Punctatosporites
favouring "a more herbaceous type of flora" is characteristic of the reed-marsh
facies. Dull coals with spore-rich clarites and durites, together with clay part-
ings have a wide regional distribution and are interpreted as subaquatic de-
posits (see Fig. 8 ). This scheme, however, contrasts with ideas by Smith (1962,
1968) who, for the first time, developed a concept of oligotrophic durites, rich
in thick-walled microspores ("crassispores" of Stach, 1955) (see pp. 60-62).
It is not easy to reconstruct the vegetation of Carboniferous coals. The many
identifiable compressed plant remains occurring in accompanying rocks may
not be characteristic of the swamp flora. Spores may be blown into the mires.
According to Scott and Rex (1985) "vegetation preserved in coal balls repre-
sents the most complete data set for any plant habitat". Coal balls are early
petrified (mostly carbonatized, sometimes silicified) peats and occur in cer-
tain, often marine-covered seams.
Recently, valuable results were obtained by Phillips and Peppers (1984)
who evaluated coal-ball studies as well as palynological investigations carried
out on Euramerican Carboniferous coals. Some of their conclusions referring
to the Carboniferous climatic changes and the various types of vegetation in
Carboniferous mires are reported in the following.
The flora preserved with all anatomical details in the coal balls together with
spores and compression floras in clastic rocks indicate a moist tropical climate
throughout the coal-bearing Upper Carboniferous, with nevertheless wetter
and drier periods (Fig. 9). The wettest period was the Westphalian D, the
driest the Stephanian, although the Westphalian B was relatively dry too.
In the swamps, lycopods dominated in the Westphalian. According to Figure
10 cordaites were abundant only during the Westphalian B to D, with their
maximum in the Upper Westphalian C (Middle Pennsylvanian). Most Late
Pennsylvanian (Stephanian) coal swamps of the U.S.A. were dominated by
tree ferns. Wetter and drier environments in Illinois swamps are correlated
with Lepidophloios as the wet extreme and Sigillaria as the representative of
driest conditions. The highly dominating Lepidophloios forest stood in water
without canopies and ground plants. About 90% of the sporomorph Lycospora
come from Lepidophloios haUii. Figure 11 shows the distribution of this spore
in the Herrin No. 6 seam of Illinois. Interestingly, the highest amounts of Ly-
39

}-

b-

to
w

Fig. 9. A wet/dry climatic curve for Euramericancoal swampsduring the Pennsylvanian (West-
phalian A to Stephanian). The hatched part indicatesthe degreeof climaticwetness.From Phil-
lips and Peppers (1984).

cospora (indicating very moist forests) were found near and within a shallow
lake in the east and along the banks of a river running southward. The drier
sigillarian forests are thought to have had a higher diversity of plant assem-
blages with seed ferns and tree ferns as well as with a plant ground cover (al-
though this seems to have been different at places in the Ruhr Carboniferous,
as is shown on p. 65 and in Fig. 20 ). Cordaites became abundant in swamps
only where conditions were relatively dry or brackish.
Of great interest for facies problems of coal formation are the following ob-
servations of Phillips and Peppers (1984): "Pronounced changes in the kinds
and abundances of lycopod-dominated forests in the Herrin profiles usually
occurred at mineral-rich or clastic bands which indicate abiotic disruptions of
fresh-water regimes like drops in water level, flood, fire and changes in nutrient
supply." Commonly the flora changes from the predominance of one lycopod
to another lycopod (e.g., from Lepidophloios to Lepidodendron) and tree ferns
become more abundant. Seed ferns were usually more abundant immediately
below and above mineral bands. These observations clearly indicate topogen-
ous (not ombrogenous) peat formation in the Pennsylvanian of Illinois (cf.
Section 2.4 ). Similar results were obtained by Mahaffy (1985) who compared
the spore assemblages in coal balls with the spores in the accompanying coal
of the Herrin No. 6 seam. He found a good agreement between the two. Five
horizons could be distinguished which are separated either by clastic bands or
by fusain layers.
40

(~ [[PENNSYLVAN/CARBONI
IANTUPPER I
FEROUS
I
~--~,
M,OSPORES,COAL)IM' DCuO."sT
X"E"TpESTERN EUROPEPEAT(COAL BALLS)
VIRGILIAN

STEPHANIAN

MISSOURIAN

MAXIMUM
EIR TREE F E R N S ~
|~"..:-:!t.YCOP0 DS :':'..':'..~
DECLINE ~':".~:[LYC0 P0 D$ ~'.'::'_::]82%

DECLINE
WESTPHALIAN D ~/",,~ CORDAI TES ~/',/A 1%
[///2 C0ROA GE S>"/~ DESMOINESIAN

MAXIMUM
r///~. CORDA IT E S~/.//~>
ATOKAN 1WESTPHALIAN C
in
I~XPAN$1QN B
o R
~TREE FERNS~
i WESTPHALIAN B 0
EXPANSION
MORROWAN i
(/_.//~CORDAITES~/A IO%
#:
WESTPHALIAN A
MAXIMUM
[,:~:?:': LYCOPODS~i:'",~ ! ?~:!:' LYCOPODS~!~."::.196%

Fig. 10. The stratigraphic occurrence of lycopods, cordaites and tree ferns in Carboniferouscoal
swamps of Euramerica, on the basis of palynological results from Illinois coals (left) and of bo-
tanical identifications in Euramerican coal balls (right). From Phillips and Peppers (1984).

Phillips and DiMichele (1981), in their thorough study of coal balls in the
Herrin No. 6 seam of Illinois, described the contribution of different plants and
plant organs. Lycopods contribute 63-73% of the peat, ferns 15-17%, pterido-
sperms 6-16%, sphenopsids 4-5% and cordaites less than 0.7%. The lycopods
were the most important contributors for bark and root (Stigmaria), Psaron-
ius tree ferns are largely presented by roots whereas pteridosperms like Med-
ullosa contributed much foliage, and the sphenopsids are mainly represented
by wood from stems and roots. Four to five percent of the plant tissues are
fusinized and belong to aerial tissues of lycopods, pteridosperms and sphen-
opsids, the latter two contributing particularly to the formation of fusinite.
The predominance of root tissues in the petrified peats of coal balls has been
described by most investigators. Plate II-7 shows a petrified root peat with
41

,~:,-, .............................. !

1
I
J I
I

i,
L

60 km
"t 7 ~

Present Limit of Hel


: ~ Walshville channel - - ",l,~ ,

- - 6 0 - - Percent LycOspora
• Sample site
~':'~,'~, ]<e,>~Uclry
Lake

Fig. 11. Relative abundance of Lycospora in the Herrin No. 6 seam of Illinois in relation to the
palaeogeography of the Carboniferous swamp area. From Phillips and Peppers (1984).

cross sections of stigmarian appendices from the Katharina seam of the Lower
Rhine District, and in Plate II-1 ) the rootlets are ingrown in a stem from the
seed fern Lyginopteris oldhamia. Phillips et al. (1985) gave quantitative infor-
mation: according to their observations of coal balls the "shoot:root ratio"
which would be 4:1 for a whole tree, commonly is only 1:1. This indicates enor-
mous loss of aerial plant material. The shoot: root ratio is highest ( > 1) in
lycopod peats in which periderm is predominant compared with wood (36%
periderm, 4% wood). Lycopod peats predominate in the Lower and Middle
Pennsylvanian. They are much rarer in the Upper Pennsylvanian (Stephan-
42

Jan) when the climate became drier and the shoot:root ratio declined to a
minimum of 0.4! Generally, it must be assumed that lignified roots of lycopods
are the source material for many telinites in Carboniferous coals. It is also
known that the thick cortex of lepidophytes played a much more important
role for vitrinite formation than lycopodean wood. According to Phillips et al.
(1985) the ratio periderm:wood is 8.5:1 in Westphalian coal balls which are
highly lycopod dominated and carry only small amounts of cordaitean remains.
Comparisons of petrified Carboniferous peats with the petrological compo-
sition of the surrounding coal are still very rare. Former studies (Teichmfitler,
1952; Teichm~iller and Schonefeld, 1955) have shown that:
(1) The most abundant and best preserved tissues in coal balls are from roots
that grew through the peat.
(2) Forest peats are precursors of coarsely banded bright coals with vitrites and
clarites.
(3)At least part of liptinite-rich durites is derived from highly decomposed
organic muds containing many spores and inertodetrinites. Neither the mi-
crinite nor the macrinite of these durites can be detected in the original
mud petrified in siderite concretions. The compaction ratio of these muds
is extremely high.
(4)The peat of the marine-influenced Katharina seam (Westphalian A/B
boundary) has been formed from a vegetation rich in ferns and pteridos-
perms, especially in Lyginopteris oldhamia (Plate I-l). Lycopodean root
appendices of a later vegetation are abundant. The coal formed from this
peat consists of finely laminated bright coal with vitrite and clarite. It is
poor in liptinites and inertinites, although micrinite (not present in the
peat) is a characteristic maceral.
In silicified coal balls of the Namurian, Teichmfiller and Schonefeld ( 1955 )
found small sticks, leaves and cones of calamites which are highly predominant
over fine remains from ferns and lycopods. No inrooting was observed. This
peat type represents the end of seam formation (of seam HauptflSz) and is
obviously derived from a calamitean marsh. It is shown in Plate II-8.
Recent comparative studies by Winston (1986) between coal balls and coals
from Pennsylvanian seams of Illinois showed that lycopod periderm from stems
and from roots, as well as Diaphorodendron cortex, Myeloxylon, Pennsylva-
nioxylon wood and bark, marrattialean fern leaves, Psaronius outer roots and
Stigmaria rootlets can all be traced from coal balls into vitrinite layers, whereas
tissues traced into fusinites include lycopod stem periderm, Myeloxylon, and
Psaronius inner roots, the latter pointing to charring of the peat (see p. 69 ).
Winston calculated compaction ratios from peat to coal in the range between
7:1 to 19:1, in the case of Stigmaria periderm even of 30:1. A ratio of 7:1 has
been estimated from a coal ball in the vitrinite-rich Katharina seam of the
Ruhr Carboniferous, and for a liptinite-rich durite a ratio of 20:1 was found
43

(Teichmiiller and Teichmiiller, 1982, pp. 17-18), the latter easily explainable
by the high water content of organic muds.
Irrespective of coal-ball studies, the evaluation of the literature has shown
that the opinions about the genesis of hard-coal lithotypes and microlithotypes
may vary, especially in regard to durites. Most coal petrologists agree that coals
rich in vitrite and liptinite-poor clarite have been deposited in wet-forest
swamps. It is also mostly acknowledged that cutinite-rich clarites (Plate III-
1 ) represent subaquatic deposits near lake banks, and that inertinite-rich dur-
ites ("grey durains"), same as inertites (Plate III-2) were derived from oxi-
dized peat surfaces. But there are great differences in concepts about the gen-
esis of dull coals with liptinite-rich (other than cutinite-rich) clarites and,
especially, with liptinite-rich durites ("black durains" ). One group considers
these microlithotypes as a product of oxidation which formed under relatively
dry conditions (Smith, 1962,1968; Littke, 1985,1987; Bartram, 1987) where
spore exines were enriched due to their chemical stability (see pp. 61-63).
Another group interprets spore-rich durites and clarites as subaquatic deposits
that formed in open water or in reed marshes, where, just as in sapropelic muds,
allochthonous pollen and spores ("pollen rain" ) were well protected from de-
cay under water, whereas in the forest canopy they were destroyed. This con-
cept is strengthened by observations that liptinite-rich dull coals are often
associated with clastic layers. Marchioni (1980, p. 49) found that dull coal
layers in the Liddell seam of the Sydney Basin (Australia) are closely associ-
ated with clastic bands, although these dull coals certainly do not contain the
Densospore-rich type of durites but rather belong to the inertinite-rich "grey
durains". Hower et al. (1987) report that in the western Kentucky No. 11 seam
(correlative of the Herrin No. 6 seam of Illinois) which is distinguished by a
very high vitrite content (85% vitrinitel), durites, vitrinertoliptites and pure
liptites become more common as the blue-band parting is approached. These
spore-rich microlithotypes occur below as well as above the blue band*.
It must be stressed that Smith (1962,1968) has distinguished between, on
the one hand, "black durain" comprising clean durite rich in spores and ma-
crinite and characterising his "densospore phase" and, on the other hand, "grey
durain" comprising durites with some mineral matter, fewer spores (tenui-
spores instead of crassispores) and semifusinite as the characteristic inertinite
maceral. The grey durain belongs to his "incursion phase" which is considered
to represent inundations with oxygenated water transporting poorly preserved
spores and inertinite (cf. pp. 61-62). Similarly, Smyth (1980a) found durites
below clay bands in Australian coal seams which she interpreted as oxidized
plant remains washed in by flowing water. According to Neavel (1981), durites

*On the other hand, at the edges of the same coal deposit striking concentrations of inertinites
occur, partly looking like the "steinkohlenartige Teilchen" of German brown coals which are in-
terpreted as redeposited particlesof oxidized peat (Teichmiiller,1982a, p. 281, figs.86a, b).
44

may come from peats deposited either under extreme wet or under extreme dry
conditions. Likewise, Styan and Bustin (1983b) concluded that precursors of'
durite in peats of the Frazer River delta may be deposited either under water
as organic muds or may be oxidation (dessication) products of sedge-grass
peats.
The uncertainty about the genesis of durites seems to be connected with the
uncertainty about the genesis of the typical inertinites of durites, namely ma-
crinite and semifusinite of the degrado-type. Both macerals are very rarely or
not at all found in peats and brown coals. From this point of view, studies of
differences in opacity and reflectance of huminites in recent peats carried out
by Cohen et al. (1987) are of great interest. These authors found that: (1) a
darkening process ("inertinization") of cell walls occurs under more oxygen-
ated conditions, mainly through fires, more seldom through microbial activity;
(2) that humocollinites (gels) are always higher reflecting than humotelinites
{cell walls); and (3) that gelification in peats takes place mainly in those types
which were deposited in a deeper water aquatic habitat, for example, in Nym-
phaea peat. These observations may permit the following explanation: the ba-
sis for the formation of inertinitic groundmasses in durites is founded already
in the peat stage, when gels with a relatively high, although still huminitic
reflectance did form, but the true inertinitic reflectance is attained later in the
hard-coal stage. This explanation is consistent with the concept of a secondary
inertinization during coalification (see pp. 25,29 and Fig. 1 ).

2.3.2 Gondwana coals (and peats of cool climates)


In contrast to Carboniferous coals of the northern hemisphere, which were
deposited in a tropical climate, Carboniferous and Permian hard coals of the
southern hemisphere were deposited in sub-arctic to cool climates after the gla-
ciation of Gondwanaland. The vegetation of these Gondwana coals, of course,
was very different: in contrast to the luxuriant flora of Carboniferous forest
swamps with tall trees, the Gondwana vegetation was stunted with only shrub-
like, broad-leafed trees and bushes of Glossopteris and Gangamopteris and with
many herbaceous plants. Therefore, many Gondwana coals contain relatively
little vitrinite and do not contain thick bands of vitrite. These coals tend to be
dull in lustre, due to very high contents of inertinites, mainly semifusinite and
inertodetrinite (Plate III-2 ). They may contain bands of one metre and greater
thickness which contain pure inertites (Snyman, 1961 ). Liptinites (mostly
cutinite) are very rare. The distribution of the various macerals is much finer
than in European Carboniferous coals, just as is the distribution of minerals
(clay, quartz, siderite predominate ). The finely distributed clay and sand min-
erals, according to Plumstead (1962), may have been blown into the almost
treeless mires from surrounding mountains that were still devoid of vegetation
after glaciation. Other authors think of precipitation from solutions or of a
remani~ effect (Smyth and Cook, 1976).
45

The high content of inertinites in many Gondwana coals* is generally ex-


plained by the relativelydry conditions of peat formation. But, recently Liu
and Taylor (1987), by means of electron microscope investigations (TEM)
were able to recognize very fine lipidmaterial between the inertodetrinitepar-
ticles.They assume that this material came from algae and their decomposi-
tional products. These findings point to subaquatic deposition.
It is striking that, according to Smyth (1980b), certain Triassic coals of
Australia, which were deposited in a sub-arctic climate with tundra and perma-
frost,are also particularly rich in inertinite (in part, more than 80%!). Smyth
suggests that the semifusites of these coals were deposited on tree islands,
whereas mosses and lichens tended to form the inertodetrinite in cutinite-
bearing durites and inertodetrites. However, a genetic relation between mosses
and inertodetrinite cannot be generalized, at least not for peats, because, for
example, Sphagnum peats are never rich in inertinites. In contrast, Pleistocene
sub-arctic peats, deposited during interglacial times, are very poor in inertin-
ites (0.1-2.0%) (Koch, 1966,1969b), and also modern peats of the Frazer River
delta in British Columbia (Canada) which were deposited in a cool climate
consist mainly of huminites with some liptinites. They would be transformed
to vitrites and clarites during coalification (at least in the absence of secondary
inernitization of cellulose-rich material). High amounts of inertinites were only
found where peats were exposed to flooding with oxygenated, neutral-pH water
and/or to dessication and oxidation (Styan and Bustin, 1983a).
Certainly, investigations of sub-arctic mires and peats (for example, in Can-
ada and the U.S.S.R. ) are needed to clarify the special petrological properties
of sub-arctic coals. According to Smyth (1980b), the development of sub-arctic
mires depends on the presence of continuous permafrost, which in the summer
dams the melting water at the surface. The low temperatures, the poor vege-
tation, and the annual change between extreme wetness in the summer and
dry periods in the winter certainly are very special conditions of peat formation
which might explain the special petrographic properties of Gondwana coals.
The vegetation of cool to temperature-climate mires are well known from
Europe and North America, but only few attempts have been made to study
these peats from the viewpoint of coal petrology. According to studies of Styan
and Bustin (1983a) in the Frazer River delta, where peat formation is still
going on, the vegetation is almost completely herbaceous and dominated by
sedges and mosses. The only plants with woody stems are ericacean shrubs and
stunted bushes as well as small trees of Pinus and Picea which occur in the
climax stage of the peat "biofacies". This is different in the different areas of
the delta (see p. 33-35). With the exception of the distal lower delta, coloni-
zation by Sphagnum mosses occurs after the peat substrate, which may be
several metres of sedge-grass peat, has been raised above brackish-water influ-

*Nevertheless there are some G o n d w a n a coals rich in vitrinite a n d even in pyrite!


46

ence. Thus, acidic, oligotrophic and drier conditions are created in which, be-
sides Sphagnum, Ericaceae and small pines can live.
As in North European mires (Koch, 1966,1969b; Grosse-Brauckmann,
1979a) woody plants indicate drier conditions of peat formation than do her-
baceous plants. Grosse-Brauckmann (1979a) even regarded alder (Alnus glu-
tinosa) peats as indicators of stillstand of peat formation. After the alder roots
have penetrated the peat, oxidation of the older peat takes place and may lead
to the destruction of older plant remains or to loss of a whole peat layer. Some-
times, concentrations of small sclerotia of the fungus Cenococcum geophilum,
which lives in peats above the water table, are the only relics of an older peat
layer.
Figure 12 shows various plant associations of mires in the Frazer River delta
with their peat types and the macerals and microlithotypes which might form
from them according to Styan and Bustin (1983a). These authors think that
"a high ratio of cellulose to lignin in marsh plants (sedges, grasses ) and limited
exposure of these tissues to dessication and oxidation produce primarily des-
mocollinite". Under freshwater conditions more telocollinite will form, under
marine conditions more desmocollinite. Together with the liptinite macerals
alginite, sporinite and "cerinite" (the latter from sedge lipids) clarites are
formed from sedge-grass peats. Flooding by oxygenated neutral-pH waters,

Peat type Gyttja peats Ericaceous Freshwater


Sphagnum peats Sphagnum peats Nuphar peats sedge-grass peats

do m ina nt liptodetrinite suberinite telinite liptodetrinite telocollinite


coal alginite telinite reslnite macrinite
macerals cutinite telocotiinite telinite sclerotinite
desmocollinite pyrofusinite cutinite cutinite

minor inertodetrinite desmocollinite cutinite suberinite telinite


coal micrinite oxyfusinite telocollinite desmocollinite pyrofusinite
rnacerals eutinite sclerotinite desmocollinite
se|erotinite cerinite
resinite
sporonite
dominant liptite and vitrite with vitrite liptite and clarite with thin
micro thin clarite bands o f clarite c|arite interbanded
lithotypes bands lenses o f liptite inertinite and
and durite vitrite

Fig. 12. Peat types in the Frazer River delta and the presumed macerals and microlithotypes which
will form from them. From Styan and Bustin (1983a).
47

followed by extended periods of dessication results in increases of inertode-


trinite, macrinite, sclerotinite and degradofusinite and the formation of inter-
laminated durites and vitrinertites as the microlithotypes. After colonization
by Sphagnum, lignin-rich tissues from ericaceous shrubs and pines provide
precursors of suberinite, telocollinite, and telinite. Coal seams originating from
such peats would be thick, laterally extensive and characterized by vitrites with
laminae of liptite and lenses of clarite. Alluvial peats are rich in carbargilites
with collinite, suberinite and vitrodetrinite. They may be overlain by gyttjae
that will form liptites and clarite bands.
According to the interpretations of Styan and Bustin (1983a), a character-
istic microlithotype of Gondwana coals, namely vitrinertite, may have been
formed from herbaceous peats after flooding with oxygenated water and sub-
sequent dessication. In this concept, the idea of "flooding" is important, es-
pecially when compared with hypotheses which regard inertinite-rich coal lay-
ers as oligotrophic oxidized peats formed in raised bogs (p. 62). The idea of
"flooding", however would fit well with the electron microscopical observa-
tions of "algal" material between inertodetrinites in Australian coals (Liu and
Taylor, 1987) as well as with the "grey durains" included in the "incursion
phase" of Smith (1962,1968) (see pp. 43-44, 62).

2.3.3 Tertiary brown coals of Germany and Australia


Brown coal still resembles peat. Many plant remains can be recognized with
the naked eye (woods, leaves, roots) or under the microscope (spores and pol-
len, cuticles, fungal remains) and are still identifiable by palaeobotanists. Par-
ticularly in Germany and in Australia attempts have been made to reconstruct
the peat-forming vegetation of Tertiary brown coals. By using the analogy of
conditions in modern mires and peats, palaeobotanists and coal petrologists,
working in cooperation, were able to distinguish certain coal facies. In most
cases lithotypes were the basis for facies differentiation.
Although the varying colour is the most striking property when viewing a
dried brown coal face, other megascopic properties are also used to classify
brown-coal lithotypes. In Germany, they are mainly banding {"non-banded"
to "highly banded"), occurrence of plant remains which are visible with the
naked eye (wood, leaves, resins, fusain, etc. ), degree of gelification (mainly as
seen under the microscope), and physical properties like hardness ("soft",
"hard", "crumbly").
Figure 13 shows an old, well-known attempt to correlate plant associations
and lithotypes of the Miocene Rhenish brown coal, which was deposited in a
subtropical climate. It is based on palaeobotanical, mainly palynological stud-
ies of Thomson (1950,1951) and on coal petrological observations by the pres-
ent author (Teichmfiller, 1950) as well as on chemical analyses of the various
facies types (Teichmfiller and Thomson, 1958). It should be stressed that
Thomson was a palynologist who had an excellent knowledge of Baltic mires
48

moor type: seq . . . . . . . . ! Myn . . . . . . Cyrgr. . . . . . . . . : Nyssa [axog . . . . . . p reed marsh !


!
[1~. . . . . ter

resulting coal,
dark brown coal with coalified tree stems (xylilicl lighter brown coaJ dark. lough brown coal
megascopic:t~ with stump horizons les~ stems more stems without stems (detmall (delritall

much humolelinile much humodelrinite


microscopic: 1 (lexlinite AL much homo/elinite, poorly much humo'lelinite, beller much hurnodetrinite.
and much liplinite,
, well preserved lissues preserved tissues preserved tissues very few tissues
i often clay minerats

Fig. 13. The presumed moor types and the relating petrograpmc compos~tlonot the Miocene
Rhenish brown coal. From Teichmiiller(1982a).

and peats. Figure 13 shows from the right to the left side, the horizontal se-
quence of different environments and plant communities from open water to
increasingly drier conditions: from submerged plants to a reed marsh, followed
by a wet Nyssa-Taxodium swamp, followed by a drier Myricaceae-Cyrillaceae
bush moor and a Sequoia forest as the driest, final peat-forming vegetation.
Although this scheme has been challenged by later authors (see below), the
author thinks that it still may be used to understand the main facies types of
the Rhenish brown coal which are indicated on Figure 13 according to their
megascopic and microscopic properties. Accordingly, open-water coals are dis-
tinguished by a dark, black colour (due to biochemical gelification) and high
amounts of humodetrinite and liptinite. Reed coals correspond to the "light
layers" of the Rhenish brown coal and carry the highest amounts of humode-
trinite ( > 90% ) and ten times more pollen than the forest coals (Plate III-3).
Forest coals are darker brown in colour, carry stems and stumps from trees,
and are microscopically distinguished by much more humotelinite (Plate III-
4 ) whose cell tissues are better preserved in the coniferous than in the angios-
permous forest coals. Sequoia coal is distinguished by thick stumps and espe-
cially well preserved wood-cell structures. Mukhopadhyay (1986) found sim-
ilar relationships in Tertiary lignites of Texas.
Nevertheless, the scheme of Figure 13 has been challenged by Von der Brelie
and Wolf (1981a,b) and Hagemann and Wolf (1987), the latter authors ap-
plying the combined palynological and coal petrological studies of the first
mentioned authors. Von der Brelie and Wolf (1981a) used the same methods
as Teichmiiller and Thomson (1958) but studied more single coal samples.
Their new interpretations are, however, based on three very short coal sections
with a total thickness of only ten metres (against 50-100 m for the whole
seam ). In contrast to Van der Burgh ( 1973 ) (see below ) the authors concluded
that the peat-forming vegetation consisted mainly of shrubs and trees from
angiosperms. Three types are distinguished on the basis of pollen: (1) a "mi-
crohenrici-type" with oak-like shrubs and trees; (2) a "megaexactus-exactus-
49

type with Cyrillaceae and Clethraceae; and (3) a "Triporates-type" with Myr-
icaceae. No indication of a Nyssa-Taxodium swamp or of reed marshes were
found. Instead, the reed coals of Teichmiiller and Thomson (1958) were inter-
preted as the oxidation product of the same peat that formed in the micro-
henrici-forest swamp with oak-like plants. Their high content of humodetrin-
ite and of pollen (which are of extrapaludal origin according to Teichmiiller
and Thomson, 1958) is regarded as the consequence of strong aerobic decom-
position. The microhenrici-forest is compared with "evergreen oak forests of
central Florida (which, however, grow on dry soils and not in mires ). Accord-
ing to Von der Brelie and Wolf (1981b) Sequoia peats were deposited in "very
moist to moist" environments and not under driest conditions as interpreted
originally by brown-coal palaeobotanists such as Jurasky (1940), Thomson
( 1950/1951 ) and Schneider (1986). The modern Sequoia sempervirens of the
U.S. Pacific coast (mentioned by Von der Brelie and Wolf, 1981b, to document
the requirement for wetness) grows on dry soils and not in mires. Moreover, it
is hard to understand why Sequoia-forest coals do not contain many well-pre-
served wood tissues, i.e., telinites (p. 184 of Von der Brelie and Wolf, 1981b),
because Sequoia wood is known as the most resistant wood of the present time.
Generally, Von der Brelie and Wolf (1981a) stressed the uniformity of the
flora throughout the time of peat formation and compared the vegetation of
the Rhenish brown coal with the uniform vegetation of modern topogenous
forest swamps ("Hochmoorwaldflachmoore") in Sumatra (see fig. 3, p. 152 of
their paper). On the other hand, they assumed a "more or less ombrogenous
origin" for the whole Rhenish brown coal (p. 150). Sphagnum spores were
found in the "Triporates"-forest type with Myricaceae.
The different results of Teichmfiller and Thomson (1958) on one side and
of Von der Brelie and Wolf (1981a,b) on the other side may be partly explained
by the different samples studied. For example, the latter authors (p. 154) ad-
mit that they probably did not study the same type of "light layers" that were
interpreted as reed coals by the earlier authors. This, indeed seems to be the
case. Moreover, in relation to the short profile studied by Von der Brelie and
Wolf (1981a) it must be assumed that only part of the facies of the Rhenish
brown coal has been included, particularly as environmental conditions and
plant associations normally vary in time and space (Cohen, 1968; Koch, 1966;
1970a; Styan and Bustin, 1983a,b). In addition, the present author thinks that
reconstructions of peat-forming vegetation cannot be founded solely on paly-
nological results (cf. p. 17). Thus, for example, it has to be considered that
pollen of Cyperaceae (sedges) and Gramineae (grasses) are easily decomposed
(Kilpper, 1960, p. 295) and that the amount of Cyperaceae-pollen in the sedge
peat of the Everglades is only 3-6% of the total pollen content (Riegel, 1965).
According to Figure 14, in the Everglades, Riegel (1965) found a good corre-
lation between peat-forming vegetations and pollen only in the case of Taxo-
dium-peat. No correlation was found for any other peat types. In the sedge
5/)

peat, allochthonous pollen (of Pinus and Chenopodiaceae) predominate and


the abundance of pollen is highest in this peat type, just as in the "reed coal"
of Teichm~iller and Thomson (1958).
Van der Burgh (1973), as a wood histologist, found a high predominance of
coniferous wood (76-84% of all woods studied in various mines of the Rhenish
brown coal). Most woods stem from Taxodiaceae (40-70%). Van der Burgh
could not find correlations between palynological and wood anatomical results
in the Rhenish brown coal, except for Sciadopitys horizons.
Moreover, a comparison of cuticles and pollen from the same coal samples
of the Rhenish brown coal showed that only rarely did these two plant entities
indicate the same flora (Kilpper, 1960, p. 301 ): leaves of Quercus, Betula, Al-
nus, Fagus, Rhus, Ulmus and Corylus were lacking although the pollen of these
plants generally was present in considerable amounts. As the above-mentioned
leaves are resistant their cuticles should have been found if they were depos-
ited. These findings of Kilpper ( 1960 ) support the opinion of Thomson ( 1950/
1951 ) that the quercoid pollen (Quercoidites) of the henrici- and micro-henrici-
types, which are so characteristic of the "light layers" ("reed coals"), have
been blown into the open marshes from a vegetation existing outside the
swamps.
The Miocene Lusatian brown coal of the German Democratic Republic shows
many similarities to the Rhenish brown coal. The lithotypes of the Lusatian
brown coal have been studied in detail by many brown-coal petrologists, mainly
in regard to their technical behaviour. To clarify the related vegetational gen-
esis and the peat facies, palynological methods were combined with coal pe-
trology, especially by Sontag (Sontag and Schneider, 1982 ). The most valuable
results were obtained through the "cuticular analyses" applied by Schneider
(1978,1986, and many other papers) who recently also used coal petrological
methods (Schneider, 1984) to clarify the genesis of the Lusatian brown coal.
Figure 15 shows a schematic reconstruction of the vertical sequence of veg-
etations in the main Lusatian seam with the relating types of coal facies. The
F-facies with very wet forests of Alnus, Liquidambar, Taxodium, ferns and
Glumiflorae represents a pre-stage of seam formation since it is combined with
clastic sediments. The real seam begins with a dark coal rich in stems, twigs,
leaves and roots of Taxodiaceae, i.e., the K-facies, represented by an eutrophic,
wet Glyptostrobus swamp. The succeeding A-facies generated from an angio-
spermous bush moor, drier than the Glyptostrobus swamp, with small trees and
bushes from Fagaceae (Quercus rhenana) and Ericaceae (Kalmia), as well as
with Magnolia and Kadsura as characteristic plants. This facies formed a dense,
relatively homogeneous non-layered to weakly layered brown coal that is dis-
tinguished by having the best briquetting properties. The mires of the A-facies
have been flooded periodically so that organic muds with well preserved oak
leaves were deposited from time to time. The thin G-facies is represented by
Glumiflorae, i.e., by grasses and sedges whose cuticles are well preserved be-
51
a
Abundance LEGEND
in pefcenl
Site 62-26
8O
Site 62-28
70
Site 6 2 - 2 9 -
6O

50
Site EE-11I
0
~ 40 Site 62-27I

TrlporahD Sallx Viti$ Cephalanthu$ Polypodium


S- 2o

I=
L_o Vt_-__ ~'~__ ~L ~-~
Chenopod Grami . . . . Cyp. . . . . . . ~agit~l~rla ' Compositoe ~voldites'
r~
~~'0~') 20
tl0 ~Type I

L~ Pinus Quercus Rhizophora

Pollen Type

-~ ~ 8

i
SITEE7 ~ L ~

20
I
SITE16

SITE16A

$awgrossenvlronmenti
Fig. 14. The abundance of pollen and spores in Taxodium-forest peats (a) and in marsh peats (b)
of the Everglades, southern Florida. From Riegel (1965). Note the scarcity of C~y]oeraceae and
Gramineae pollen (from the peat-forming vegetation) in the marsh peats.
52

M O ! c 0 ~J d L r Q Myrico Glumiiloroe Ouercus (ganzrcmdKj) Daahnoaene


,~,,,: P a I m a e X~suro Ainu,"

"~ ~ :/ r,li~.Smm, ~ lib . . . . . . . . ZM .

Fig. 15. Scheme of the verticalsuccession of moor types and the resultinglithotypes in the Miocene
Lusatian brown coal (German Democratic Republic). From Schneider (1986). Conifers are in-
dicated sidewards, angiosperms are indicated at the top. Lithotypes are indicated by the following
signs: ~ ~ streakily banded, = = well banded, - - w e a k l y banded, empty white = non-banded.
H B = light band; Z M = bone coal; L = basal coal.

cause the leaves were deposited under still water so that, later, a gelified, well
stratified black band developed as a special lithotype ("sedimentary Glumi-
florae-facies", in contrast to the "sedentary Glumiflorae facies" of Schneider,
1980). At the boundary between the A-facies and the P-facies, Palmae occur
as characteristic plants. The P-facies generated in an oligotrophic, relatively
wet moor with Pinus, Lauraceae, Cyrilla and Myrica. It is represented by a
dark, gelified, well-layered brown coal lithotype with many inclusions of leaves,
barks and roots. The driest and final stage of seam formation (or of a cycle
within the seam) is the M-facies with Sciadopitys and Sequoia, and with (from
the angiosperms ) probably Sapotaceae and Apocynaceae. Most characteristic
of this facies is the abundance of Marcoduria, striking long bands, oriented
approximately parallel to the bedding, thus causing a well layered, streaky
lithotype which is also distinguished by stump horizons. The M-facies usually
is followed by a pale band rich in clastic minerals which represents the end of
seam formation or of a seam cycle.
The origin of Marcoduria has been discussed ever since Weyland ( 1957 ) who
erroneously interpreted the remains of this plant as the epidermis of an inun-
dated water plant (Helobiae). According to the most recent results of Schnei-
der (1989), however, it represents a tissue which separated aerenchymous tis-
sues from the cortex of certain coniferous roots. Marcoduria is a very striking
constituent of German Miocene brown coals, also as seen under the micro-
scope: the thick, sclerenchymatous walls of large, open cells are very low re-
flecting and highly fluorescent (Plate III-6). In thin sections they are aniso-
tropic, obviously due to high cellulose contents.
The brown coals of Victoria, Australia, although originating from types of
vegetation that were different from those of the German brown coals, have
53

similar petrographic properties. According to George (1975) the prevailing


plant associations had a great influence on the megascopic and microscopic
appearance of the different lithotypes. George's earlier results have been
strengthened by the findings of palaeobotanists and geochemists such as
Blackburn (1980a,b,1981), Luly et al. (1980), Johns et al. (1981), Verheyen
and Johns (1981), Kershaw and Sluiter (1982), Verheyen et al. (1982,1984),
Chaffee and Johns (1985) and Kershaw et al. (1984). For vegetational details
the reader is referred to these papers. Blackburn (1980,1981), who identified
numerous plant remains in the various lithotypes of the Yallourn seam, con-
cluded, on the basis of cluster and discriminant analyses, that the lithotypes
are closely related to the fossil flora contained in them. This Australian pal-
aeobotanist even developed a "diagnostic key" to the lithotypes of the Yallourn
seam by using fossil-plant assemblages with frequency data of certain plant
fossils. He found an increased prevalence of darker lithotypes towards the top
of the seam section (Blackburn 1980, p. 48), in contrast to Mackey et al. ( 1985 )
who studied the lithotypes of the older Morwell seam (see p. 54). Blackburn
(1981) emphasizes the strong influence of subsidence and fires upon vegeta-
tion and lithotypes in the Yallourn brown coal. Figure 16 shows the presumed
plant associations of the variously coloured lithotypes in the Latrobe Valley
brown coal according to Luly et al. (1980). Similarily, as shown in Figure 13
for the German Miocene brown coal, the pale and light layers of the Australian
brown coal are interpreted also by these Australian authors as deposits in open-
water lakes, whereas the darkest lithotype is represented by a plant community
with a significant sclerophyll component, thought to have colonized raised bogs.
According to an overview of Gloe (1984), the palaeobotany (plant groups and

A Group AI Group B IGroup C IGroup D


D
I I I
I Group I Group 2 1 Group 3 Group •
oPE, c0,
SAMPLES ' f.. "'I i...*.'.': . . . . .I
j i i
LITHOTYPES MEDIUM
\ \ \ \LIGHT
~\\'~ ~ \ PALE

,NOEX I I I I
I I I I
I I I

C Eaacfldaceoe/

B ®L2,1 N'::::

DEPOSITIONAL C ~ , ~ , i , ~ Gleick~Jall*lllo~e*oe Epaerld~e~ . " 2i':. ".'i:2"i;i.):::;: ::':)i


~.NVtRONMENTS swamp swamp s~mp tronl,hon Open Woler

Fig. 16. Relations between the colour of lithotypes (A) and the presumed vegetation and deposi-
tional environments (B) in the Latrobe Valley brown coal of Australia. From Luly et al. ( 1980 ).
54

genera, pollen types) of the Latrobe Valley coals is closely associated with
lithotypes: the darker lithotypes generated from swamp forests while the lighter
layers were deposited in an open-water environment.
Very recently, Kershaw et al. (1988) evaluated palaeobotanical studies on
pollen and macrofossils as well as coal-petrological studies on the lithotypes
within all major coal seams of the Latrobe Valley. They found that in all seams
there is a development of hydrological stages from open water through swamp
forests to raised bogs, running parallel with increasing darkness of the litho-
types. The pale and light layers are interpreted as deposits in open water (see
Fig. 16).
Nevertheless, Mackay et al. (1985) offered another concept. They studied
the lithotypes of the Morwell No. I brown coal in dried core samples from four
boreholes covering a seam having a thickness of 100 m. In contrast to Von der
Brelie and Wolf (1981a) and Hagemann and Wolf (1987), they stress the value
of coal colour for the determination and correlation of lithotypes which they
regard as facies dependent. These Australian authors found that the Morwell
No. 1 seam tends to exhibit lightening in brown colour upwards, and that also
single cycles in the seam begin with a dark band and end with a light band.
These cycles are compared with the cycles that Smith (1962) and Smyth and
Cook (1976) suggested for hard coals and interpreted as a development from
wet to drier conditions of peat formation (see pp. 60-63). Referring to the
interpretion of light layers as products of aerobic decomposition of the same
peats that formed dark layers in the Rhenish brown coal (Von der Brelie and
Wolf, 1981a), Mackay et al. (1985) concluded from their soley megascopic
studies that "if the moisture condition....is indeed the dominant factor in de-
termining lithotype, then the most likely explanation for the environment of
formation of these sequences is a wet or waterlogged peat swamp commencing
the cycle, with the peat accumulating on a gradually drying surface." These
results are, however, not in accordance with geochemical findings of Chaffee
and Johns (1985) on a 100-m-thick section of bore cores. These authors state
a "strong support for the view that different lithotypes were derived from dif-
ferent, yet fairly specific palaeobotanical communities" (p. 349) "and are not
merely the result of varied mechanisms or conditions of post-depositional al-
teration upon similar organic source materials" (p. 363 ). Figure 17 shows ear-
lier results of Johns et al. (1981) : compared to the dark layers, the pale and
light layers of the Latrobe Valley brown coal are maximal in hydrocarbon po-
tential, H:C ratio, extracts of aromatic hydrocarbons, triterpenoids, carboxylic
acids and aliphatic carbon, whereas higher amounts of aromatic carbon, phen-
ols and methoxyl groups were found in the dark layers. Thus, the light layers
show a remarkable similarity with type II kerogen of oil shales. These findings
are not consistent with the concept of light layers representing oxidation
products.
55

Structural Parameters as a function of lithot~pe


higher plant
wax derived
lignin ~

Pale
Light
M. Lt.
H. Dk.
Dark
Evidence IR.SS IR.SS IR.Py IR.SS IR.SS Py. Ex Ex El Py.RE
SS Fn Fn Pr

IR = infrared; Py = Pyrolysis/GC; SS = Solid State C13-NMR;


Fn = Functional group analyses; Ex = Extractable components;
El = Elemental analyses; RE = Rock-Eval; Pr = Proximate analyses.

Fig. 17. Chemicalcharacteristics of the variouslycolouredlithotypes in the Miocenebrown coal


of the LatrobeValley,Australia.From Johns et al. (1981). M.Lt. = mediumlight;M.Dk = medium
dark.

2.3.4 Origin of "light layers" in Tertiary brown coals


This topic deserves special treatment because many authors have discussed
it and very contradictory opinions exist. One would think that more strongly
humified material results in darker peat and darker brown coal than less hu-
miffed material, and, indeed, Figure 17 shows that the dark colour seems to be
the result of more lignin (methoxyl) and of more humic substances (phenols
and other aromatics). However, Koch (1969b) reported that Holocene reed
peats of Germany are darker in colour than forest peats, whereas Cohen ( 1973 ),
indeed, observed in the Okefenokee Swamp of Georgia that the (cellulose-
rich ) herbaceous peats tend to be lighter in colour than the (lignin-rich) forest
peats. The great difference in colour (when dried) between the older and
younger Holocene "moss peats" (Sphagnum peats) of northern Germany is
probably a consequence of climatic changes. The older "black peat" (with a
degree of humification, H = 6-9) has been deposited in a warmer and drier
climate than the younger "white peat' (H = 3-4). Both are separated by the
"Grenzhorizont" (ca. 700-1000 B.C.), a very dark and black layer with plant
remains from Ericaceae and Eriophorum which indicate a relatively dry peat
surface 2700 to 3000 years ago. Similarly, in North America, dark peats were
deposited in a dry environment (Cohen et al., 1987).
The discussion on the origin of darker and lighter brown-coal layers began
in early times when, on the basis of merely megascopic observations on the
dried coal face, WSlk (1935) distinguished in the Rhenish brown coal dark-
brown layers, stump horizons and light-brown layers which he could correlate
over wide distances (Plate III-7 ). In a "normal section" of eighty metres thick-
5~

ness, he ~'ound in the "Main Seam" (Hauptfl6z) 31 light layers that are 25-50
cm thick. They consist of a crumbling coal without visible plant remains, ex-
cept some fusinite and the upper part of stumps reaching from the underlying
stump horizons into the light layers. WSlk interpreted these light layers as
highly decomposed peats of the same vegetation that formed the dark layers.
Because he knew that oxidized peats generally are darker in colour he ex-
plained the colour of the light layers due to their higher content of bituminous
substances.
WSlk was obviously influenced by ideas of Gothan (1924) who regarded the
pale layers of German Eocene brown coals as oxidized peats, although, later,
Gothan (1937, p. 131) admitted that their formation might have been influ-
enced "by the kind of vegetation". In contrast to the light layers of the Miocene
German brown coals which differ from the dark layers by a merely slightly
lighter brown colour, when dried, the Eocene brown coals of Germany contain
much lighter layers that may attain a yellow tint. For a better differentiation,
these very light layers should be called "pale layers". Technologically, these
pale layers belong to the "Schwelkohlen" which were mined to produce montan
wax and low-temperature (500~C) tar. They are distinguished by high H:C
ratios, high yields of extract ! montan wax) and of tar and relatively high ash
contents. Under the microscope, these pale layers show a groundmass of bi-
tuminite in which are embedded high amounts of other liptinites, especially
sporinite, resinite (also wax), cutinite, suberinite and liptodetrinite, besides
some sclerotinite and inertodetrinite. Huminites are very rare and represented
almost only by ingrown roots (Plate III-8). Sponge spicules and high propor-
tions of clastic minerals indicate a subaquatic deposition.
According to Pflug ( 1957 ) the pale layers contain their own pollen spectrum.
The typical pollen, although very delicate, are well preserved. Laterally and
vertically the pale layers may change over into clay or silt bands. They have
been interpreted by Teichmtiller ( 1950, pp. 463,466-467 ), Pflug ( 1957 ), Koch
( 1969a, 1970a), Klein-Reesink et al. (1982) and Minnigerode and Riegel (1983)
as bituminous gyttjae, whereas Winkler (1986) regards these layers as dry,
oligotrophic forest peats (see below). Klein-Reesink et al. (1982) and Minni-
gerode and Riegel (1983) stated that colour and lightness (Helligkeit) corre-
late closely with the microscopic composition of lithotypes in Eocene brown
coals of Hessen. In the Borken coal, the pale layers are distinguished by a
"strongly fluorescing groundmass" (=bituminite), by high yields of extract-
able bitumen and large proportions of anemophilous pollen which are thought
to have been blown in. Klein-Reesink et al. (1982) found that the increase in
the proportion of bituminite in the pale layers correlates with a shift from the
"coal band" in the H:C/O:C diagram of Van Krevelen (1961) into the band
for organic matter type II (for example, of the Posidonia oil shale). The bitu-
minite groundmass is interpreted as of algal origin and the pale layers as limnic
deposits. On the other hand, the dark layers which are distinguished by a highly
57

gelified humic groundmass (densinite) and lack of well preserved tissues (hu-
motelinite ) were interpreted as the product of aerobic decomposition in a rel-
atively dry environment (Minnigerode and Riegel 1983, p. 315 ).
Winkler (1986) found that the pale layers of the Helmstedt Eocene brown
coal yield three times more extract than the dark layers, and that asphaltenes
and resins are the main components of the extracts. Although the increase of
n-C25-alkanes, fluctuations of homohopane yields, and the higher degree of
humification in the dark layers are considered to indicate a strong microbial
activity in the peat, Winkler interpreted the light layers as degradation prod-
ucts of the same peats that formed the dark layers.
The Eocene brown coals of Germany were deposited in a tropical climate
and, generally, reveal a much stronger decomposition of plant remains than
the Miocene brown coals of Germany which have been deposited in a sub-
t~opical to warm-temperate climate. Moreover, the Miocene brown coals con-
tain many plant remains that can still be correlated easily with modern plants,
living today in the southern United States or in southeastern Asia.
Whereas the pale layers of the Eocene brown coals are interpreted at present
as bituminous gyttjae by the great majority of coal petrologists, the origin of
the light layers in Miocene brown coals of Germany is still highly controversial.
The theory of oxidized peats (W51k, 1935) had been rejected for 25 years by
palaeobotanists and palynologists (such as Jurasky, 1940; Thomson, 1950,1951;
Hunger, 1953; Pflug, 1957 and others) in favour of the concept of a reed-marsh
peat or even of a subaquatic organic mud deposited in open lakes. TeichmiJller
and Thomson (1958) argued that the extremely high content of humodetrinite
(> 90 percent) (Plate III-3) was caused by the high cellulose and low lignin
content of the herbaceous reed vegetation, and that the amount of (well pre-
served!) pollen of the Quercoidites henrici and microhenrici type ( 10-times more
pollen than in the dark layers) came from extrapaludal forests, just as in the
pollen-rich gyttjae of the Rhenish brown coal (Plate III-5). The stumps and
the reworked material at the base of the light layers were regarded as indicators
of an inundation of the underlying forest peat and the initial cause for the reed-
marsh vegetation.
First objections came from Benda (1960) and Kilpper (1960) who both ap-
plied cuticular analysis to Rhenish brown coal. They found only sparse cuticles
of marsh plants in the light layers. Instead, well preserved monocotyledonous
cuticles were found in thin, very dark layers, distinguished by a high degree of
gelification, a pronounced stratification and, in part, by an appearance like
that of sapropelic coals. These dark layers contain plankton and allochthonous
pollen from a forest vegetation rich in conifers. Benda and Kilpper interpreted
these layers as representative of the reed-marsh facies in the Rhenish brown
coal. However, they did not take into account that ( 1 ) concentrations of well-
preserved leaves in finely laminated, dark and highly gelified brown coals are
characteristic of humic gyttjae (Teichmiiller, 1950; Schneider, 1980; Klein-
5~

Reesink and Riegel, 1983): whereas (2) reed-marsh peats like, for example,
~hose of the Okefenokee Swamp of Georgia (Cohen, 1973 ) and the sedge-grass
peats of the Everglades in Florida (Cohen and Spackman, 1977 ) do not contain
well preserved leaves. Instead, the surface litter of the reed plants breaks down
very rapidly, so that a fine granular material occurs just a few inches below the
surface. According to Cohen and Spackman (1977) the ratio between "f'rame-
work" (plant tissues) and "matrix" (detritus) is very low in marsh peats. Like-
wise, Koch (1969b, p.342) describes only root tissues from Holocene peats
deposited in marshes of Cyperaceae and Gramineae, and Styan and Bustin
(1983a) found high amounts of "amorphous" humic matter (humodetrinite),
together with high cellulose/tignin ratios in sedge-grass peats of the Frazer
River delta. Schneider (1980) distinguished between "sedentary" (autochtho-
nous) and "sedimentary" (allochthonous) Glumiflorae peats as facies of the
Lusatian brown coal, the first consisting of dense flat (German: Filz) from
unidentifiable plant remains in which only roots are preserved as tissues,
whereas the allochthonous Glumiflorae brown coals bear well preserved leaf
tissues of Glumiflorae and are distinguished by a dark black colour, high degree
of geliftcation, and pronounced stratification, i.e., the same properties that
Benda (1960) and Kilpper (1960) described for their "reed marsh facies" in
the Rhenish brown coal. Also in modern peats, Cohen et al. (1987) found ge-
lifted masses in deeper subaquatic habitats. The same authors point to the
allochthonous origin of finely layered peats. These examples may suffice to
show that, just as with the palynological results (see p. 50), the results of cu-
ticular analyses must be evaluated with caution when reconstructing vegeta-
tion and peat types of various coal facies.
Weyland (1958), Kilpper (1960), Benda (1960) and, partly, also Hittmann
(1976) assumed that the light layers of the Rhenish brown coal are the result
of large inundations leading to redeposition of detrital peat particles from the
underlying autochthonous peat layers. Jacob (1968) likewise interpreted the
light layers as peats deposited in a wet environment under oxygenated water.
A completely different origin is, however, proposed by Von der Brelie and Wolf
(1981a) and Hagemann and Wolf (1987). These authors returned to the idea
of WSlk (1935) who regarded the light layers as oxidized forest peats (see pp.
55 and 56 ). Hagemann and Wolf ( 1987 ) stated that "in coals with a high amount
of Quercoidites microhenrici pollen" (as those regarded as reed coals by Teich-
mfiller and Thomson, 1958) "the proportion of fine detrital groundmass in-
creases, a fact which points to drier conditions" (p.344) and "never was an
enrichment of the easy scattering Pinus pollen observed in the pale bands of
the so-called Reed-facies (i.e., open water facies)" (p.345). Another argument
is based on geochemical results of Hagemann and Hollerbach (1980): "assum-
ing that high amounts of homohopanes are equivalent to intense microbial
activity and low oxygen concentrations,..., the unbanded light bands show evi-
dence of aerobic decomposition" (p.342). Homohopanes are oxidation prod-
59

ucts of bacterial membranes: their role as indicators of redox potential is un-


known (Dr. J. RullkStter, KFA Jiilich; pets. commun., 1988). The lack of Pinus
pollen in the reed coals of Teichmfiller and Thomson (1958) is easily explained
by the lack of a corresponding flora and/or by unfavourable wind directions,
particularly as Von der Brelie and Wolf (1981a, p.137 and 1981b, p.189) stated
that Pinaceae forests were rare during the deposition of the Rhenish brown
coal. According to Thomson (1950/1951) extrapaludal forests with oak-like
Cupuliferae, rather than with Pinaceae, contributed to the high amount of
allochthonous pollen in the open-water facies. Moreover, the very high pollen
content of the "reed coals" and the excellent preservation of the pollen exines
are incompatible with a strong aerobic decomposition of the peat. The same
applies to Thomson's (1950/1951) finding that the pollen spectra and fre-
quencies in the light layers are the same as those in lacustrine layers of the
same coal, that cuticles and the few parenchymous tissues of roots are well
preserved, and that reworked peat and rounded pieces of wood occur at the
bottom of the light layers.
In the light layers of Australian brown coals, the Nothofagus pollen plays
the same role as does the Quercoiditespollen in the Cologne brown coal. This
pollen is highly enriched in the light layers and is thought to be derived from
an extrapaludal vegetation. As reported in the foregoing chapter (pp. 53,54)
the very well founded results of Australian palaeobotanists (for megafossils as
well as for sporomorphs) led to the conclusion that the pale and light layers of
the Miocene Yallourn coal represent subaquatic deposits (cf. Fig. 16). Ker-
shaw et al. (1988) drew the conclusion "the fact that there is such a good
relationship between pollen composition and petrology suggests that altera-
tion (in the sense of peat oxidation according to Von der Brelie and Wolf 1981 )
has not been an important factor".
The fundamental idea of Von der Brelie and Wolf (1981a,b) and of Hage-
mann and Wolf (1987) is that the ratio humodetrinite:humotelinite is the
most important indicator for dry/wet conditions of peat formation. They did
not consider the different resistance of different plant materials, especially of
cellulose and lignin which play a decisive role for the ratio detritus:telinite.
This has also been shown by experiments of Masran and Pocock (1981) who
subjected specific plant tissues to physical, chemical and biological breakdown.
As a matter of fact, the preservation of plant tissues in peat depends on:
(1) the resistance of the plant material (coniferous: angiospermous wood, her-
baceous plants, more or less lignified);
(2)autochthonous or allochthonous deposition (transport by wind, water, with
or without minerals);
(3)redox potential (which depends on dry/wet conditions and on water flow):
(4) acidity (Sphagnum peats contain, for example, well-preserved tissues due
to acids produced by the Sphagnum mosses);
(5)peat temperature (in tropical peats microbial life is more intense than in
peats of cool climates).
Finally, it should be mentioned that Thomson (1954) described thin, rela-
tively light bands occurring within the dark forest coals of the Fortuna Mine.
These light bands differ distinctly from the light reed coals in their petro-
graphic as well as their palynological properties. They bear Sphagnum spores
as well as pollen of Ericaceae and Pinus and well preserved coniferous wood.
These light bands are interpreted as ombrogenous peat layers and may be due
to climatic changes.

2.4 Verticalsuccession of coal facies and the influence of nutrient supply


(eutrophic-oligotrophic)

This subject has become a favoured one in recent times, especially in regard
to "ombrogenous coal". One of the first authors who studied the vertical se-
quence of lithotypes and microlithotypes in many seams of the Ruhr Carbon-
iferous was E. Hoffmann (1933). Carboniferous seams of the northern hemi-
sphere generally have a seat earth with Stigmaria, the roots of lycopod trees.
According to Hoffmann ( 1933 ) the lower part of the seam is rich in vitrite and
liptinite-poor clarite. The vitrite layers are thick and form megascopic bands
of bright coal, often representing single thick stems of lycopods. Dominating
spores in the clarites were derived from Lepidodendron, Lepidophloios and other
lycopod trees (Scott and King, 1981). Upwards, the vitrite content decreases
and the predominance of bright coal is followed by an increase in semi-dull
and dull layers consistiag of liptinite-rich clarites and durites. In the upper-
most part of the seam, the dull coal may contain more and more clay minerals,
ending in a dirt band. If sapropelic coals occur, they usually lie at or near the
top of the seam. But in many cases the amount of bright coal increases again
just below the top: vitrite and liptinite-poor clarite, associated with carbargilite
represent the end of seam formation. This overall sequence, described by Hoff-
mann (1933) (which may be interrupted by dirt bands) has been interpreted
as a development from drier (forests) to wetter (marshes, open-water) envi-
ronments (see Fig. 8).
It is to the great merit of Smith (1962,1968) (who first combined coal-pet-
rological with palynological methods in studying the sequence of facies in Eng-
lish Carboniferous coals) that he drew attention to a possible development
from initially eutrophic to eventually oligotrophic conditions of peat formation
in Carboniferous coals. Figure 18 shows the spore "phases" and the corre-
sponding microlithotypes within a seam according to Smith (1968). The seam
begins with the "lycospore phase" with vitrites and clarites, followed by a
"transition phase" with trimacerites, followed by the "densospore phase" with
durites rich in spores and in macrinite ("massive micrinite"). This sequence
is interpreted as the development from a wet, eutrophic forest swamp with
61

lycopod trees, through a less wet mire with a more open vegetation rich in
species (sphenopsids, ferns, pteridosperms, and herbaceous lycopods ) to a rel-
atively dry, oligotrophic high moor (raised bog) with an unknown, stunted
vegetation. The characteristic microspore of the densospore phase, namely
Densosporites sphaerotriangularis may have been derived from a lycophyte tree.
Nemejc (1931) and Chaloner (1955) have recovered megaspores of the type
usually found with densospores from species of Sporangiostrobus. Recently,
Wagner (1985) has isolated microspores of the Densosporites type from a large
species of Sporangiostrobus with sigillarian affinites.
A reversal of the plant successions after the densospore phase is thought to
be due to subsidence. The "incursion phase" may be present anywhere within
Phases Petroqraph/
(llcrolithotypes)
Roof

Lycospore Vitrite
and
Clarite

Transition Duroclarite
Clarodurite

Ourite
Densospore (Inertinite mainly massive
micrlnite)

Transition Clarodurite
Duroclarite

VItrite
Lycospore and
Clarite

Seat-earth

Fig. 18. Microspore phases and the related microlithotypes in Carboniferous coal seams from
Yorkshire, England. From Smith (1968).
62

the seam, usually combined with spores either of the "lycospore phase" or the
"'transition phase". It represents temporary flooding of the peat and is char-
acterized by mineral sediment and by thin spores ("tenuispores"), often
weathered, and by assemblages rich in species and partly of allochthonous or-
igin. According to Smith (1962,1968) "crassidurite" (the raised-bog facies) is
distinguished by a low ash content and lack of' pyrite. It occurs preferentially
in relatively thick seams in which peat formation was not interrupted, and
nutrient deficiency could develop. Interestingly, Westphalian C coals do not
contain these crassidurites; according to Smith, this was probably a conse-
quence of a drier climate.
Littke (1985,1987) studied 20 seams of the Ruhr Carboniferous (of' which
10 seams were of Westphalian C age!) using coal petrological (only one seam
also palynological) methods. Four seams showed a sequence similar to that
described by Smith from English Carboniferous coals, and only six seams con-
tained more than 5% macrinite, the maceral which is regarded to be charac-
teristic of the oligotrophic phase. Likewise, Hacquebard and Donaldson ( 1969 )
did not find crassidurite in Canadian Carboniferous coals that are of West-
phalian C and D age. According to detailed coal petrological and palynological
studies in the thick Harbour seam of Nova Scotia, the vertical sequence of
facies is the same as described from Ruhr coals by Hoffmann (1933) (see
above): in the lower part bright coal with vitrites and clarites thought to be
deposited in wet forest swamps (more seldom in reed marshes) predominate,
whereas the upper part of the seam contains more dull layers with spore-rich
clarites, durites and clay partings which were deposited in reed marshes and
open water, respectively. A certain coal band with much degradofusinite and
macrinite is interpreted as a period of dessication, although without assuming
oligotrophic conditions.
On the other hand, recently Bartram (1987), on the basis of the succession
of lycopod megaspores, distinguished in one British Westphalian B coal five
phases that are broadly related to petrographic types, although most changes
in spore phases and petrography are not coincident. At the base of the seam,
carbominerites with vitrites and clarites, containing megaspores of Lepidod-
endron and Lepidophloios predominate. They are interpreted as peat deposits
below the water table in a nutrient-rich environment. Phase 2, with vitrite,
clarite, and cutinite-rich clarite, bears the megaspores (Triangulatissporites,
Valvisisporites) of a herbaceous lycopod vegetation (SellagineUites, Chalo-
neria) and is thought to be nutrient-poor. Phase 3, with trimacerite (i.e. with
increasing amounts of inertinite), is assumed to be formed just at the ground
water table; and phase 4, with durite and trimacerite, containing Zonalesspor-
ires, is interpreted as a peat formed in a raised bog under ombrogenous condi-
tions. Megaspores of the Zonalessporites type occur with the microspores of
the Densosporites type in heterosporous cones of Sporangiostrobus type (Wag-
63

ner, 1985). The association of Zonalessporites brasserti and Densosporites


sphaerotriangularis was noted by Bartram (1987).
Smyth and Cook (1976) also tended to acknowledge Smith's ideas of a peat
cycle for Australian Gondwana coals, although there are, of course, marked
differences in the petrographic composition between Euramerican Carbonifer-
ous and Australian Gondwana coals (see p. 44,45). On the basis of statistical
methods, Smyth and Cook (1976) found that the vitrite + clarite content gen-
erally decreases upwards in a seam (or in a cycle within a seam) and that a
vitrite + clarite-rich coal is the most likely lithotype after a dirt band. Crassi-
durites are unknown in Australian coals, just as most other liptinite-rich mi-
crolithotypes. The inertinite-rich layers at the top of a cycle are very poor in
liptinites but rich in mineral matter. The latter does not agree with the concept
of oligotrophic peats deposited in a raised bog, although Smyth and Cook (1976)
explain the high amount of mineral matter with a remani~ effect caused by
oxidation.
Cecil et al. (1985) assumed an influence of climatic changes during the Car-
boniferous on the occurrence of ombrogenous, oligotrophic coals. For the latest
Mississippian and earliest Pennsylvanian (Namurian) they assumed an ever-
wet tropical climate in which many peats of Appalachian coals were deposited
in domed bogs with forests. With change of this climate to "seasonal tropical"
in later Pennsylvanian times, the coal-forming peats are assumed to have been
deposited in planar areas between river channels. These suggestions are based
on the spatial form of coal seams, ash and sulfur contents, kind of minerals,
seatrocks and lithology of other accompanying rocks in the central Appala-
chian basin. Coal petrological studies have not been carried out. Another ex-
ample of an "ombrogenous coal" deposited in the tropical climate of the Car-
boniferous has been given by Esterle and Ferm (1986) who based their
conclusion on the sequence of lithotypes, their maceral composition and their
ash and sulphur contents in relation to the geometry of the Hance seam in
Kentucky. The authors found that non-split portions of the seam in the inte-
rior of the basin are relatively poor in vitrinite, ash and sulfur, whereas the
thinned and split marginal seam is richer in these constituents. The interpre-
tation is that ombrogenous domed peats with a stunted vegetation formed in
the interior and that peats from larger trees grown in a nutrient-rich environ-
ment were deposited in the marginal parts of the basin (Fig. 19).
In the book "Coal and Coal Bearing Strata: Recent Advances", edited by
A.C. Scott (1987), many authors (most of them are not coal petrologists)
stressed the idea of oligotrophic, ombrogenous peats as precursors of coal seams,
especially in the Carboniferous (Moore, Clymo, McCabe, Fulton, Bartram).
The ombrogenous forest bogs of Indonesia (Anderson, 1964,1983; Teichmiiller
and Teichmiiller, 1982, pp.9,14,15,30) are often compared with the deposi-
tional environment in which Carboniferous coals have been deposited. Only
WELLPRESERVEDWOOD
...BSZ V
] MIXEDMATERIAL
65-84~[ V
] POORLYPRESERVEDSHRUBS
•C6SZ V
] FIRECLAY

] SANDSTONE

Fig. 19. Presumed moor types of the lower bench of the Upper Hance seam, Kentucky (U.S.A.),
with an ombrogenous domed bog in the centre. From Esterle and Ferm (1986).

Bartram based her conclusions on new coal petrological and palynological data
( see pp. 62,63 ).
According to Fulton (1987), Anderson (1964) differentiated freshwater
"rheotrophic" ( = topogenous) peats with ash contents of > 25 and pH-values
of > 4 from "ombrotrophic" ( = ombrogenous ) peats with < 25% ash and pH-
values of < 4. This would imply that most mineable coal seams have been de-
posited in domed bogs under ombrogenous, oligotrophic and relatively dry con-
ditions. This concept, however, cannot be accepted unreservedly by coal geol-
ogists and coal petrologists because the development from topogenic to
ombrogenic requires that peat formation exceeds subsidence which is possible
only under stable geological conditions. However, such conditions are lacking
in foredeeps and other unstable areas where most coals have been deposited.
Foredeep coals are rich in vitrites and liptinite-poor clarites (Smith, 1962),
often associated with clastic layers and with pyrite inclusions, particularly in
the numerous thin, often unmineable seams of the Hercynian foredeep. In most
cases, periods of stability were apparently too short for a bog surface to build
up high enough above the groundwater table. Consequently, the oligotrophic
"densospore phase" known from English Carboniferous coals (Smith,
1962,1968) could develop only occasionally. Nevertheless, at least ten seams
have substantial beds of crassidurite in the Westphalian A and B of Yorkshire
(pers. commun, of Dr. A.H.V. Smith, Sheffield University).
According to Grosse-Brauckmann (1979b) who investigated many modern
peats in the temperate-climatic zone, "the successional direction from + eu-
trophic to + oligotrophic peat types seems to be an endogenous tendency of all
peat-forming vegetations, but the real process of such successions is highly
65

controlled by environmental, especially by water conditions". Many plant


communities do not change during the span of many hundred years, due to
constant conditions of water supply. These important observations are in ac-
cordance with Schlesinger (1978) who found that in wet swamps with pro-
longed periods of standing water the nutrient cycling is confined to the water
column with the peat acting as a permanent nutrient sink. This is the case in
Taxodium swamps of the Okefenokee Swamp (Georgia) and very probably was
the case in lycopod forest swamps of the Carboniferous. Phillips and Di-
Michele (1981), for example, assumed that Lepidophloios hallii, the dominant
tree in the swamps of the Herrin No. 6 seam in Illinois (Sahara Mine) could
grow under minimal nutrient levels standing in relatively high water for pro-
longed periods. Although sigillarians are thought to be indicative of relatively
low water levels, and Phillips and DiMichele (1978) reported that they tend
to be associated with relatively high amounts of fusinite, sigillarians are ac-
companied by the "blue band" in the Herrin No. 6 seam, i.e., a distinctive
parting of claystone. Likewise, in the Ruhr Carboniferous, near Essen, a sub-
mersed, petrified sigillarian forest with calamites in the undergrowth has been
mapped and reconstructed (Fig. 20). The trunks, preserved in cast molds up
to a height of 7.5 m, were standing on a thin layer of bony coal with sigillarian
spores and leaves from ferns and pteridosperms. It seems that at least in this
case the sigillarians lived in an open forest standing in high water.
The central part of the Everglades in Florida (known as the "slough") ob-
tains a steady nutrient supply through a very slow flow (30 m / h r ) of water
moving from the Lake Okeechobee in the north, southwestwards through the
reed marsh. Similar conditions may have ruled in palaeomires where topogenic
coals poor in mineral matter and ash have been deposited, for example, in the
mires of the Rhenish brown coal (Teichmiiller and Teichmfiller, 1982, p.31 ).
The assumption of Von der Brelie and Wolf (1981a) that the whole Rhenish
brown coal is of "more or less ombrogenous origin" is not argued. For a seam
with a maximum thickness of 100 m (HauptflSz) this would imply that oligo-
trophic conditions must have prevailed for more than 200,000 years (Teich-
miiller and Teichmiiller, 1982, pp.17-18) or according to Zagwijn and Hager
(1987) for even nine million years! This does not seem probable, bearing in
mind the steady, slow subsidence of the coal-bearing area in a graben zone
during the Miocene. Thomson (1954) interpreted thin light bands with Sphag-
num and Ericaceae pollen as ombrogenous types (see p. 60). It is not clear if
the vertical development from more darker to more lighter lithotypes in the
Australian Morwell No. 1 brown coal that Mackay et al. (1985) interpreted as
a development from wet to drier environments of peat formation (see p. 54) is
assumed to represent also a development from more eutrophic to more oligo-
trophic conditions.
Recently, Kosters et al. (1987) and Cohen et al. (1987) drew attention to
the possibility of a secondary leaching out of mineral matter from peat. Sili-
66

Fig. 20. Lycopod (mostly sigillarian) forest swamp with calamites in the undergrowth. Recon-
struction on the basis of cast molds of upright stems mapped in the WestphalianA of a quarry
near Essen/Ruhr. See Klusemannand Teichmiiller (1954).

ceous phytoliths and sponge spicules may disappear completely and detrital
mineral matter may be dissolved too by water circulating in peats. Thus, coals
poor in ash may not necessarily be of ombrogenous origin.
Undoubtedly, the problem concerning the recognition, properties and the
significance of coals t h a t have been deposited under oligotrophic and ombro-
genous conditions remains one of the most interesting to be solved in the future.
67

2.5 Influence of acidity and of marine and calcareous conditions

Acidity has a great influence on the bacterial activity in peats. Most bacteria
thrive best under neutral to weakly alkaline conditions (pH 7-7.5). Because
most peats are acid (due to the formation of humic acids), bacterial life in
peats is much less intensive than in "normal" soils. Sphagnum peats have
especially low pH values (3.3-4.6), due to: (1) certain acid secretions of the
Sphagnum plant; (2) to more oxidative conditions of peat formation in a raised
bog; and (3) to higher concentrations of humic acids because water exchange
is lacking. Moreover, the substrata of these peats are commonly represented
by other peat types or sandy (acid) subsoils. The peat of raised forest bogs in
Indonesia is similarly distinguished by low pH values (3.5-4.5), but, whereas
Sphagnum peats of the temperate climate contain excellent preserved plant
tissues, most ombrogenous forest peats of the tropics are strongly decomposed,
due to the high peat temperatures that favour fungal and bacterial decompo-
sition. This should be considered when interpreting Carboniferous coals of
Euramerica as of ombrogenous and oligotrophic origin.
Marine influences on coal facies have been discussed for many years and in
detail (Teichmiiller and Teichmiiller, 1982, pp. 27-30). It is well known that
coals that are covered by marine clastic rocks are rich in sulfur-organic sulfur
as well as pyrite. Only part of the sulfur stems from the peat-forming plants,
most of it comes from SO4-ions of the seawater. According to Neavel (1981)
organic sulfur as well as pyrite owe their emplacement in the seam to the ac-
tivity of sulfur-reducing bacteria. Desulfovibria desulfuricans reduces sulfate
to HeS which is needed for pyrite formation. The iron most likely enters the
swamp adsorbed on clays. Therefore, pyrite is often found adjacent to clay-rich
zones which, for their part, are mostly associated with vitrites and clarites.
According to Westgate and Anderson ( 1984 ), organic sulfur is enriched in 34S
unlike pyritic sulfur.
Recent studies of various peats in the southern United States and in British
Columbia carried out by Cohen et al. (1984), Given and Miller (1985) and
Styan and Bustin (1983a) confirmed that brackish and marine peats are en-
riched in organic and pyritic sulfur. Of course, mangrove peats are extremely
rich in sulfur. According to Cohen et al. (1984), burial of freshwater peat be-
neath marine or brackish peat also leads to an increase of sulfur in the fresh-
water peat, particularly if clay-rich deposits overlie it. Given and Miller (1985)
found total sulfur contents of up to 6% in dried peats of coastal swamps in
Florida. They report that sulfate-reducing bacteria prefer fatty acids of chain
C3 and up, and grow in association with the anaerobic bacterium Clostridium
which can ferment cellulose, thereby converting it to lower fatty acids. Casa-
grande et al. (1977) reported that sulfate-reducing bacteria grow best at high
pH values (6.5-8). In the peat from the Frazer River delta the abundance of
H2S generally correlates with these pH values (Styan and Bustin, 1983a):
t3~

whereas Sphagnum peats have the lowest S-contents (0.22%) brackish peats
are rich in sulfhr (2.9%) and reach maxima of 5.6% S, 70~; of it as organic
sulfur.
Marine-influenced coals are not only enriched in sulfur. The Katharina seam
of the Ruhr Carboniferous, which is overlain by a dark marine or brackish
claystone, is also distinguished by relatively high amounts of hydrogen, nitro-
gen and volatile matter. The characteristic microlithotype is a liptinite-poor,
micrinite-bearing clarite with a perhydrous collinitic groundmass (Teich-
mfiller, 1955). Alteb~iumer (1983) found in the 10 uppermost centimetres of
the Katharina seam, which was studied at 15 different localities, high amounts
of micrinite-bearing bituminite (probably of bacterial origin) as well as abun-
dant exsudatinite and "oil expulsions" in the strongly fluorescing vitrinites,
all indications of intense oil generation from the perhydrous coal which, cor-
respondingly, yields high amounts of extract. The well-known anomalous cok-
ing properties of the Katharina seam (high fluidity, strong swelling) are a
consequence of its high bitumen content. It seems that marine-influenced coals
are especially suited to be oil source rocks (Teichmfiller, 1987, p.143).
Like marine-influenced coals, coals deposited in a calcareous, non-marine
environment show similar petrographic and chemical properties. They are rich
in sulfur (organic sulfur and pyritic), their vitrinites tend to be perhydrous
and their coking behaviour is anomalous and distinguished by low softening
temperatures and high swelling indices. Calcareous solutions from the floor or
the roof, or from partings within the seams cause neutral or alkaline conditions
for a strong bacterial activity which resulted in the formation of fatty acids
(see above) and of other finely divided lipoid substances within the humic
groundmass.

2.6 Influence of fires

According to Cohen (1974) and Cohen et al. (1987) severe fires in the swamps
of Florida and Georgia may cause the death of the peat-forming vegetation.
During such fires the uppermost part of the peat may burn too. In these cases
the whole peat-forming environment changes suddenly from, for example, a
dense forest swamp to an open-lake or a reed marsh. Plate IV-4 shows such a
situation in the Okefenokee Swamp of Georgia. Such an event would be marked
in the case of hard coal by the appearance of a layer of fusain, underlain by a
bright coal rich in vitrite and clarite {forest peat) and overlain by a dull coal
with liptinite-rich clarites and durites (subaquatic muds or marsh peats) or by
a clastic band. Accordingly, Phillips and DiMichele ( 1981 ) found in the Herrin
No. 6. coal of Illinois the maximum fusain zone just below a shale parting, and
Scott and Collinson (1978) refer to fusain concentrations in English Carbon-
iferous coals that are overlain by lake deposits.
Styan and Bustin (1983a) observed a rapid increase of pyrofusinite (and
69

degradofusinite) in peat types of the Frazer River delta which were deposited
marginal to levees. Small ponds on the peat surface resulted from periodic fires
that burned into the peat and formed depressions (Bustin et al., 1985). Char-
ring of peat was observed after a very dry summer (1959) in a mire in the
Netherlands (Teichmiiller, 1961). Although the fire had been extinguished,
charring was going on locally down to 50 cm below the burned peat surface,
and at some places less than 1 m away from small open-water sinks. During
charring, peat-coke and pyrofusinite did form, the latter with excellent pres-
ervation of cell tissues from roots of living plants. Plate I-9 shows a fusinized
root tissue of a birch that still existed at the surface but was totally dried. The
observations suggest that relatively large pyrofusinites with especially well-
preserved cell tissues represent charred roots. The recent pyrofusinites that
had formed above the peat surface were smaller, fragile and often like fine dust
in size. They are easily blown away by the wind, just like the very fine ash that
formed during the surface burn. Therefore, small particles of fusinite, known
as inertodetrinite, are found in almost all sedimentary rocks where they have
been blown in or washed in. It is surprising that layers or patches of peat coke
seem to be practically unknown from coals. The charred peat is very fragile
because of an enormous porosity (Plate I-8) and, therefore, obviously only
small particles of it remain preserved in coals, mainly as macrinite and
inertodetrinite.
The influence of fires on the succession of plant associations and of litho-
types in Australian brown coals has been stressed by Blackburn (1981) as well
as by Kershaw et al. (1988). Fusinite is enriched in the dark layers but may be
washed into the open-water muds of the light layers as well. The maximum of
fusinite content was found in the darkest layer of the Yallourn brown coal
which is interpreted as a raised-bog peat.

3 CONCLUSIONSAND FURTHERWORK

Since the era of descriptive coal petrography is more or less over, modern
coal petrologists attempt to elucidate the origin of the microscopic entities
("macerals") of coal, by working in cooperation with palaeobotanists and or-
ganic geochemists. Some macerals still show their origin from certain plant
parts ("phyterals"); others do not due to a severe decomposition before or
during the peat stage. Therefore, microscopic studies of peat constituents are
very important in understanding the origin of macerals, which depends on the
source material as well as on its fate during peatification and coalification.
Over the last 15 years coal petrology has greatly profited from results ob-
tained through the microscopic investigation of oil-source rocks and has broad-
ened to "organic petrology" which deals with the microscopic investigation of
organic constituents in all rocks. It is mainly on the basis of organic petrology,
in combination with organic geochemistry, that new insights have been ob-
7(}

tained into the genesis and the coalification/maturation of" macerals. One im-
portant result is the distinction between primary and seconda©' macerals, the
latter generating through disproportional reactions within primary macerals
during coalification/maturation, especially as a result of bituminization. Bi-
tuminization is part of the coalification process and is comparable with the
generation of petroleum in clastic oil-source rocks.
Coal as a source rock [or oil t not only for gas ) has become an important new
subject for coal petrologists (as well as for hydrocarbon exploration! ). It has
been found that not only the H-rich liptinites but also perhydrous vitrinites
generate bitumen, even earlier than many liptinites. These findings were pos-
sible mainly by the use of' fluorescence microscopy. Close relationships were
found between vitrinite fluorescence and coking behaviour, and, in more recent
times, also between the fluorescence of some inertinites and their coking ability.
Microscopic studies of modern peats, partly combined with chemical and
microbiological studies, contributed to the understanding of the genesis and
early fate of huminites/vitrinites which formed from the tissues of higher plants.
Humification takes place in the peat and brown-coal stages. It is a microbial
and chemical process. Although cellulose is much more easily attacked by mi-
croorganisms than lignin, there remain cellulose-rich tissues preserved up to
the brown-coal stage. The question of whether their later coalification prod-
ucts are vitrinites or fusinites is not yet solved. Coalification experiments un-
der water and at elevated temperatures and pressures have shown that cellu-
lose generates bitumen very early, possibly leaving behind secondary "rank
fusinite". The causes of biochemicalgelification in peats and brown coals which
leads to the formation of various kinds of gelinite remain doubtful, although it
has been learnt that this process is promoted in subaquatic muds.
Electron microscopical studies allowed the recognition of tiny lipoid parti-
cles in huminites/vitrinites. These particles are not visible under the light mi-
croscope, but they obviously cause the relatively low reflectance and high flu-
orescence of perhydrous vitrinites. These lipoid inclusions were found to be
associated with indications of fungal and bacterial activity around huminites/
vitrinites.
Geochemicalgelification converts huminites of the peat and brown-coal stages
into the vitrinites of hard coals. It causes the most striking petrographic changes
during the whole coalification process and is obviously related to the early
generation of bitumen (beginning of the "oil window"). The hydrogel brown
coal is converted into the bitumogel hard coal.
For the liptinite maceral group, fluorescence microscopy allowed new results.
This method showed that many liptinites cannot be properly classified when
using the Stopes-Heerlen system of maceral classification. Some new liptinite
macerals were detected by the use of fluorescence microscopy but there remain
other liptinitic constituents (distinguished by a low reflectance and a high
fluorescence ) whose classification and origin remain doubtful. Although there
71

is a tendency to classify all strongly yellow-fluorescing material as "alginite"


in analogy to the well-known yellow-fluorescing Botryococcus alginite of sap-
ropelic coals and to the many newly described algae of oil-source rocks, the
true algal nature of all these constituents has not yet been ascertained by pa-
laeobotanists. Resinites were found to vary widely in their fluorescence prop-
erties (5 to 6 different kinds have been described from the same coal). The
new liptinite maceral bituminite is amorphous. Its origin from anaerobic bac-
terial degradation of algal and faunal plankton, inclusive of much bacterial
substance, is assumed because of its occurrence in sapropelic coals and in mi-
crolithotypes that were obviously deposited under water. Fluorinite occurs with
phyllovitrinites, surrounded by cutinite and, therefore, has been interpreted as
former essential oils. Recent measurements of fluorescence alteration suggest
a very early bituminization of this new liptinite maceral. The newly found
liptinite maceral exsudatinite which represents secondary (probably asphal-
tenous) bitumen, filling cracks and other voids, testifies to the generation of
bitumen in coals.
During coalification all liptinite macerals pass through "coalification jumps"
that are caused by the generation of bitumen and its later cracking to gaseous
hydrocarbons. These jumps occur at different rank stages for the different lip-
tinites. Resinites, fluorinite and bituminite form bitumen very early, whereas
the bituminization of Botryococcus-alginite occurs relatively late.
It seems that in the future the genesis of inertinites will acquire a special
interest. Besides the well known pyro-inertinites, formed through charring of
plants or peat, the degrado-inertinites formed under access of oxygen through
mouldering with microbial, especially fungal, activity before or during the early
peat stage, and the primary inertinites, already existing or pre-existing in the
living plants as melanin-rich material, there are in particular the rank-inertin-
ites that obviously form during coalification. These rank-inertinites must be
regarded as secondary macerals and will become a very worthwhile subject for
research. A well-known example of secondary inertinite is micrinite. This is
the solid residue ("dead carbon") left after bituminization of liptinites like
bituminite as well as certain resinites and submicroscopical lipoid substances
hidden in the perhydrous vitrinites. To clarify the genesis of rank-inertinites
artificial coalification experiments (pyrolysis at elevated pressures) with var-
ious source materials (inclusive cellulose-rich and lignin-rich woods from brown
coals) are recommended. The assumed genesis of some degrado-inertinites
through fungal attack has been confirmed by recent observations in modern
peats. The genesis of macrinite whether from humic gels formed under water,
or from charred peat, or other causes, remains an open question.
Although many questions of maceral genesis remain open, a widely acknowl-
edged concept is that huminites/vitrinites indicate wet and anoxic conditions
of peat formation, whereas inertinites indicate more oxidative, and generally
also drier conditions. This concept seems to be too simple and has to be checked
72

from case to case. For instance, the well-known types of Gondwana coals that
are distinguished by extremely high amounts of inertinites seem to contain
submicroscopic algal remains according to recent studies under the electron
microscope. This would indicate deposition under water (and not dry, oxida-
tive conditions). Not only artificial coalification but also artificial peatifica-
tion experiments may lead to a better understanding of the origin of macerals,
especially of the amorphous types. An ideal example is the work of Masran and
Pocock (1981) who subjected tissues of specific plants to physical, chemical
and biological breakdown under varying laboratory conditions. Electron mi-
croscopical (TEM) studies as applied recently by Taylor and Liu (1987) cer-
tainly will become another method to clarify the (sub-light-)microscopical
composition of macerals and, thereby, possibly, their origin.
Lithotypes and microlithotypes, that is typical maceral associations, may be
regarded as the different coal facies. Modern coal petrology tends to draw con-
clusions from the type and the succession of coal facies to the palaeogeographic
position of peat formation, the peat-forming vegetation, the water and the nu-
trient supply, the peat acidity, the climate of peat formation and accidents like
crevasses and fires. For such studies the cooperation with sedimentotogists,
palaeobotanists and peat experts is necessary. Such studies have shown that
mineable coal seams have been deposited usually in upper delta plains or in
back-barrier areas of coastal plains, and less commonly in limnic environ-
ments along rivers and lakes. Following the pioneering work of Hacquebard
and Donaldson (1969), Australian coal petrologists, especially, tried to find
relationships between coal facies and sites of deposition such as lower and upper
deltaic, fluvial, lacustrine (Smyth 1984) or lower and upper deltaic and pied-
mont plain (Diessel, 1986), using indicators like the ratio vitrite+clarite :
durite + inertite, or a "Gelification Index" and a "Tissue Preservation Index",
respectively. Generally coals rich in vitrite and clarite are thought to have been
deposited in wet and more anoxic environments, for example of foredeeps,
whereas the inertites of Gondwana coals are assumed to have been deposited
in small basins on a stable craton or in piedmont plains. Fluviatile environ-
ments lead to coals rich in vitrites and clarites but also rich in mineral matter.
Lacustrine deposits are distinguished by microlayered clarites and durites,
usually with high liptinite contents and perhydrous collinites.
Vegetation, climate and water supply play a major role in the development of
the various coal facies. The literature of recent decades is full of papers treating
the genesis of lithotypes (of brown coals mainly) and microlithotypes (of hard
coals) in relation to their provenance from plant associations like forests, reeds
with sedges and grasses and submerged and not submerged water plants. Most
authors agree that vitrite and liptinite-poor clarites formed from an arbores-
cent vegetation, that liptinite-rich clarites represent organic muds deposited
in still-water lakes and ponds, and that inertinite-rich durites represent rela-
73

tivelydry depositional milieus. For liptinite-richdurites the opinions vary from


wet to relativelydry conditions.
Much disputed was the origin of the lighterand darker brown coallithotypes.
Although Tertiary brown coals contain the remains of plants which are similar
to plants living today, and although palynology as well as cuticular analyses
have been applied, many questions remain unresolved. The reasons are that
palynological results need a careful evaluation since anemophelous pollen are
always overrepresented, the resistance to decay of pollen from differentplants
may be very different,and spores and pollen may be extrapaludal. The cuti-
cular analysis gives more meaningful results regarding the peat-forming veg-
etation although leaf epiderms too may be of different resistance, and leaves,
for example, of Cyperaceae, may be deposited in subaquatic muds. The best
results from cuticular analyses have been obtained by Schneider (1978,
1980,1984,1986) in the Miocene brown coal of Lusatia. Wood histology should
be much more used to clarifythe vegetational problems of brown-coal genesis.
Because brown coal stillresembles peat in its microscopic appearance, com-
parison with the maceral composition of peats from different environments
and types of vegetation are especially helpful to determine brown-coal facies.
It must be taken into account that the ratio humotelinite:humode-
trinitedepends primarily on the plant material deposited: for example, lignin-
rich woods, especially of conifers, remain commonly well preserved structur-
ally whereas the lignin-poor herbs like sedges and grasses are easily decom-
posed, yielding mainly humodetrinite. Very important for future work is the
microscopic study of botanically identifiedplant remains from coals and as-
sociated rocks, using the current coal petrological methods. This method has
been successfullyapplied, e.g.,for vitrinitesfrom lycopod barks and cordaitean
woods in Carboniferous coals,for coniferous leaves in German Wealden coals,
and for the striking coniferous root tissue of Marcoduria inopinata in Tertiary
brown coals.
Most effective were the microscopic studies of peats from different types of
vegetation and environments, carried out by Spackman and his school, espe-
cially by Cohen in the southeastern United States. A very recent paper of Cohen
et al. (1987) contains important new results. For Carboniferous hard coals, the
study of petrified peats preserved in coal balls has been very informative. These
studies should be carried out with the help of palaeobotanists to guarantee the
true identification of plants and plant organs, as well as to ensure a correct
interpretation with regard to the ecology of peat formation. Recent papers of
T.L. Phillips and co-authors (see references) revealed, for example, that in-
grown roots participate to the same amount in the composition of Carbonifer-
ous peats as do aerial plant materials, and that in lycopod peats the ratio
bark:wood is as high as 9:1. The high proportion of root tissues is a character-
istic of all peats, especially of those which were deposited under relatively dry
74

conditions or in a tropical climate where the high temperatures promote dis-


integration and bacterial decomposition of aerial plant remains.
The great difference between the petrographic composition of Carboniferous
coals from the northern hemisphere and the Permian Gondwana coals of the
southern hemisphere is certainly mainly due to the great disparity in climatic
conditions of peat formation. To clarify the distinctive composition of many
Gondwana coals with their extremely high inertinite contents, it seems nec-
essary to study more modern peats that are deposited in cool, especially in sub-
arctic climates. Smyth (1980b) has drawn attention to the specific conditions
in sub-arctic mires (permafrost with dry conditions in the winter and very wet
conditions during the summer). These periodically changing conditions from
dry and oxidative to wet and probably more anoxic, together with the stunted,
predominantly herbaceous and possibly mossy vegetation may explain the un-
usual composition of many Gondwana coals. As mentioned already, electron
microscopical observations by Liu and Taylor (1987) point to very fine lipoid
material between the inertodetrinites of inertites in Gondwana coals. These
results suggest a subaquatic deposition.
A very modern topic for coal petrologists is the possible ombrogenous, oligo-
trophic origin of most mineable coal seams, especially of those with a relatively
low ash content. Because it is known that the vegetation of raised bogs is not
always mossy and stunted (as in temperate climates) the raised bogs of Bor-
neo, which carry a dense arborescent vegetation, are often cited as the modern

Fig. 21. Some well-known past masters of genetic coal petrology.


75

PLATE V

Some well-known coal petrologistswho are working on the genesis of coal.


7(5

example of ombrogenous coal formation. Although successions from more eu-


trophic to more oligotrophic conditions of peat formation seem to be reason-
able (as first suggested by A.H.V. Smith for English Carboniferous coals) the
idea that all mineable coals have been deposited in raised bogs is certainly
much exaggerated. It is not possible to understand how whole brown coal seams
of 100 m thickness or the many mineable Carboniferous coals which were de-
posited in steadily sinking graben zones and foredeeps, respectively, were de-
rived mainly from ombrogenous, oligotrophic peats. Peat experts like Grosse-
Brauckmann (1979b) point to the fact that, instead of the common endoge-
nous successional tendency from eutrophic to oligotrophic peats, the real pro-
cess of plant successions is controlled by other environmental, especially by
water conditions. If the water supply remains constant the plant communities
do so likewise.
Severe climatic changes certainly influence the kind of vegetation in mires
and also the kind of peatification, but the main causes for abrupt facies changes
in coal seams are accidents like crevasses and fires. Therefore, striking changes
of lithotypes and microlithotypes in seam profiles are often combined either
with mineral layers or with concentrations of pyrofusites.
Marine inundations as well as calcareous environments lead to an intense
bacterial activity due to decreasing acidity and increasing nutrient supply. Ma-
rine-influenced peats and coals are rich in sulfur and in pyrite due to the ac-
tivity of sulfate-reducing bacteria which, moreover, lead to an incorporation of
fatty and proteinaceous bacterial substances into the humic matter, which
causes finally a high hydrogen content, a relatively low reflectance and strong
fluorescence of vitrinites, as well as the well known high plasticity of marine-
influenced coking coals.
Many coal petrologists have contributed to the results reviewed above. Fig-
ure 21 and Plate V present some past masters of genetic coal petrology and
coal petrologists, who are still active in this special field of research, respectively.
Finally, the author would like to stress that the microlithotypes of the inter-
national nomenclature (Stopes-Heerlen System) have been classified mainly
from the viewpoint of technological behaviour, the proportions of the three
maceral groups being decisive. To study the genesis of microlithotypes it is
necessary to consider the nature and proportion of individual macerals and
submacerals, the type and amount of mineral matter and the texture (e.g.,
microbedding), as has been mentioned in the foregoing. A genetic classifica-
tion of coal facies which is applicable to all coals probably will not be possible
for coals of different climates, geological ages and types of vegetation that in-
fluenced the petrographic composition of coal layers. Facies classifications have
to be developed from case to case for certain comparable coal deposits.

ACKNOWLEDGEMENTS

The author would like to thank the following authors for their permission
to reproduce published figures: R.M. Bustin, C.F.K. Diessel, J.S. Esterle, R.D.
77

Harvey, R.B. Johns, A.P. Kershaw, T.L. Phillips, W. Riegel, A.H.V. Smith, W.
Schneider, M. Smyth and G.H. Taylor. A.M. George was very helpful in pro-
viding literature on Australian brown coals.
The author is very grateful to B. Alpern, and especially to P.C. Lyons, D.G.
Murchison and A.H.V. Smith who revised the manuscript and improved the
English language. Finally, she would like to thank the President of the Geo-
logisches Landesamt Nordrhein-Westfalen, Krefeld, for facilities to prepare
this paper.

REFERENCES

Alpern, B., 1966. Un example int~ressantde houillificationdans lebassin Lorrain et ses prolonge-
ments. Adv. Org. Geochem., 1964: 129-145.
Alpern, B. and Lemos de Sousa, M.J., 1970. Sur lepouvoir r~flecteurde la vitriniteet de la fusinite
des houiUes. C.R. Acad. Sci. Paris, 271: 956-959.
Alpern, B. and Quesson, A., 1956. Etude par autoradiographie de la r~partition des cendres de
charbons actives.Bull. Soc. Ft. Min~r. Cristallogr.,L X X I X :449-463.
Alpern, B. and Teichmliller,M., 1971. Classificationet corrdlationde la vitrinite (houilles)et de
l'huminite (lignites).C.R. Acad. Sci. Paris,272: 775-778.
Altebiiumer, A.M., 1983. Geochemische Untersuchungen zur Kliirung der Fazies sowie der pri-
m~iren Migration auf Menge und Zusammensetzung des organischen Materials im FlSz Ka-
tharina (Westfal A) und in den hangenden Schiefertonen (Westfal B) im Ruhrgebiet. Disser-
tation Technische Hochschule Aachen, Aachen, 305 pp.
Anderson, J.A.R., 1964. The structure and development of the peat swamps of Sarawak and Bru-
nei. J. Trop. Geogr., 18: 7-16.
Anderson, J.A.R., 1983. Tropical peat swamps of Western Malasia. In: A.J.P. Gore (Editor),
Ecosystems of the World, 4B. Mires: Swamp, bog, fen and moor. Regional studies.Elsevier,
Amsterdam, pp. 181-199.
Bartram, K.M., 1987. Lycopod succession in coals: an example from the Low Barnsley seam
(Westphalian B), Yorkshire, England. In A.C. Scott (Editor), Coal and Coal-Bearing Strata:
Recent Advances. Geol. Soc. Spec. Publ., 32: 187-199.
Bartram, K.M., Jeram, A.J. and Selden, P.A., 1987. Arthropod cuticlesin coal.J. Geol. Soc. Lon-
don, 144: 513-517.
Belyaev, S.S., Yy-Lein, A. and Ivanov, M.V., 1981. Role of methane-producing and sulphate-
reducing bacteria in the destruction of organic matter. In: P.A. Trudinger and M.R. Walter
(Editors),Biochemistry of Ancient and Modern Environments. Springer,Berlin,pp. 235-242.
Benda, L., 1960. Beitriigezur Stratigraphie und Fazies des rheinischen HauptbraunkohlenflSzes
auf Grund einer kutikularanalytischen Untersuchung der Tagebaue Vereinigte Ville,Berren-
rath, Liblar,Lucretia,Sybilla,Fischbach und Fortuna. Neues Jahrb. Geol. Pal~iont.,Abh., 109:
225-260.
Benedict, L.G., Thompson, R.R., Shigo, J.J. and Aikman, R.P., 1968. Pseudovitrinite in Appa-
lachian coking coals.Fuel, 47: 125-143.
Blackburn, D.T., 1980. Floristic,environmental and lithotype correlations in the Yallourn for-
mation, Victoria. Paper presented at the Conference on Cainozoic evolution of continental
south-eastern Australia,Canberra, November 1980.
Blackburn, D.T., 1980. Megafossil/lithotype correlationsin a transect of the Yallourn open cut
mine. Unpubl. Rep. to the State ElectricityCommission of Victoria, 82 pp.
Blackburn, D.T., 1981. Floristiccontrol on lithotype banding within the Yallourn coal seam in
78

the Yallourn open cut, Victoria: evidence from megaibssil assemblages. Unpublished report to
the State Electricity Commission of Victoria, SECV Palaeobotany project, major report No.
2, April 1982, 67 pp.
Brown, H.R., Taylor, G.H. and Cook, A.C., 1964. Prediction of coke strength from the rank and
petrographic composition of Australian coals. Fuel., 43: 43-54.
Brown, K., Diessel, C.F.K., Mc Hugh, E.A., Wolf, M. and Wolff-Fischer, E., 1985. The fluorescence
properties of Carboniferous and Permian coking coals. Proc. Int. Conf. Coal Sci., Sydney,
N.S.W., pp. 649-652.
Bustin, R.M., Cameron, A.R., Grieve, D.A. and Kalkreuth, W.D., 1985. Coal Petrology: its prin-
ciples, methods and applications. Geol. Soc. Canada, Short Course Notes, 3: 2nd ed., 273 pp.
Casagrande, D.J., 1971. Geochemistry of amino acids in selected Florida peats. Dissert. Abstr.
Ann Arbor, 32: No. 1.
Casagrande, D.J., Siefert, K., Berschinski, C. and Sutton, N., 1977. Sulfur in peat forming systems
of the Okefenokee swamp and Florida Everglades: origin of sulfur in coal. Geochim. Cosmo-
chim. Acta, 41: 161-167.
Casagrande, D., Ferguson, A., Boudreau, J. Predny, R. and Fotden, Ch., 1985. Organic geochemical
investigations in the Okefenokee Swamp, Georgia: the fate of fatty acids, glucosamine, cellulose
and lignin. C.R., IX. Congr. Int. Strat. G6ol. Carbonif'ere, 4: 193-204.
Cecil, C.B., Stanton, R.W., Neuzil, S.G., Dulong, F.T., Ruppert, L.F. and Pierce, B.S., 1985. Pa-
leoclimate controls on late Paleozoic sedimentation and peat formation in the central Appa-
lachian Basin (U.S.A.). Int. J. Coal Geol., 5: 195-230.
Chaffee, A.L. and Johns, R.B., 1985. Aliphatic components of Victorian brown coal lithotypes.
Org. Geochem., 8: 349-365.
Chaffee, A.L., Johns, R.B., Baerken, M.J., Leeuw, J.W., Schenck, P.A. and Boon, J.J., 1984.
Chemical effects in gelification processes and lithotype formation in Victorian brown coals.
Adv. Organic Geochem., 1983 (6):409-416.
Chaloner, W.G., 1955. On Sporangiostrobus langfordi sp. nov., a new fossil lycopod cone from
Illinois. Am. Mid. Nat., 55: 437-442.
Clymo, R.S., 1987. Rainwater-fed peat as a precursor of coal. In: A.C. Scott, (Editor) Coal and
Coal-Bearing Strata: Recent Advances. Geol. Soc. Spec. Publ., 32: 17-23.
Cohen, A.D., 1968. The petrology of some peats of southern Florida (with special reference to the
origin of coal ). Ph. D. Thesis, Pennsylvania State University, University Park, Pennsylvania,
352 pp.
Cohen, A.D., 1973. Petrology of some Holocene peat sediments from the Okefenokee swamp-
marsh complex of southern Georgia. Geol. Soc. Am. Bull., 84: 3867-3878.
Cohen, A.D., 1974. Petrography and paleoecology of Holocene peats from the Okefenokee swamp-
marsh complex of Georgia. J. Sediment. Petrol., 44: 716-726.
Cohen, A.D. and Spackman, W., 1977. Phytogenic organic sediments and sedimentary environ-
ments in the Everglades-mangrove complex. Palaeontographica, Abt. B, 162, 144 pp.
Cohen, A.D. and Spackman, W., 1980. Phytogenic organic sediments and sedimentary environ-
ments in the Everglades mangrove complex of Florida. Part III. The alteration of plant material
in peats and the origin of macerals. Palaeontographica, Abt. B, 172: 125-149.
Cohen, A.D., Spackman, W. and Dolson, P., 1984. Occurrence and distribution of sulfur in peat-
forming environments of southern Florida. Int. J. Coal Geol., 4: 73-96.
Cohen, A.D., Spackman, W. and Raymond, R., 1987. Interpreting the characteristics of coal seams
from chemical, physical and petrographic studies of peat deposits. In: A.C. Scott (Editor),
Coal and Coal-Bearing Strata: Recent Advances. Geol. Soc. Spec. Publ., 32: 107-125.
Cook, A.C., 1975. The spatial and temporal variation of the type and rank of Australian coals. In:
A.C. Cook (Editor), Australian Black Coal. Australasian Inst. Min. Metall., Iltawarra Branch,
A.B.C. Symp. Melbourne, pp. 63-84.
79

Cook, A.C., 1982. Organic facies in the Eromanga Basin. In: P.S. Moore and T.J. Mount (Edi-
tors), 1982. Eromanga Symposium. Geol. Soc. Aust. & Pet. Explor. Soc. Aust. pp, 234-257.
Cook, A.C. and Struckmeyer, H., 1986. The role of coal as a source rock for oil.In: R.C. Glenie
(Editor), Second south-eastern Australiaoilexploration symposium, Melbourne, pp. 419-432.
Corr~a da Silva,Z.C., Hagemann, H.W., P(ittmann, W. and Wolf, M., 1986. Petrographische und
geochemische Eigenschaften ausgewiihltersildbrasilianischerKohlen. Zentralbl.Geol. Paliion-
tol.I, 1985, Heft 9/10: 1565-1578.
Crelling,J.C., Dutcher, R.R. and Lange, R.V., 1982. Petrographic and fluorescence properties of
resinitemacerals from western U.S. coals,In: Proceed. 5th Symp. Geology of Rocky Mountains
Coal (1982, Utah). Geol. Mineral. Surv. Bull.,118: 187-191.
Diessel,C.F.K., 1983. Carbonization reactions of inertinitemacerals in Australian coals.Fuel, 62:
883-892.
Diessel, C.F.K., 1985. Fluorometric analysisof inertinite.Fuel, 64: 1542-1546.
Diessel,C.F.K., 1986. On the correlationbetween coal faciesand depositionalenvironments. Proc.
20th Syrup., Dep. Geol.,Univ. Newcastle, N.S.W., pp. 19-22.
Durand, B. and Paratte, M., 1983. Oil potential of coals:a geochemical approach. In: J. Brooks
(Editor), Petroleum Geochemistry and Exploration of Europe. Blackwell, Oxford, pp. 225-
265.
Esterle, J.S. and Ferm, J.C., 1986. Relationship between petrographic and chemical properties of
coal seam geometry, Hance seam, Breathitt Formation, southeastern Kentucky. Int. J. Coal
Geol., 6: 199-214.
Flaig, W., 1965. Chemie der Humusstoffe und Inkohlung. Paper read at the Colloquium "Chemie
und Physik der Kohle", Aachen, 28 pp.
Frey-Wyssling, A., 1959. Die pflanzliche ZeUwand. Springer, Berlin-GSttingen-Heidelberg, 367
pp.
Fulton, I.M., 1987. Genesis of the Warwickshire thick coal: a group of long-residence histosols. In:
A.C. Scott (Editor), Coal and Coal-Bearing Strata: Recent Advances. Geol. Soc. Spec. Publ.,
32: 201-218.
George, A.M., 1975. Brown coal lithotypes in the Latrobe Valley deposits. State Electricity Comm.,
Victoria, Petrol. Rep. No. 17, 36 pp.
Given, P.H. and Miller, R.N., 1985. Distribution of forms of sulfur in peats from saline environ-
ments in the Florida Everglades. Int. J. Coal Geol., 5: 397-409.
Gloe, C.S., 1984. The geology, discovery, and assessment of the brown coal deposits of Victoria.
In: J.T. Woodcock (Editor), Victorian Brown Coal - A Huge Fortune in Chancery. Australas.
Inst. Min. Metall. Monogr., 11: pp. 79-109.
Goodarzi, F., 1984. Chitinous fragments in coal. Fuel, 63: 1504-1507.
Gothan, W., 1924. Studium fiber die Bildung der Schwelkohle und des Pyropissits. Braunkohle,
23: 725-733.
Gothan, W., 1937. Kohle. Enke, Stuttgart, 432 pp.
Grosse-Brauckmann, G., 1979a. Zur Deutung einiger Makrofossil-Vergesellschaftungen unter dem
Gesichtspunkt der Torfbildung. In: O. Williams and R. Tiixen (Editors), Werden und Verge-
hen yon Pflanzengesellschaften. Ber. Int. Symp. Int. Vet. Vegetationskunde, J. Cramer, Vaduz,
pp. 111-132.
Grosse-Brauckmann, G., 1979b. Sukzessionen bei einigen torfbildenden Pflanzengesellschaften
(nach Ergebnissen yon Grossrest-Untersuchungen an Torfen). In: R. Tiixen and W.H. Som-
mer (Editors), Gesellschaftsentwicklung (Syndynamik). Ber. Int. Symp. Int. Vet. Vegetations-
kunde, J. Cramer, Vaduz. pp. 393-412.
Guennel, G.K., 1960. Re-evaluation of the Russian paper coal. Geol. Soc. Am. Bull., 71: 1875-
1876.
Guennel, G.K. and Neavel, R.C., 1959. Paper coal in Indiana. Science, 129: 1671-1672.
80

Gutjahr, C.C.M., 1983. Introduction to incident-light microscopy of oil and gas source rocks. Geol.
Mijnbouw, 62: 417-425.
Hacquebard, P.A. and Donaldson, J.R., 1969. Carboniferous coal deposition associated with flood-
plain and limnic environments in Nova Scotia. In: E.C. Dapples and M.E. Hopkins (Editors).
Environment of Coal Deposition. Geol. Soc. Am., Spec. Pap., 114: 143-191.
Hagemann, H.W., 1987. Mikrofluorometrische Untersuchungen an Extraktfilmen wm Kohlen
einer Inkohlungsreihe. Zeiss Information, 29: 40-43.
Hagemann, H.W. and Hollerbach, A., 1980. Relationship between the macropetrographic and
organic geochemical composition of lignites. In: A.G. Douglas and J.R. Maxwell (Editors),
Advances in Organic Geochemistry 1979, pp. 631-638.
Hagemann, H.W. and Wolf, M., 1987. New interpretations of the facies of the Rhenish brown coal
of West Germany. Int. J. Coal Geol., 7: 337-348.
Harvey, R.D. and Dillon, J.W., 1985. Maceral distributions in Illinois coals and their paleoenvi-
ronmental implications. Int. J. Coal Geol., 5: 141-165.
Hatcher, P.G., Breger, I.A. and Earl, W.L., 1981. Nuclear magnetic resonance studies of ancient
buried wood, I. Observations on the origin of coal to the brown coal stage. Org. Geochem., 3:
49-55.
Hickling, H.G.A. and Marshall, C.E., 1932. The microstructure of the coal of certain fossil trees.
Trans. Inst. Min. Eng., 84: 13-23.
Hickling, H.G.A. and Marshall, C.E., 1933. The microstructure of the coal of certain fossil tree
barks. Trans. Inst. Min. Eng., 86: 56-75.
Hiltmann, W., 1976. Pollenanalytische Untersuchungen im rheinischen HauptbraunkohlenflSz
der Tagebaue Frechen und Fortuna unter besonderer Berticksichtigung der makropetrogra-
phischen Ausbildung der Kohle. Dissertation, Technische Hochschule, Aachen, 162 pp.
Hirsch, P.B., 1954. X-ray scattering from coals. Proc. R. Soc. London. Ser. A, 266: 143-169.
Hoehne, K., 1954. Sekretvitrit. Geologic, 3: 560-575.
Hoffmann, E., 1933. Neue Erkenntnisse fiber die Vorg~inge der F15zbildung. Bergbau, 7: 89-94.
Hower, J.C., Trinkle, E.J., Graese, A.M. and Neuder, G.L., 1987. Ragged edge of the Herrin (No.
11 ) coal, western Kentucky. Int. J. Coal Geol., 7: 1-20.
Hunger, R., 1953. Mikrobotanisch-stratigraphische Untersuchungen der Braunkohlen der slid-
lichen Oberlausitz und die Pollenanalyse als Mittel zur Deutung der F15zgenese. Freiberger
Forschungsh. C, 8: 1-38.
Hutton, A. and Cook, A., 1980. Influence of alginite in the reflectance of vitrinite from Joadja,
N.S.W. and some other coals and oil shales containing alginite. Fuel, 59: 711-716.
Hutton, A.C., Kantsler, A.J., Cook, A.C. and McKirdy, D.M., 1980. Organic matter in oil shales.
APEA J., 20: 44-68.
International Handbook of Coal Petrography, 2nd edition 1963:184 pp., reprinted 1985.1st sup-
plement to 2nd edition, 1971:197 pp., reprinted 1985.2nd supplement to 2nd edition, 1975.:
60 pp. Centre National de la Recherche Scientifique, Paris.
Jacob, H., 1968. Genesis der hellen Straten yon Weichbraunkohlen-F15zen. Z. Dtsch. Geol. Ges.,
118: 102-110.
Jacob, H., 1985. Disperse solid bitumens as an indicator for migration and maturity in prospecting
for oil and gas. Erd~il Kohle, 38: 365.
Johns, R.B., Verheyen, T.V. and Chaffee, A.L., 1981. Chemical characterization of Victorian brown
coal lithotypes. Proc. Int. Conf. Coal Science, D/Jsseldorf, September 1981, pp. 863-868.
Jurasky, K.A., 1934/35. Kutikular-Analyse. Biol. Gen., 10: 383-402, 1934.; 11, I: 227-244; 11, II:
1 26. Wien-Leipzig.
Jurasky, K.A., 1940. Kohle. Naturgeschichte eines Rohstoffs. Springer, Berlin, 170 pp.
Kaegi, D.D., 1985. On the identification and origin of pseudovitrinite. Int. J. Coal Geol., 4: 309-
319.
81

Kalkreuth, W., 1982. Rank and petrographic composition of selectedJurassic-Lower Cretaceous


coals of BritishColumbia, Canada. Bull.Can. Pet. Geol.,30: 112-139.
Kershaw, A.P. and Sluiter,I.R., 1982. The applicationof pollen analysisto the elucidationof
Latrobe Valley brown coal depositionalenvironments and stratigraphy.Aust. Coal Geol.,4:
169-186.
Kershaw, A.P., Bolger, P.F., Sluiter, I.R.K., Baird, J.G. and Whitelaw, M., 1988. The nature and
evolution of lithotypes in the Tertiary brown coals of the Latrobe Valley, southeastern Aus-
tralia. Int. J. Coal Geol. (in press).
Kilpper, K., 1960. Pflanzenfiihrung, Fazies und Bildungsverh~iltnisse im "HauptflSz der Ville",
eine kutikularanalytische Untersuchung in den Tagebauen Neurath und Frimmersdorf-Stid
des rheinischen Braunkohlenreviers. Neues Jahrb. Geol. Pal~iontol., Abh., 109: 261-308.
Klein-Reesink, J. and Riegel, W., 1983. Kohlenpetrographische Aspekte der obermiozRnen Braun-
kohle van Viehhausen (Oberpfalz). Weltenburger Akademie, Erwin Rutte Festschrift, pp. 125-
131.
Klein-Reesink, J., Riegel, W. and Schaub, K., 1982. Zur petrographischen Konstitution alt-und
jungtertiRrer Braunkohlen Niederhessens. Z. Dtsch. Geol. Ges., 133: 309-337.
Klusemann, H. and Teichmiiller, R., 1954. Begrabene W~ilder im Ruhrkohlenbecken. Natur Volk,
84: 373-382.
Koch, J., 1966. Petrologische Untersuchungen an jungpleistoz~inen Schieferkohlen aus dem Al-
penvorland, der Schweiz and Deutschlands mit Vergleichsuntersuchungen an holoz~inen Tor-
fen. Dissertation, Technische Hochschule, Aachen, 186 pp.
Koch, J., 1969a. Uber ein naeheiszeitRhnliches Aquivalent der HeUen Schichten (Bitumenkohle)
von Weichbraunkohlen. Braunkohle, 21: 240-242.
Koch, J., 1969b. Mikropetrographische Untersuchungen an einigen organischen Komponenten
jungpleistozRner und holozRner Torfe Stiddeutschlands und der Schweiz. Geol. Jahrb., 87: 333-
360.
Koch, J., 1970. Petrologische Untersuchungen an nieders~ichsischen Torfen und Weichbraunkoh-
len. Geol. Mitt., 10: 113-150.
Koch, J. and Scheuermann, L., 1970. Elektronenmikroskopische Untersuchungen tiber die Struk-
turen der humosen Grundmasse von lufttrockenen Weichbraunkohlen. Braunkohle, W~irme
Energ., 22: 88-94.
Kosanke, R.M., 1955. Petrographic and microchemical studies of coal. 2nd Conference on the
Origin and Constitution of Coal, Crystal Cliffs, N.S., 1952, pp. 248-267.
Kosanke, R.M. and Harrison, J.A., 1957. Microscopy of resin rodlets of Illinois coal. Ill. State
Geol. Surv., Circ. 234, 14 pp.
Kosters, E.C., Chmura, G.L. and Bailey, A., 1987. Sedimentary and botanical factors influencing
peat accumulation in the Mississippi delta. J. Geol. Soc. London, 144: 423-434.
Lin, R., Davis, A., Bensley, D.F. and Derbyshire, F.J., 1986. Vitrinite secondary fluorescence, its
chemistry and relation to the development of a mobile phase and thermoplasticity in coal. Int.
J. Coal Geol., 6: 215-228.
Littke, R., 1985. Fl~zaufbau in den Dorstener, Horster und Essener Schichten der Bohrung Wul-
fener Heide 1 (n~rdliches Ruhrgebiet). Fortschr. Geol. Rheinland Westfalen, 33: 129-159.
Littke, R., 1987. Petrology and genesis of Upper Carboniferous seams from the Ruhr region, West
Germany. Int. J. Coal Geol., 7: 147-184.
Liu, S.Y. and Taylor, G.H., 1987. Liptodetrinite of algal origin in inertinite-rich coals. Adv. Stud.
Sydney Basin, 21st Newcastle Symposium, Proc., pp. 57-60.
Luly, J., Sluiter, I.R. and Kershaw, A.P., 1980. Pollen studies of Tertiary brown coals: preliminary
analyses of lithotypes within the Latrobe Valley, Victoria. Monash Publ. Geography, 23, Mon-
ash Univ., 78 pp.
Lyons, P.C., Finkelman, R.B., Thompson, C.L. Brown, F.W. and Hatcher, P.G., 1982. Properties,
82

origin and nomenclature of rodlets of the inertinite maceral group in coals of the central Ap-
palachian basin U.S.A. Int. J. Coal Geol., 1: 313-346.
Lyons, P.C., Hatcher, P.G., Minkin, J.A., Thompson, C.L., Larson, R.R., Brown, Z.A. and Pheifer,
R.N., 1984. Resin rodlets in shale and coal (Lower Cretaceous), Baltimore Canyon Trough.
Int. J. Coal Geol., 3: 257-278.
Lyons, P.C., Hatcher, P.G. and Brown, F.W., 1986. Secretinite, a proposed new maceral of the
inertinite maceral group. Fuel, 65: 1094-1098.
Mackay, G.H., Attwood, D.H., Gaulton, R.J. and George, A.M., 1985. The cyclic occurrence of
brown coal lithotypes. State Electricity Commission of Victoria, Report No. 256, 23 pp.
M~idler, K., 1963. Die figurierten organischen Bestandteile des Posidonienschiefers. Beih. Geol.
Jahrb., 58: 287-407.
Mahaffy, J.F., 1986. Profile patterns of coal and peat palynology in the Herrin (No. 6) coal mem-
ber, Carbondale Formation, Middle Pennsylvanian of southern Illinois. In: J.T. Dutro and
H.W. Pfefferkorn (Editors), Paleontology, Paleoecology, Paleogeography. 9th Int. Congr. Strat.
G~ol. Carbonif~re, Washington, 1979, C.R., 5:25-34
Marchioni, D.L., 1980. Petrography and depositional environment of the Liddell seam, Upper
Hunter Valley, New South Wales. Int. J. Coal Geol., 1: 35-61.
Masran, Th. C. and Pocock, St. A.J., 1981. The classification of plant-derived particulate organic
matter in sedimentary rocks. In: J. Brooks (Editor), Organic Maturation Studies and Fossil
Fuel Exploration. Academic Press, London, pp. 157-175.
Mc Cabe, P.J., 1987. Facies studies of coal and coal-bearing strata. In: A.C. Scott (Editor), Coal
and Coal-bearing Strata: Recent Advances. Geol. Soc. Spec. Publ., 32: 51-66.
Minnigerode, C. and Riegel, W., 1983. Makropetrographische und palynologische Untersuchun-
gender Braunkohle von Borken (Bez. Kassel). Neues Jahrb. Geol. Pal~iontol., Monatsh. 1983,
5: 300-320.
Moore, P.D., 1987. Ecological and hydrological aspects of peat formation. In: A.C. Scott (Editor),
Coal and Coal-Bearing Strata: Recent Advances. Geol. Soc. Spec. Publ., 32: 7-15.
Mukhopadhyay, P.K., 1986. Petrography of selected Wilcox and Jackson group lignites from the
Tertiary of Texas. In R.B. Finkelman and D.J. Casagrande (Editors), Geology of Gulf Coast
Lignites. 1986 Annu. Meet. Geol. Soc. Am., Coal Geology Div., Field Trip, pp. 126-145.
Mukhopadhyay, P.K. and Gormly, J.R., 1984. Hydrocarbon potential of two types of resinite. Org.
Geochem., 6: 439-454.
Murchison, D.G., 1966. Infra-red spectra of resinites and their carbonized and oxidized products.
Coal Science, Adv. Chem. Ser. 55, Washington, D.C., pp. 307-331.
Murchison, D.G., 1987. Recent advances in organic petrology and organic geochemistry: an over-
view with some reference to "oil from coal". In: A.C. Scott (Editor), Coal and Coal-Bearing
Strata: Recent Advances. Geol. Soc. Spec. Publ., 32: 257-302.
Nandi, B.N. and Montgomery, D.S., 1967. Thermal behaviour of massive and granular micrinite.
Fuel, 46: 394-398.
Neavel, R.C., 1981. Origin, petrography, and classification of coal. In: M.A. Elliot (Editor), Chem-
istry of Coal Utilization, 2nd supplementary volume. John Wiley, New York, NY, pp. 91-158.
Nelson, W.J., Eggert, D.L., DiMichele, W.A. and Stecyk, A.C., 1985. Origin and discontinuities
in coal-bearing strata at Roaring Creek (Basal Pennsylvanian of Indiana). Int. J. Coal Geol.,
4: 355-370.
Nemejc, F., 1931. A study of the systematic position of the fructification called Sporangiostrobus
Bode. Bull. Int. Acad. Sci. Prague, 32: 68-80.
Ottenjann, K., 1985. Fluoreszenzeigenschaften von FlSzproben aus stratigraphisch jungen Schi-
chten des Oberkarbons. Fortschr. Geol. Rheinld. Westfalen, 33: 169-195.
Ottenjann, K., 1988. Fluorescence alteration and its value for studies of maturation and bitumin-
ization. Organic Geochem., 12: 309-321.
Ottenjann, K., Wolf, M. and Wolff-Fischer, E., 1982. Das Fluoreszenz-Verhalten der Vitrinite zur
83

Kennzeichnung der Kokungseigenschaften von Steinkohlen. Gliickauf Forschungsh., 43: 173-


179.
Petrascheck, W., 1947. Die Metamorphose der Kohle und ihr Einfluss auf die sichtbaren Bes-
tandteile derselben. Sitzungsber. ~)sterr.Akad. Wiss., Math.-naturwiss. KI., Abt. I, 156: 375-
444.
Pflug, H.D., 1957. Zur Altersfolge und Faziesgliederung mitteleuropiiischer (insbesondere hes-
sischer) Braunkohlen. Notizbl. Hess. Landesamt Bodenforsch., 85: 152-178.
Phillips, T.L. and Cecil, C.B., 1985. Paleoclimatic controls on coal resources of the Pennsylvanian
system of North America.-Introduction and overview of contributions. Int. J. Coal Geol., 5: 1-
6.
Phillips, T.L. and DiMichele, W.A., 1981. Paleoecology of Middle Pennsylvanian age coal swamps
in southern Illinois, Herrin coal member at Sahara Mine No. 6. In: K.J. Niclar (Editor), Pal-
eobotany, Paleoecology and Evolution, 1. Praeger, New York, NY, pp. 231-284.
Phillips, T.L. and Peppers, R.A., 1984. Changing patterns of Pennsylvanian coal-swamp vegeta-
tion and implications of climatic control on coal occurrence. Int. J. Coal Geol., 3: 205-255.
Phillips, T.L., Peppers, R.A. and DiMichele, W.A., 1985. Stratigraphic and interregional changes
in Pennsylvanian coal-swamp vegetation: environmental inferences. Int. J. Coal Geol., 5: 43-
109.
Plumstead, E.P., 1962. The Permo-Carboniferous coal measures of the Transvaal, South Africa-
an example of the contrasting stratigraphy in the southern and northern hemispheres. 4th
Congr. Int. Stratigr. G~ol. Carbonif'ere, C.R., 2: 545-550.
Radke, M., Schaefer, R.G., Leythaeuser, D. and Teichmiiller, M., 1980. Composition of soluble
organic matter in coals: relation to rank and liptinite fluorescence. Geochim. Cosmochim. Acta,
44: 1787-1800.
Raistrick, A. and Marshall, C.E., 1939. The nature and origin of coal and coal seams. English
Univ. Press, London, 282 pp.
Riegel, W.L., 1965. Palynology of environments of peat formation in southwestern Florida. Ph.D.
Thesis, Pennsylvania State University, University Park, PA, 97 pp.
Rouzaud, J.N., 1984. Relations entre la microtexture et les propri~t~s des matdriaux carbon,s-
Application ~ la caract~risation des charbons. Th~se, Universit~ d'OrlSans, OrlSans, 150 pp.
Russell, N.J., 1984. Gelification of Victorian Tertiary soft brown coal wood. I. Relationship be-
tween chemical composition and microscopic appearance and variation in the degree of geli-
fication. Int. J. Coal Geol., 4: 99-118.
Russell, N.J. and Barton, P.F., 1984. Gelification of Victorian Tertiary soft brown coal wood. II.
Changes in chemical structure associated with variation in the degree of gelification. Int. J.
Coal Geol., 4: 119-142.
Schlesinger, W.H., 1978. Community structure, dynamics and nutrient cycling in the Okefenokee
cypress swamp-forest. Ecol. Monogr., 48: 43-65.
Schneider, W., 1978. Zu einigen Gesetzmiissigkeiten der faziellen Entwicklung im 2. Lausitzer
F15z. Z. Angew. Geol., 24: 125-130.
Schneider, W., 1980. Mikropaliiontologische Faziesanalyse in der Weichbraunkohle. Neue Berg-
bautech., 10: 670-675.
Schneider, W., 1984. Angewandte Pal~iobotanik und Braunkohlenpetrologie - pflanzliche Gewebe
und GeRigebildner in der Braunkohle. Freiberger Forschungsh., C 381: 14-19.
Schneider, W., 1986. Cryptomeria DON (Taxodiaceae) - ein Kohlenbildner im mitteleuro-
piiischen Terti~ir.Z. Geol. Wiss., 14: 735-744.
Schneider, W., 1989. Die neue Deutung von Marcoduria inopinataWEYLAND und ihre kohlen-
geologische Konsequenz. Z. Geol. Wiss., Berlin, in press.
Scott, A.C., 1980. The ecology of some Upper Palaeozoic floras.In: A.L. Panchen (Editor), The
Terrestrial Environment and the Origin of Land Vertebrates. System. Assoc., Spec. Vol. 15.
Academic Press, London, pp. 87-115.
84

Scott, A.C. (Editor), 1987. Coal and Coal-Bearing Strata: Recent Advances. Geol. Soc. Spec.
Publ., Blackwell, Oxford, 32:332 pp.
Scott, A.C. and Collinson, M.E., 1978. Organic sedimentary particles: results from scanning elec-
tron microscope studies of fragmentary plant material. In: W.B. Whaliey (Editor), Scanning
Electron Microscopy in the Study of Sediments. Geoabstracts, Norwich, pp. 137-167.
Scott, A.C. and King, G.R., 1981. Megaspores and coal facies: an example from the Westphalian
A of Leicestershire, England. Rev. Palaeobot. Palynol., 34: 107-113.
Scott, A.C. and Rex, G., 1985. The formation and significance of Carboniferous coal balls. In: H.B.
Whittington (Editor), Extraordinary Fossil Biotas; their Ecological and Evolutionary Signif-
icance. Philos. Trans. Soc. London, Set. B: Biological Sciences, 311 (1148): 123-137.
Seyler, C.A., 1928. The Dictyoxylon cortex of Lycopodiales as a constituent of coal. Trans. R. Soc.
London, Serie B, 216: 353-360.
Sherwood, N.R. and Cook, A.C., 1986. Organic matter in the Tolebuc formation In: Proc. Ero-
manga Basin Symposium Geol. Soc. Aust., Spec. Publ., 12: 255-265.
Shibaoka, M. and Smyth, M., 1975. Coal petrology and the formation of coal seams in some
Australian sedimentary basins. Econ. Geol., 70: 1463-1473.
Smith, A.H.V., 1962. The palaeoecology of Carboniferous peats based on the miospores and pe-
trography of bituminous coals. Proc. Yorkshire Geol. Soc., 33: 423-474.
Smith, A.H.V., 1968. Seam profiles and seam characters. In: D.G. Murchison and T.S. Westoll
(Editors), Coal and Coal-Bearing Strata. Oliver Boyd, Edinburgh, pp. 31-40.
Smith, G.C. and Cook, A.C., 1980. Coalification paths of exinite, vitrinite and inertinite. Fuel, 59:
641-647.
Smyth, M., 1975. Coal petrology and the formation of coal seams in some Australian sedimentary
basins. Econ. Geol., 70: 1463-1473.
Smyth, M., 1980a. Thick coal members: products of an inflationary environment? J. Coal Geology
Group, Geol. Soc. Aust., 2: 53-72.
Smyth, M., 1980b. Coal encounters of the third kind: Triassic. J. Coal Geology Group, Geol. Soc.
Aust., 2: 161-177.
Smyth, M., 1984. Coal microlithotypes related to sedimentary environments in the Cooper Basin,
Australia. Spec Publ., Int. Assoc. Sediment., 7: 333-347.
Smyth, M. and Cook, A.C., 1976. Sequence in Australian coal seams. Math. Geol., 8: 529-547.
Snowdon, L.R. and Powell, T.G., 1982. Immature oil and condensate - Modification of hydrocar-
bon generation model for terrestrial organic matter. Bull. Am. Assoc. Pet. Geol., 66: 775-788.
Snyman, C.P., 1961. A comparison between the petrography of South African and some other
Palaeozoic coals. University of Pretoria, N. Reeks, 15, 58 pp.
Sontag, E. and Schneider, W., 1982. Zur mikropaliiobotanischen Gliederung des 2. Niederlausitzer
FlSzhorizontes. Z. Angew. Geol., 28: 470-480.
Soos, L., 1963.0bet das sogenannte dunkle Harz der terti~iren Kohlen, insbesondere Ungarns.
Ann. Univ. Sci. Budapest, Sect. Geol., 6: 129-151.
Soos, L., 1966. Die Kennzahlen der Braunkohlen-Gemengteile, 3. Phlobaphenit. Acta Geol. Hun-
garia, 10: 65-68.
Stach, E., 1936. Zur Petrographie der Saarfettkohle. Abh. Preuss. Geol. Landesanst., Neue Folge,
171: 83-104.
Stach, E., 1955. Crassidurain - a means of seam correlation in the Carboniferous coal measures
of the Ruhr. Fuel, 34: 95-118.
Stach, E., Mackowsky, M.Th., Teichmtiller, M., Taylor, G.H., Chandra, D. and Teichmtiller, R.,
1982. Stach's Textbook of Coal Petrology. Borntraeger, Stuttgart, 535 pp.
Stopes, M., 1919. On the four visible ingredients in banded bituminous coal. Proc. R. Soc., London,
B, 90: 470-487.
Stout, S.A. and Bensley, D.F., 1987. Fluorescing macerals from wood precursors. Int. J. Coal Geol.,
7: 119-133.
85

Stout, S.A. and Spackman, W., 1987. A microscopic investigationof woody tissuesin peats: some
processes active in the peatificationof ligno-cellulosiccellwalls.Int.J. Coal. Geol. 8: 55-68.
Stout, S.A., Spackman, W. and Boon, J.J.,1987. Peatificationand early coalificationof wood as
deduced by analyticalpyrolysis and microscopic methods. 4th Ann. Meet. Soc. Org. Petrol.,
Sept. 30-Oct. 3 1987, San Francisco, Abstracts and Program, pp 32-33.
Styan, W.B. and Bustin, R.M., 1983a. Petrography of some Frazer River delta peat deposits:coal
maceral and microlithotype precursors in temperate-climate peats. Int.J. Coal Geol., 2: 321-
370.
Styan, W.B. and Bustin, R.M., 1983b. Sedimentotogy of Frazer River delta deposits: a modern
analogue for some deltaic coals. Int. J. Coal Geol., 3: 101-143.
Taylor, G.H., 1966. The electron microscopy of vitrinites. Coal Science, Adv. Chem. Ser., 55: 274-
283.
Taylor, G.H. and Cook, A.C., 1962. Sclerotinite in coal-its petrology and classification. Geol. Mag.,
XCIX: 41-52.
Taylor, G.H. and Liu, S.Y., 1987. Biodegradation in coals and other organic-rich rocks. Fuel, 66:
1269-1273.
Taylor, T.N., and Millay, M.A., 1981. Morphologic variability of Pennsylvanian lyginopterid seed
ferns. Rev. Palaeobot. Palynol., Spec. Iss., 32: 27-62.
Teerman, S.C., Crelling, J.C. and Glass, G.B., 1987. Fluorescence spectral analysis of resinite
macerals from coals of the Hanna Formation, Wyoming, U.S.A. Int. J. Coal Geol., 7:315-334.
Teichmtiller, M., 1944. Zur Petrographie zweier oberschlesischer Flt~ze mit ~ihnlichem Inkohlungs-
grad, aber verschiedenen Kokungseigenschaften. Z. Prakt. Geol., 52: 1-6.
TeichmiJller, M., 1950. Zum petrographischen Aufbau und Werdegang der Weichbraunkohle (mit
Beriicksichtigung genetischer Fragen der Steinkohlenpetrographie). Geol. Jahrb., 64: 429-488.
Teichmiiller, M., 1952. Vergleichende mikroskopische Untersuchungen versteinerter Torfe des
Ruhrkarbons und der daraus entstandenen Steinkohten. C.R. 3 i~me Congr~s Int. Stratigr.
Gdol. Carbonif'ere, 2: pp. 607-613.
Teichmtiller, M., 1955. Anzeichen mariner Beeinflussung bei der Kohle aus Fl~iz Katharina der
Zeche Friedrich Heinrich. Neues Jahrb. Geol. Pal~iontol., Monatsh., pp. 193-201.
Teichm~iller, M., 1961. Beobachtungen bei einem Torfbrand. Geol. Jahrb., 78: 653-660.
Teichm~iller, M., 1962. Die Genese der Kohle. C.R. 4 i~me Congr~s Int. Stratigr. Gdol. Carbonif'ere,
Heerlen, 1958, 3: 699-722.
Teichmiiller, M., 1974a. Uber neue Macerale der Liptinit-Gruppe und die Entstehung des Micrin-
its. Fortschr. Geol. Rheinl. Westfalen, 24: 37-64.
Teichmtiller, M., 1974b. Entstehung und Veriinderung bituminSser Substanzen in Kohlen in Be-
ziehung zur Entstehung und Umwandlung des Erd~ls. Fortschr. Geol. Rheinl. Westfalen, 24:
65-112.
Teichm~iller, M., 1974c. Generation of petroleum-like substances in coal seams as seen under the
microscope. In: B. Tissot and F. Bienner (Editors), Adv. Org. Geochem.,1973, pp. 321-348.
TeichmUller, M., 1982a. Origin of the petrogrographic constituents of coal In: E. Stach, M.Th.
Mackowsky, M. Teichmiiller, G.H. Taylor, D. Chandra and R. Teichm~iller (Editors), Text-
book of Coal Petrology, 3rd ed. Borntraeger, Stuttgart, pp. 219-294.
Teichmiiller, M., 1982b. Fluoreszenzmikroskopische Anderungen von Liptiniten und Vitriniten
mit zunehmendem Inkohlungsgrad und ihre Beziehungen zu Bitumenbildung und Verko-
kungsverhalten. Geologisches Landesamt, Krefeld (Editor), Van Acken, Krefeld, 119 pp.
Teichmfiller, M., 1984. Fluorescence microscopical changes of liptinites and vitrinites during coal-
ification and their relationship to bitumen generation and coking behaviour. Translated from
German by N.H. Bostick. Soc. Org. Petrol., Spec. Publ., No. 1. Houston, 74 pp.
Teichmiiller, M., 1987a. Recent advances in coalification studies and their application to geology.
In: A.C. Scott (Editor), Coal and Coal-Bearing Strata: Recent Advances. Geol. Soc. Spec.
Publ., 32: 127-169.
~6

Teichmiiller, M., 1987b. Organic material and very low-grade metamorphism. In: M. Frey (Edi-
tor ), Low Temperature Metamorphism. Blackie, Glasgow and London, pp. 114-161.
Teichm~iller, M. and Durand, B., 1983. Fluorescence microscopical rank studies on liptinites and
vitrinites in peats and coals and comparison with results of the Rock-Eval pyrolysis. Int. J.
Coal Geol., 2: 197-230.
Teichmtiller, M. and Ottenjann, K., 1977. Art und Diagenese von Liptiniten und lipoiden Stoffen
in einem Erd~lmuttergestein auf Grund fluoreszenzmikroskopischer Untersuchungen. ErdS1
Kohle, 30: 387-398.
Teichmiiller, M. and Schonefetd, W., 1955. Ein verkieselter Karbontorf im Namur C yon Kupfer-
dreh. Geol. Jahrb., 71: 91-112.
Teichmfiller, M. and Teichmfiller, R., 1982. Fundamentals of coal petrology. In: E. Stach, M.Th.
Mackowsky, M. Teichmtiller, G.H. Taylor, D. Chandra and R. Teichmfiller (Editors), Text-
book of Coal Petrology. Borntraeger, Stuttgart, pp. 5-86.
Teichmtiller, M. and Thomson, P.W., 1958. Vergleichende mikroskopische und chemische Unter-
suchungen der wichtigsten Fazies-Typen im HauptflSz der niederrheinischen Braunkohle.
Fortschr. Geol. Rheinl. Westfalen, 2: 573-598.
Thiessen, R., 1920. Structure in Paleozoic bituminous coals. U.S. Bur. Mines, Bull. 117,295 pp.
Thiessen, R., 1947. What is coal? U.S. Bur. Mines, Circ. 7397, 53 pp.
Thiessen, R. and Sprunk, G.C., 1941. Coal paleobotany. U.S. Bur. Mines, Tech. Pap. 631, 56 pp.
Thiessen, R. and White, D., 1913. The origin of coal. U.S. Bur. Mines, Bull. 38, 390 pp.
Thomson, P.W., 1950/1951. Grunds~itzliches zur terti~iren Pollen-und Sporenmikrostratigraphie
auf Grund einer Untersuchung des HauptflSzes der rheinischen Braunkohle in Liblar, Neu-
rath, Fortuna und Brfihl. Geol. Jahrb., 65: 113-126.
Thomson, P.W., 1954. Der Fazieswechsel im Hauptfl6z der rheinischen Braunkohle im Gebiet
der Grube Fortuna. Geol. Jahrb., 69: 329-338.
Treiber, E., 1957. Die Chemie der Pflanzenzellwand. Springer, Berlin, 511 pp.
Van der Burgh, J., 1973. HSlzer der niederrheinischen Braunkohlenformation. 2. HSlzer der
Braunkohlengruben Maria-Theresia zu Herzogenrath, Zukunft West zu Eschweiler und Victor
(Ztilpich Mitte) zu Ztilpich. Nebst einer systematisch-anatomischen Bearbeitung der Gattung
Pinus L. Rev. Paleobot, Palynol., 15: 73-275.
Van Krevelen. D.W., 1961. Coal. Elsevier, Amsterdam, 514 pp.
Verheyen, T.V. and Johns, R.B., 1981. Structural investigation of Australian coals. I.A character-
ization of Victorian brown coal lithotypes and their kerogen and humic acid fractions by i.r.
spectroscopy. Geochim. Cosmochim. Acta, 45: 1899-1908.
Verheyen, T.V., Johns, R.B. and Blackburn, D.T., 1982. Structural investigation of Australian
coals. II. A ~:~C-NMR study of humic acids from Victorian brown coal lithotypes. Geochim.
Cosmochim. Acta, 46: 2061-2067.
Verheyen, T.V., Johns, R.B., Brysont, R.L., Maciel, G.E. and Blackburn, D.T., 1984. A spectro-
scopic investigation of the banding of lithotypes in Victorian brown coal seams. Fuel, 63: 1629-
1635.
Verma, L. and Martin, J.P., 1976. Decomposition of algal cells and components and their stabili-
zation through complexing with model humic acid-type phenolic polymers. Soil Biol. Biochem.,
8: 85-90.
Von der Brelie, G. and Wolf, M., 1981a. Zur Petrographie und Palynologie heller und dunkler
Schichten im rheinischen HauptbraunkohlenflSz. Fortschr. Geol. Rheinl. Westfalen, 29: 95-
163.
Von der Brelie, G. and Wolf, M., 1981b. "Sequoia" und Sciadopitys in den Braunkohlenmooren
der Niederrheinischen Bucht. Fortschr. Geol. Rheinl. Westfalen, 29: 177-191.
Wagner, R.H., 1985. Upper Stephanian stratigraphy and palaeontology of the Puertollano Basin,
Ciudad Real, Spain. In: M.J. Lemos de Sousa, and R.H. Wagner (Editors), Anais de Faculdade
de Ciencias Universidade de Porto. Suppl. vol. 64, pp. 171-231.
87

Weyland, H., 1956. Die Bedeutung der Kutikular-Analyse fiir die Braunkohlenforschung. Frei-
berger Forschungsh. C, 30: 7-16.
Weyland, H., 1957. Kritische Untersuchungen zur Kutikular-Analyse terti~irer Bliitter, III. Mon-
ocotylen der rheinischen Braunkohle. Paliiontographica B, 103: 34-74.
Weyland, H., 1958. Die Monocotylen des HauptflSzes der Ville. Forschr. Geol. Rheinl. Westfalen,
1 and 2: 527-538.
Winkler, E., 1986. Organic geochemical investigations of brown coal lithotypes. A contribution to
facies analysis of seam banding in the Helmstedt deposit. In: D. Leythaeuser and J. RullkStter,
(Editors), Adv. Org. Geochem. 1985, Part 1, pp. 617-624.
Winston, R.B., 1986. Characteristic features and compaction of plant tissues traced from per-
mineralized peat to coal in Pennsylvanian coals (Desmoinesian) from the Illinois Basin. Int.
J. Coal Geol., 6: 21-41.
WSlk, E., 1935. M~ichtigkeit, Gliederung und Entstehung des niederrheinischen Hauptbraunkoh-
lenfl~zes. Decheniana, 92: 81-162.
Wolf, M. and Wolff-Fischer, E., 1984. Alginit in Humuskohlen karbonischen Alters und sein Ein-
fluss auf die optischen Eigenschaften des begleitenden Vitrinits. GliJckauf Forschungsh., 45:
243-246.
Zagwijn, W.H. and Hager, H., 1987. Correlations of continental and marine Neogene deposits in
the south-eastern Netherlands and the Lower Rhine District. Meded. Werkgr. Tert. Kwart.
Geol., 24: 59-78.

CONTENTS
Abstract 1
Introduction 2
I. Origin of macerals 3
I.I Genesis of huminites/vitrinites 4
1.1.1 Humification and biochemical gelification 4
1.1.2 Macerals of the huminite/vitrinite group 9
1.1.3 Coalification of huminites/vitrinites 13
1.2 Genesis of liptinites 16
1.2.1 Macerals of the liptinite group 17
1.2.2 Coalification of liptinites 22
1.3 Genesis of inertinites 23
1.3.1 Macerals of the inertinite group 23
1.3.2 Coalification of inertinites 29
2. Origin of coal facies (microlithotypes and lithotypes) 29
2.1 Distinction of coal facies 29
2.2 Influence ofdepositional (palaeogeographic)environment 30
2.3 Influence of climate, vegetation and water supply 35
2.3.1 Carboniferous hard coals of Euramerica 35
2.3.2 Gondwana coals (and peats of cool climates) 44
2.3.3 Tertiary brown coals of Germany and Australia 47
2.3.4 Origin of "light layers" in Tertiary brown coals 55
2.4 Vertical succession of coal facies and the influence of nutrient supply (eutrophic- 60
oligotrophic)
2.5 Influence of acidity and of marine and calcareous conditions 67
2.6 Influence of fires 68
3. Conclusions und further work 69
4. References 77

You might also like