You are on page 1of 54

Coal petrology and petrographic analysis

3.1

INTRODUCTION

Close examination of coal in hand specimen generally


shows it to be composed of different layers. Under the
microscope, these layers in turn are seen to be composed
of mixtures of discrete entities, each class of which is
distinguished by having different optical characteristics.
Coal petrology is the study of the origin, composition
and technological behaviour of these different materials,
while the systematic quantification of their proportions
and characteristics under the microscope is sometimes
known as coal petrography.
The different layers or entities occurring in a single
coal may possess quite different physical and chemical
properties, and hence their relative abundance and
manner of admixture is vital in determining the overall
characteristics of a coal seam or mined coal product.
Coal petrography has been widely applied to the
selection and blending of coals for production of
metallurgical coke, and is one of the major
considerations in research directed towards coal
liquefaction operations. The techniques of coal petrology
are also used in geological investigations aimed at
assessing the potential of rocks and sedimentary basins
as sources of petroleum.
3.2 MEGASCOPICALLY RECOGNIZABLE
CONSTITUENTS
The petrology of coal may be studied at either a
megascopic or a microscopic scale. From a megascopic
point of view, coal may be classified into two broad
groups, the banded or humic coals and the non-banded
(massive) or sapropelic coals. The humic coals are
visibly stratified, consisting of layers or bands of organic
material of varying appearance, with individual layers
usually no more than a few centimetres in thickness.
Such coals are derived from a heterogeneous mixture of
a wide range of plant debris. The sapropelic coals, on the
other hand, are homogeneous, tough materials, often
displaying a marked conchoidal fracture. They are made
up of specific kinds of fine grained organic matter,
notably masses of spores or algal material.
3.2.1

Lithotypes in banded bituminous coals

A large part of the terminology used in coal petrology is


derived from the work of Stopes (1919) in recognizing
four basic ingredients of banded bituminous coal that
can be distinguished in hand specimens. These
constituents, regarded in current usage as lithotypes,
were identified by Stopes, a palaeobotanist, as follows:
(a) Vitrain (L. vitrurn, glass)
Vitrain is the black, glassy, vitreous material that is
probably the most striking component of bituminous
coals. It occurs as thin bands, commonly less than 6 or 8
mm in thickness and is usually very closely jointed.
Vitrain tends to be more brittle than other megascopic
coal constituents, often breaking with a conchoidal
fraction (Fig. 3.1a).
(b) Clarain (L. clarus, bright)
This lithotype is represented by bright to semi-bright
bands of finely laminated coal. Clarain generally
exhibits an overall silky lustre, and commonly contains
fine vitrain bands alternating with a duller attrital
groundmass (Fig. 3.1b).
(c) Durain (L. durus, hard)
Durain occurs as grey to black bands with a dull to
slightly greasy lustre. The material is relatively hard
compared to other lithotypes, and tends to break into
large, blocky fragments. Durain may sometimes be
confused with impure coal or carbonaceous shale, which
are also often dull and hard, but it can be distinguished
by its lower density (Fig. 3.1c).
(d) Fusain (L. fusus, a spindle)
In French, the word fusain means charcoal, which at one
time was made from the wood of the spindle tree. The
suffix ^ain was adopted for the other lithotypes. Where
it is unmineralized, fusain is a soft, friable material that
closely resembles the charcoal from which it takes its
name. Soft, or unmineralized, fusain easily disintegrates
into a black fibrous powder, but hard fusain,
impregnated with mineral matter, may be found in some
coals as well. Fusain usually occurs as thin lenses,
seldom more than a few millimetres thick, and is only a
very minor constituent of most bituminous coal seams
on a volumetric basis (Fig. 3.Id).
The terms duroclarain (Cady 1942) and clarodurain

Coal Petrology and Petrographic Analysis

(a)

(c)

dark in colour with a dull to greasy lustre and typically


display a marked conchoidal fracture (Fig. 3.2a).
Sapropelic coals may occur as layers or plies within
seams of banded or humic coal, often at the roof. They
also occur as seams made up mainly of homogeneous,
non-banded material in their own right.
The two major types of sapropelic coal are cannel
coal5, composed largely of spores or fine organoclastic
detritus, and boghead coaP composed largely of algal
material. These are, however, effectively end-members
of a range of materials representing mixtures of these
two types of components, and transitional or
intermediate forms such as cannel- boghead and
boghead-cannel may be recognized as well. Bogheads
may grade laterally or vertically into oil shales.
When viewed under the microscope, cannel coal can
be distinguished from boghead coal both by the
abundance of spores and by the presence of a regular
microstratification. The materials are, however, almost
impossible to distinguish from each other in hand
specimen.
3.2.3 Field description of coal seams

(b)

(d)

Fig. 3.1 Lithotypes in banded bituminous coal, (a)


Vitrain in polished surface. Note that the three
bands of vitrain at the top of the block are more
highly jointed (cleated) than the rest of the coal
(x<3.8). (b) Clarain in polished surface. The lower
three quarters of the block are composed mainly of
clarain; the upper quarter consists of two durain
bands (grey) and a vitrain band (black)
(x0.8). (c) Durain in hand specimen (x0.3).
(d) Fusain in bedding surface of hand specimen
(x0.2).

have been added to this list by some workers to extend


the number of terms available for megascopic coal
description. They represent material that is intermediate
in character between clarain and durain.
Ul OLl l lVl lVl l Wt V* Wll

KricrVit

(vitrain and clarain) and dull components


components (durain and fusain) is
most apparent in coals of bituminous rank. In
anthracites, however, the lithotypes all tend to develop
a relatively bright lustre, and the contrast between them
decreases.
3.2.2. Lithotypes in sapropelic bituminous coals
Unlike the banded or humic coals, which were
deposited as peats made up of large to small fragments
of plant debris, the sapropelic coals represent
accumulations of fine organic mud containing concentrations of algae or spore remains. Sapropelic coals are
characteristically fine grained, faintly bedded to
homogeneous, massive materials. They are generally

Although the terms vitrain, clarain, durain and fusain


are widely used for the description of individual
specimens or discrete horizons within a coal seam, a
number of difficulties arise with their employment in
routine logging of seam sections in bore cores or coal
exposures (Davis 1978a). One major problem is that
these four terms refer to quite different kinds of units
within the coal. Vitrain bands represent coalified
fragments of wood or bark and are generally no larger
than an individual tree trunk in size. Clarain and durain,
on the other hand, are usually more extensive units,
each possibly representing a aepositional environment
within the peat swamp. In a rather exaggerated analogy,
the vitrain bands might be compared to an individual
pebble, while the clarains and durains are like the
conglomerate in which the pebbles occur.
Another disadvantage associated with use of Stopes
terminology lies in the fact that the individual layers or
lenses of the four lithotypes may be very thin, usually
only some millimetres in thickness. Even with the
accepted minimum layer thicknesses of different
countries (3-10 mm), strict application of the Stopes
terms can result in extremely detailed descriptions.
Many field descriptions of coal seams, however,
especially those of very thick seams, are based, for
expediency, on sub-division into a relatively small
number of megascopically distinct units. The

75

76

Chapter 3

Stopes system was not designed for, and is not


particularly effective in, this kind of usage.
Recognizing these and other difficulties inherent in
the Stopes terminology of coal lithotypes, Schopf
(1960) established a descriptive system for use by the
U.S. Geological Survey, and this has been subsequently
employed by many others for field use. Schopf s terms
are out in Table 3.1. In summary, three constituents are
described, namely vitrain, fusain and attrital coal.
The first two are regarded as larger clastic units
occurring within a matrix of finely divided attrital coal.
The thickness and concentration of the vitrain and
fusain are described in terms which are quantitatively
precise, and the attrital coal is described as having one
of five levels of lustre, ranging from bright to dull.
Another alternative has been to describe the coal
with reference only to its relative brightness.
Employing terms used in the German coal industry,
Diessel (1965) has described coal seams as being
composed of megascopically distinct layers of the
following types of materials:
(a) bright coal;
^
(b) banded bright coal; ^
(c) banded coal;
(d) banded dull coal;
(e) dull coal.
There are many similar descriptive systems in use
around the world, and all can be used on as broad or as
narrow a scale as necessary or convenient for the
particular purpose for which the description is being
prepared. A seam may be described on a centimetre
scale (e.g. for research purposes), or on a metre scale,
as desired.
TABLE 3.1 U.S. Geological Survey terms for megascopic description
of banded bituminous coal (Schopf 1960).
Vitrain
Fusain
f bright
moderately bright
midlustrous

Attrital coal J

thick bands
very thick bands

5-50
>50

Concentration classes (%)


sparse
<15
moderate
15-30
abundant
dominant

3.2.4

30-60
>60

A combination of the bright-dull system and the


Schopf system has been devised by the Coal Research
Section of The Pennsylvania State University for
effective seam descriptions at a megascopic level, and

Impure coals

Where the coal contains a significant amount of mineral


matter, its overall density and ash yield increase
significantly. Although the distinction between clean
coal and impure coal is generally based on the
economic constraints of mining, marketing and use,
most materials regarded in the latter category have an
ash yield greater than 25%, and sink when placed in a
liquid with a relative density of 1.60. However, where
the coal has more than about 40-50% ash, it is usually
more correctly described in non-coal terms, for
example as a carbonaceous shale or a coaly shale.
The mineral impurity in the coal may be in the form
of discrete bands, streaks or layers interbedded with the
organic constituents, or it may occur as nodules or as
fracture infillings. It may be made up of clay or shaley
material, or of pyrite, siderite or calcite. Coal with a
significant amount of fine clay disseminated throughout
the organic matter, rather than in discrete layers, is
often described as stony coal or, in the U.S.A., as Tone
coal (Fig. 3.2b). Such material is characterized by a
dull appearance and, commonly, a grey, rather than a
black, coloured streak. An indication of the terms that
may be used to describe impure coals is given in Table
3.3.
3.2.5

moderately dull
1 dull
Thickness classes (mm)
thin bands
0.5-2
medium bands
2-5

this is summarized in Table 3.2. Fusain bands or lenses


thicker than about 5 mm are recorded separately, as are
non-coal bands or partings, for which conventional
sedimentary rock terms may be used. Another system,
used extensively to describe Australian bituminous
coals in outcrops, mine exposures and drill cores, is
discussed in Section 6.5.6. Discussions of the
preparation and use of megascopic coal seam logs in
field studies are also given by Schopf (1960) and
Dutcher (1978).

Megascopic features of low-rank coals

For most practical purposes, the distinction between


hard coals, of bituminous rank or higher, and soft or
low rank coals is based on the specific energy and other
chemical properties of the materials concerned (Section
2.10.2). In European terminology, the low rank
materials are generally described as brown coaf,
whereas in the U.S.A. and elsewhere they are classed
either as lignite or sub-bituminous coal, depending
mainly on their chemical characteristics. Although the
term brown coal is, strictly speaking, applicable to a
wider range of material than lignite, it is also used as
a synonym for lignite in many contexts.
Lignite is a dull, soft, earthy material, ranging from

TABLE 3.2 Descriptive terms for coals (The Pennsylvania State University).
Coarsely banded coals
bright (banded)* coal (> 90 V)f bright interbanded
with dull coal (65-90 V)
interbanded dull and bright coal:): (35-65 V)

Attrital coals

bright (attrital) coal* (> 90 V)

midlustrous coal

fusain

dull interbanded with bright coal (10-35 V) dull


coal (< 10 V) sapropelic (non-banded) coal

* In practice, only one category of bright coal has been employed.


] Numbers in parentheses refer to approximate percentages of vitrain bands and streaks.
^ Contains approximately equal proportion of bright coal and dull attrital coal.
Midlustrous attrital coal may contain many fine vitrain streaks, but obvious vitrain bands should be less than 10%.

brown to black in colour. It may occur in a massive


sapropelic form or, more commonly, as a humic material
with recognizable wood, leaf and other plant

banding, but some are almost massive in hand specimen.


Both lignites and subbituminous coals also tend to crack
and fall apart on drying-out with exposure, a process
known generally as slacking.

Fig. 3.2
(a) Sapropelic coal. Note the massive structure,
faint bedding and conchoidal fracture (*0.5).
(b) Bone coal in broken surface of drill core. Note i
dull appearance and lack of obvious mineral
=
partings (x0.5). (c) Macroscopically visible plant
structure in polished surface of vitrain band from Indiana high volatile C bituminous coal (x2.5).

fragments of various sizes set in a finer grained organic


matrix. A sub-bituminous coal, on the other hand, is
generally black in colour and relatively hard, ranging from
dull to vitreous in lustre. Humic varieties may display

Low rank coals, especially lignites, are very difficult to


describe by megascopic examination. The colour and
lustre may vary with different degrees of dessication, and

no generally accepted classification of lithotypes has yet


been devised. A classification of brown coal lithotypes is
currently being prepared by the International Committee
for Coal Petrology (I.C.C.P.), with the different categories
distinguished on the basis of colour and texture rather than
chemical composition. The terms that have been proposed
in this classification are:
(a) xylitic (woody);
(b) attritic;
(c) fusitic;
(d) mineralized.
3.2.6
Applications of megascopic seam
descriptions
Detailed megascopic logging of coal seam sections is a
time consuming and often difficult task. Conditions of
lighting, surface moisture and accessibility may pose
problems for the geologist working at a mine face, and
even in the somewhat less harried situation of bore core
logging, the friability or degree of oxidation of the coal
may impede the logging process.

TABLE 3.3 Descriptive terms for impure coals.


impure coal, undifferentiated bone coal*
impure coal, with shale bands/streaks (alternative name: shaley coal)
impure coal, with pyrite layer(s)/band(s)/nodule(s)/etc. impure coal, with
carbonate band(s)/nodule(s)/etc.
* Bone coal is an American miners term for describing coal in
which a significant amount of fine clay is disseminated through
the coal rather than occurring in obvious shale partings. Consequently, bone coal has a dull appearance and a grey streak.

Fresh, clean surfaces are needed for good descriptions


and cores should be carefully broken open, while in situ
seam exposures should be cleared of any weathered
debris or fire retardant stone dust (as used in
underground mines) prior to logging.
A geologist inspecting a seam exposed at a mine face
or in an exploratory bore core has a unique opportunity
to record the exact structural characteristics of that seam
before it is destroyed either by mining or by the analysis
process. Even though, in some instances, there may be a
lack of consistency between results from different
workers, experienced personnel are often able to obtain a
considerable amount of useful data that may be evaluated
in conjunction with other ply-by-ply analyses for very
little additional cost.
Some coal seams have characteristic lithotype
profiles that remain more or less constant over wide
areas, or contain marker beds of distinctive character that
can be recognized in many parts of the field. A detailed
megascopic log, perhaps expressed in graphic form, may
be very useful in correlating the individual seams in a
coal-bearing succession, and in the interpretation of
displacements in faulted strata. Durains and fusains have
proved especially useful in this regard (Cameron 1971;
Austin & Davis 1979) as have individual bands of noncoal material (e.g. Mackowsky 1968a).
Megascopic profiles or logs of the seam are also
potential sources of information on variations in coal
quality that may affect the mining, preparation or
utilization of the material. A brightness log based on the
descriptive system of Diessel (1965) has been used, for
example, as a rough guide to the coking potential of
individual seams in Australia (Hawthorne & Tweedale
1967). Although it is not necessarily proper nor wise to
draw inferences on the maceral composition of seams
from such data, it has proved possible to make at least
some correlation with micropetrographic characteristics
in a number of cases (e.g. Diessel 1965; Cameron 1978;
Marchioni 1980).
3.3 MICROSCOPIC APPEARANCE OF COAL
MACERALS
When viewed under the microscope, coal is seen to consist
of particles and bands of different kinds of carbonaceous

material. These discrete entities represent the coalified


remains of the various plant tissues or plant-derived
substances that existed" at the time of peat formation. They
are distinguished from each other on the basis of their
morphology, hardness and optical properties, and also
exhibit differences with respect to their chemical
characteristics or technological behaviour in coal
utilization.
The different entities that make up a coal in this way
are known as macerals, a term coined by Stopes (1935)
as an analogy to the minerals of inorganic rocks. Several
of the maceral names suggested by Stopes were adopted at
the 1935 Heerlen Congress, and as a consequence the
universally adopted classification of these components is
referred to as the Stopes-Heerlen system.
Much of the early work on the micropetrology of coal,
including Stopes original classification, was carried out
using thin sections viewed in transmitted light. However,
Stach (1927) pioneered the use of polished sections studied
in reflected light, under oil immersion. This development
paved the way for more efficient and consistent practice of
coal petrography. Although the techniques tend to
complement each other to some extent, almost all routine
petrological work at present is based on polished section
methods. The criteria by which the various macerals are
identified depend mainly on their appearance and optical
characteristics under reflected light illumination.
All maceral names in the Stopes-Heerlen system have
the suffix ending finite. Two of these macerals are entities
which, when observed under the microscope, are seen to
make up the bulk of the lithotypes vitrain and fusain, and
they have been named vitrinite and fusinite, respectively.
The other lithotypes, namely clarain and durain, are
generally composed of a heterogeneous mixture of
macerals of different kinds.
Coal macerals in the Stopes-Heerlen system are
classified into three groups on the basis of their physical
appearance, chemical characteristics and biological
affinities (Table 3.4). The appearance of the members of
each group, however, changes with advance in coal rank,
and the distinctions between the groups that are easily seen
in low rank bituminous coal may, for example, be lost in
semi-anthracite. In some cases, particularly with brown
coals (lignites and sub-bituminous materials), different
names may be used for macerals of similar origin to
emphasize these characteristics.

The appearance of the different macrais in transmitted


light is the antithesis of that in reflected light, since a
material that is relatively transparent is inherently a poor
reflector. Coal for thin section study must be ground to a
thickness of about 10/un (Section 3.7.1), one-third that
needed for other rocks, and this makes the preparation
process a highly skilled and time-consuming operation.
A comprehensive summary of the features that
characterize the various members of the macrai groups is
given in the International Handbook of Coal Petrography
published by the International Commission on Coal
Petrology (I.C.C.P. 1963, 1971, 1976). This all-embracing
reference gives a complete definition for each macrai,
including the derivation of its name and a list of any
synonyms, as well as its morphography,
nrnnprfi AC rhprmrol phoro/^tpricti^c or-i rl 1WCU
IOLIVO

VllVllllVUl VliUiaVLCl

anu

TABLE 3.4 Stopes-Heerlen classification of maceral groups, macerals and


submacerals of hard coals.
Maceral group
telocollinite

Maceral

^ telinite V
vitrinite

Submaceral
gelocollinite
desmocollinite
corpocollinite

collinite

liptinite
(exinite)

/ sporinite
cutinite
suberinite
resinite
alginite

liptodetrinite

botanic affinities. It also includes theories on the mode of


origin of the various macrais, and an indication of their
respective significance to commercial processes.
3.3.1 The vitrinite group
Vitrinite is the preponderant macrai in most coals. It
originates mainly from the preservation of the stems, roots
and leaves of plants, including the wood, periderm and leaf
mesophyll tissues and some cell fillings, all with varying
degrees of mechanical degradation but relatively minor
oxidative alteration. Vitrinite is also formed from colloidal
humic gels. Plant cell structure may often be observable
under the microscope (Fig. 3.3a), and sometimes even by
the naked eye (Fig. 3.2c) in the larger vitrinite occurrences.
In thin section, vitrinite is moderately transparent and
appears coloured in various shades of red, orange and
brown. In reflected light under oil immersion, however, it
appears medium grey in contrast to the darker liptinite and
lighter inertinite macrais (Fig. 3.3b).
Vitrinite occurring in vitrain bands of about 3-12 mm
in thickness represents the mummified and coalified
products of larger roots, bark and stems of plants. This
type, or sub-maceral of vitrinite is known as telocollinite.
Where a distinct cell structure is visible the term telinite
has been used, although some authors prefer to restrict this
particular term to the cell wall material only. In some
instances it is possible to identify the actual plant genus
from which such vitrinite was derived.
Apart from the relatively thick bands derived from
woody components, other vitrinite originates from

fluorinite
bituminite V
exudatinite

inertinite

s fusinite
semifusinite )
macrinite \
micrinite I
sclerotinite ^
inertodetrinite

smaller plant tissues, such as grass and reeds, and tissues


that have been degraded into finer sized fragments. This
material often occurs in attrital admixtures with other
macerals and minerals, and is known as desmocollinite.
Vitrinite of this type is a typical constituent of clarains.
The individual particles have lost much of their integrity,
and are cemented together as a result of glification.
Desmocollinite often appears darker in reflected light than
the telocollinite in the same coal (Fig. 3.3c) and this may
be due in part to derivation from a more cellulose-poor
type of plant material. However, the fact that
desmocollinite layers sometimes fluoresce in blue light
illumination (Section 3.3.4) suggests that some lipid
substances may have been absorbed into the material.
The breakdown of ligno-cellulosic tissues by bacterial,
fungal or chemical action produces a colloidal gel, of
which the dopplerite found in peats is an example. Where
it is possible to recognize that cell lumens of vitrinite or
fusinite, or cracks and other cavities, have been filled with
a substance which must have precipitated as a gel, the type
of vitrinite that forms the infilling material is known as
gelocollinite.
Yet another type of vitrinite, called corpocollinite, is
represented by circular, elliptical or rod-shaped

bodies, occurring either in isolation or as cell fillings. This


is the high-rank equivalent of corpohuminite, a maceral
of low rank coals described more fully in Section 3.3.5.
3.3.2

The inertinite group

The macerals of the inertinite group are characterized by a


high reflectance, and have higher carbon and lower
hydrogen contents than other macerals in coals of
equivalent rank. They are essentially opaque in thin
section, but appear white or light grey in polished section.
The inertinite macerals are mainly derived from the same
basic types of organic matter as vitrinite, but owe their
properties to oxidation of
.t _

____. 1 1

r* -> QilfllT ctOtTP

tnose materials ai an wn)

tViP mills

t-v ~

formation.

The name inertinite was selected as a group name to


imply the relative inertness of these macerals in
technological processing (e.g. coke manufacture) in
comparison to the members of the other two maceral
groups. However, one of these macerals (semifusinite)
sometimes displays sufficient reactivity in processes,
including carbonization, that the term semi-inert has been
used by some workers to describe such materials.
Microscopic observations made of a heated coal sample by
Nandi and Montgomery (1967) also suggest the possibility
that another maceral (micrinite) is far less inert than had
originally been thought, but this conclusion has been
disputed by other investigators in the field.
(a) Fusinite and semifusinite
Although, in certain circumstances, plant materials are
transformed into vitrinite, in other circumstances identical
materials are seen to have undergone a radically different
process, giving rise to a brittle, opaque maceral called
fusinite. A plant origin is recognizable for this material
from the well-preserved cell structure, and. fine detail of
the cell walls may be commonly seen (Fig. 3.3d). The cell
lumens of fusinite may be open cavities or they may be
infilled with minerals (carbonates, pyrite, clays) or with
gelocollinite. Where a lack of infilling has resulted in the
eventual collapse of the brittle cell walls due to
compression, a broken bogen structure may be developed
(Fig. 3.3b).
Fusinite is opaque in thin section and in reflected
light is the most highly reflecting maceral in most ranks
of coal. It generally appears white, sometimes even
yellowish in polished section studies. Where there is the
same detail of preservation of cell

structure, but the level of reflectance is grey and


intermediate between that of the vitrinite and fusinite in
the same coal, the maceral, which also usually appears
brown in thin section, is known as semifusinite (Fig. 3.3b
and d). For convenience, some petrologists have set an
arbitary reflectance threshold of 2.0% to distinguish
between fusinite and semifusinite in petrographic analysis.
^
Organic cell fillings which were emplaced early in the
coals history may also be subject to the fusiniza- tion
process. Thus, oval resin bodies (Section 3.3.3) may be
converted into masses which, because of their shape, high
reflectance and the presence of cavities (Fig. 3.3e), can be
easily mistaken for fungal sclerotia (Section 3.3.2e). These
bodies may be called fusinite, semifusinite or macrinite
(Taylor & Cook 1962), sclerotinite (I.C.C.P. 1971) or
scierotioids

1 Q5 1 Y serretinn

-----------

resino-sclerotinite (Stach 1966; Lyons et al 1982). Several


authors, including Kosanke and Harrison (1957) and Lyons
et al (1982), believe that many of the resin rodlets from
which resino-sclerotinite was derived were probably
/if marlnllnCilfl 01*1^171.

(b) Macrinite
Although it has a similar level of reflectance to fusinite,
macrinite occurs most often as small, rounded but
irregularly shaped bodies without cell structure, usually
ranging from 10 to 40 /m in diameter. Macrinite often
appears homogeneous, but it is also apparent that some
macrinite has originated through the fusinization of
gelified tissues. Certain durains contain relatively large
amounts of macrinite in association with sporinite (Fig.
3.4a).

(c) Micrinite
Micrinite is another highly reflecting coal maceral. It
occurs as very small, rounded grains, rarely more than a
few microns in size (Fig. 3.4b). The grains are in fact
discrete particles, but they tend to form accumulations,
either as lenses or layers or in cell lumens. In transmitted
light, micrinite is opaque and, in large aggregates it may be
difficult to distinguish from fusinite or macrinite.
Micrinite is a ubiquitous component of bituminous
coal, particularly in durains and sapropelic coals, but it is
rarely present in large quantities. It commonly occurs in
association with sporinite and other liptinite group
macerals. Micrinite occurs far less commonly in lignites
and sub-bituminous coals, a fact which supports the
hypothesis that it is generated as a result

Fig. 3.3 Reflected-light photomicrographs of coal (under oil immersion), (a) Vitrinite showing the cell structure of lycopod
periderm (x270). (From Davis et al 1976.) (b) Vitrinite appears medium grey in contrast to the light grey fusinite and semifusinite and the dark grey
liptinite macerals sporinite and cutinite (centre). The semifusinite seen at the top of the photomicrograph has a lower reflectance and less distinct cell
wall outlines than the fusinite, which displays bogen structure.Pyrite, occurring mainly as small euhedral crystals in the centre, appears white (x480).
(c) The telocollinite in coarse bands at the top left and bottom right has a higher reflectance than the fine desmocollinite occurring in the central attrital
layer with fragments of other minerals (x 480). (d) Fusinite (white) and semifusinite (light grey), both showing well defined plant cell structure (x480).
(From Davis et al 1976.) (e) Resino-sclerotinite or sclerotioid in Permian high volatile bituminous coal from southern Africa. Note the deep notch
which can be a feature of these bodies (x480). (f) Sclerotinite (white) derived from fungal sclerotia with isolated resinite bodies (dark, oval) in a Late
Eocene sub-bituminous coal from Washington (x330).

Coal Petrology and Petrographic Analysis

iHS~

of the chemical changes that accompany coalification


(Stach 1968; Teichmiiller 1974a). However, it can also
be observed in cell fillings of very low rank coals, in
which case it may represent an end product of the decay
of woody tissues.

(d) Inertodetrinite

Material derived from the outer layer, or exine of spores and


pollens was originally referred to by Stopes (1935) as
exinite. However, the meaning of the term was expanded to
include coalified cuticular (leaf cuticle) material (Jongmans
et al 1935), and eventually macerals derived from algae and
resin bodies as well. The term liptinite is more appropriate
to encompass all of these macerals, together with the more
recently identified components suberinite, liptodetrinite,
fluorinite, bituminite and exudatinite.
The macerals of the liptinite (or exinite) group appear
darker in reflected light than the associated vitrinite, and are
generally pale in colour under transmitted illumination. They
represent a diverse assemblage of small organic particles,
characterized, particularly in low rank coals, by a high
hydrogen content and a high proportion of volatile matter.
The optical properties and chemical characteristics, however,
change significantly as rank advances from high volatile to
medium volatile bituminous coal, and many of the
distinguishing features are lost in the higher rank materials.

Inertodetrinite is composed of broken fragments of


inertinite macerals (Fig. 3.4f, 3.5a and d). According to
I.C.C.P. (1971) a fragment of fusinite or semifusinite
which has less than one complete cell should be classed
as inertodetrinite. However, when a petrographic study
is being conducted for the purpose of interpreting the
depositional environment of the coal, it may be
reasonable to identify as inertodetrinite any small piece
of fusinite which has been detached from a larger mass
and deposited away from the immediate vicinity of
other similar fragments in any attrital coal band.
Discrete fragments of fusinite (or semifusinite) which
represent less than an entire cell may also be observed
in the resin binder of polished section grain mounts.
Such particles are, however, better classed as fusinite
(or semifusinite) because they are more likely to have The most common of the liptinite macerals in humic coals
been detached from a fusain lens than from an attrital are the coalified remains of spore and pollen exines called
coal lithotype during sample preparation.
sporinite. The typical appearance of these materials is that
of a flattened oval shaped particle with a central cavity or
line indicating that the inner layer and protoplasm of the
structure has decayed (Fig. 3.4a). The size of these particles
(e) Sclerotinite
ranges from about five to several hundred microns, while the
shape, including thickness and ornamentation, is extremely
With the exception of most micrinite, all of the macerals
varied.
described above have probably been derived, in one
In thin section, the colour of sporinite is yellow or orange
way or another, from the ligno-cellulosic tissues of in high volatile coals. Like the other low- reflecting macerals
plants. Sclerotinite is another maceral that is opaque in described below, sporinite can display a high level of
thin section and highly reflecting, but this material fluorescence under blue or ultraviolet irradiation (Section
originated instead from fungal remains. The high 3.3.4). Sporinite is abundant in some durains and clarains
reflectance in this case is due to the presence of dark and is the characteristic constituent of cannel coals.
pigment (melanin) rather than the main component of
In coals with a volatile matter yield of less than 29.5%
such bodies, the polymer chitin (Stach et al 1975).
(d.a.f.), a point known as the coalification jump, the
Material classed as sclerotinite includes all coalified reflectance curve of sporinite (Section 3.6.4) begins to
sclerotia, fungal spores, hypae and plectenchyme. True converge with that of vitrinite. The two curves are coincident
sclerotinite is a ubiquitous component of Tertiary coals, at a volatile matter of about 21% and a reflectance of about
occurring as rounded spores and sclerotia some tens of 1.50%, and it is not possible to distinguish between the two
microns in diameter and having one or more cell macerals under oil immersion in these circumstances. In high
cavities (Fig. 3.3f). Bodies which appear similar to rank coals, optical differentiation of these macerals may be
sclerotinite may be observed in Carboniferous coals improved to some extent by using methylene iodide
(Fig. 3.3e), but these are usually oxidized or fusinized immersion techniques.
resin rodlets (Section 3.3.2a). The cavities of these
bodies can generally be discerned as vesicles rather than (b) Cutinite
a regular cell structure (Stach et al 1975; Lyons et al
The waxy cuticular coatings on certain aerial epidermal
1982).
tissues, notably leaves, are preserved in coal as cutinite. The
3.3.3 The liptinite (or exinite) group
functions in life of this cutin, an insoluble polymer, are to
prevent the delicate tissues from rapid desiccation, and to

10

Chapter 3

give them physical support and protection from biochemical


agencies. An entire organ with a double layer of cutinite may
be preserved as in Fig. 3.4c, where the inner mesophyll of a
leaf has been converted into vitrinite. Frequently, however,
the resistant cutinite is all that remains of the parent plant
structure.
Cutinite usually occurs as very thin elongate bodies. A
series of cusps or teeth may be seen on one side of the cuticle
to indicate where it originally extended between the radial
walls of epidermal cells. Cutinite can also be relatively thick
(Fig. 3.4d), indicative in some instances of a dry, sunny
climate. Thick cuticles may result from extensive
cutinization of the epidermal tissues or the build-up of many
layers. Unusually rich accumulations of cutinite occur in rare
but well-known occurrences of paper or leaf coals (Auerbach
& Trautschold 1860; Guennel & Neavel 1959; Cook &
Taylor, 1963).
(c) Suberinite
The corky cells of the plants that contributed to coals,
particularly those of the Tertiary, contained the waxy
polymer suberin, which is similar in many ways to the cutin
of cutinites. Both are mixtures of substances, mainly fatty
acids, and are consequently impervious to water. The
functions of suberin are similar to those of cuticle, but
suberin is deposited within the cell walls rather than outside
them.
In reflected light, suberinite stands out as the dark walls
of the relatively large cork cells. These are filled with more
highly reflecting materials (Fig. 3.4e).
(d) Resinite
Coalified resins occur in coals as more or less oval or rodshaped bodies at their original sites of deposition in cell
lumens (these are primary resinite). Other material
(secondary resinite) clearly has been mobilized at some
stage, and occurs as veins or cleat fillings and in pods or
cavities such as fusinite cell lumens. Crelling and Dutcher
(1980) showed that secondary resinite can have quite
different fluorescence properties to either primary resinite or
exudatinite (see below) in the same coal. A third and
common mode of occurrence for resinite, however, is as
bodies which have been weathered out of the other plant
tissues and incorporated as transported particles within
attrital coal layers (Fig. 3.3f, 3.4f).

The colour in thin section and the reflectance resinite


can vary widely, even within a single coal. It can approach
that of the vitrinite with which it may be intimately
associated, so that microscopic differentiation of the two
becomes difficult. Polished sections sometimes show
internal reflections, while in transmitted light, resinite may
be shades of yellow, orange, red or brown. Oxidized
resinite bodies can have rims which are relatively higher in
reflectance than the interiors. Resino-sclerotinite is
fusinized resin with a high reflectance and is often
vacuolated (Section 3.3.2a).
Some discrete resin-rich bands are also encountered in
coals. These may be layers of secondary resinite or attrital
accumulations of weathered resin bodies. Tertiary and
Cretaceous coals contain relatively large amounts of
resinite because of the contribution from conifers. The
resinite may be concentrated from low rank coals for
commerical use either by a special coal preparation process
or, because it is more soluble in benzene than other liptinite
macerals, by a solventextraction technique.
(e) Alginite
Alginite represents the coalified remains of algae. Such
material is rare but not unknown in humic coals, and
abundant in the variety of sapropelic coal known as
boghead coal or torbanite. It is also abundant in some oil
shales.
The individual algal colonies are oval in shape, often
with scalloped outlines, and these help to differentiate
alginite from sporinite in reflected light (Fig. 3.5a).
Alginite also has a somewhat lower reflectance than other
liptinite macerals. In thin section alginite has a pale straw
colour in low rank coals, but is somewhat orange at higher
ranks.
Examination in blue or ultraviolet light reveals more of
the details of the colonial structure of alginite than can be
seen in ordinary white light (Fig. 3.5b). Alginite has a high
intensity of greenish-white or yellow fluorescence in low
rank materials, but is darker in fluorescence colour at
greater levels of organic maturation.
Two of the genera of algae that may be identified in
coals (Pila and Reinschia) have been related to the living
species Botryococcus braunii. A.C. Cook (personal
communication) has suggested that much of the material in
cannel coals that has been described

Fig. 3.4 Reflected-light photomicrographs of coal (under oil immersion), (a) Sporinite (dark) and macrinite (light) in durain
from a Carboniferous high volatile bituminous coal, Kentucky (x480). (b) Micrinite (white) as lenses in vitrinite (grey) and as a thick layer (x480). (From
Davis et al 1976.) (c) Thin layers of dark grey cutinite enclose light grey vitrinite derived from leaf mesophyll. Jurassic Maghara seam, Egypt (x205). (d)
Thick cutinite in leaf coal from Leping Country, Jiang Xi Province, China (x480). (e) Thin layers of dark suberinite in a Palaeocene sub-bituminous coal
from Wyoming.
The thick black lines in the bedding plane are desiccation cracks (x480). (f) Lenses of resinite. Note the small vertical cracks joining resinite occurrences;
these represent secondary mobilization of resinite or exudatinite (x480).

as bituminite (Section 3.3.4) may very well be the material


described by Hutton et al (1980) as alginite B\ a lamellar
alginite with affinities to the genus Pediastrum.
(f) Liptodetrinite
Liptodetrinite, the member of the liptinite group equivalent
to inertodetrinite, is composed of fragments of the liptinite
macrais sporinite, cutinite, resinite and alginite (Fig. 3.5a).

3.3.4
The fluorescence of the liptinite
macerals
_______u________________________...

i ne iipimuc liiaeciaid ui cuai nave me piupcny m

displaying a fluorescence when viewed under blue light


irradiation. Descriptions of the apparatus employed for
studies of these autofluorescence characteristics are given
by van Gijzel (1971), Ottenjahn et al (1974) and I.C.C.P.
(1976). High- pressure mercury or xenon lamps are used for
illumination in qualitative fluorescence microscopy, with a
blue or ultraviolet excitation filter to remove much of the
visible light. A blue filter, for example, with a maximum
transmission wavelength of about 400 nm may be used in
conjunction with a red suppression filter to achieve the
desired results. In some microscopes, the vertical
illuminator contains a dichroic beam-splitting mirror which
reflects light of below 510 nm. A barrier filter removes the
reflected excitation rays and protects the eyes from
exposure. For the blue light assemblage just described, a
barrier filter with peak transmittance at 530 nm is generally
most suitable.
As a result of fluorescence studies, Teichmtiller
t i r \ n A - i _ \ ______i

_____________ A

/-\C\HH\

U ______

anu iciLiiiiiunci aim wun \ i y / / ; nave

named a number of materials present in coal which had not


been previously identified or distinguished from other coal
constituents. The new maceral names proposed include
fluorinite, bituminite and exudatinite. The optical properties
of these macerals, summarized in Table 3.5, reveal that these
are quite different substances from the better known liptinite
macerals, namely sporinite, cutinite, resinite and alginite.
(a) Fluorinite
Because of its black appearance, sometimes with internal
reflections, in reflected white light with oil immersion, this
pure organic substance could be mistaken for lenses or
layers of clay minerals in the coal. However, Teichmtiller
(1974) has noted that this
(b) Bituminite
Bituminite is the most frequently occurring of the three new
liptinite macerals. It is seen as irregularly- shaped shreds,
wisps and layers of a material with reflectance intermediate

material has a strong yellow fluorescence when irradiated


with blue light. Fluorinite has a maximum fluorescent
intensity at a lower wavelength than other liptinite group
macerals at the same level of rank. Consequently, the
red/green quotient, that is, the ratio of the relative intensity
at 640 nm to the relative intensity at 500 nm, is always
lower than those of the other macerals.
Fluorinite is believed to originate from plant oils and
fats. It is a ubiquitous maceral of European coals
(Teichmtiller, personal communication), but is rarer in the
coals of the eastern U.S.A.
between those of vitrinite and sporinite. Bituminite may
even form the groundmass of some durains and sapropelic
coals. Previously, it had often been identified as
liptodetrinite, the fragmented form of liptinite, but
Teichmuller (1974a) notes that bituminite has a fluorescence
property that clearly distinguishes it from all other macerals,
namely a fluorescence intensity that increases by as much as
200% after a 30 min period of irradiation. The usual
fluorescence colours are orange to brown, and the maximum
fluorescence of bituminite occurs at a longer wavelength
than does that of other macerals.
Teichmuller has suggested that bituminite represents the
decomposition products of algae, bacterial lipids and animal
proteins. However, Hutton et al (1980) believe that some of
the material that has been called bituminite is really alginite
B (Section 3.3.3.e). Teichmuller (1974a) has also suggested
that the generation of some micrinite in coals results from
A : _______ r u : . . ________
U1C UiagCllCM fc Ul UllUllllilllC.

(c) Exudatinite
The mode of occurrence of exudatinite indicates that it is a
secondary maceral which has been soft and mobile at some
stage during the coalification process. Exudatinite appears
black under reflected light in oil immersion, and it is only
by the use of a dry objective or fluorescence illumination
that what appeared to be empty cracks and cavities are
sometimes seen to be filled with a material that typically has
an orange to yellow fluorescence in blue light irradiation.
The cell lumens of fusinite or semifusinite and the
chambers of sclerotinite frequently provide the cavities in
which exudatinite may occur. Some cracks containing
exudatinite may also be joined to primary

TABLE 3.5 Origin and properties of new liptinite group macerals. (Modified after Teichmtiller 1974a.)
Appearance in
reflected light (oil)
Macrai
fluorinite

bituminite

exudatinite

lenses

Form

streaks and
groundmass

cavity
fillings

Black,
occasional
internal
reflections
as

Reflectance
intermediate
between
vitrinite &
sporinite
black

Intensity
strong

weak

variable

Colour
brilliant
yellow

lipinite macerals such as cutinite and resinite. The


reflectance and fluorescence intensity of the exudatinite,
however, are respectively higher and lower than those of the
primary liptinite occurrences.
Teichmtiller (1974a) has observed that the maximum
fluorescence of exudatinite occurs at a significantly different
wavelength to that of sporinite, regardless of the rank of the
coal. Other distinguishing features of exudatinite are the
broad maximum in its fluorescence spectrum, and a
tendency to display an initial increase in fluorescence
intensity, followed by a decrease as the time of exposure is
extended (Table 3.5).
(d) Other fluorescent materials
In addition to the presence of the three new macerals,
fluorescence microscopy has revealed certain phenomena
which Teichmiiller (1974b) has associated with the
generation of mobile bitumen. These include oil
exudations, smear films or the darkening of vitrinite as a
result of irradiation, and fluorescent vitrinite, presumably
due to the incorporation of lipoid substances. The
fluorescent vitrinite is usually that which occurs in attrital
bands rather than as bands of telocollinite.
(e) Quantitative fluorescence photometry

Red/green
quotient

510-570 nm

ca. 0.5

weak, even
negative

ca. 2.6

very strongly
positive -

orange to brown ca. 635 nm

mostly
orange
brown

Alteration of
fluorescence
intensity with
time

Maximum
fluoresence
intensity (X)

yellow to ca. 635 &


and red580 nm

(Q)

ca. 2.2

often an increase then


decrease

In quantitative spectral fluorescence photometry, the


recommended optical apparatus is somewhat different to
that described above for qualitative work (Ottenjahn et al
1974; I.C.C.P. 1976; van Gijzel 1979). A mercury lamp is
used with an ultraviolet filter orfilters to produce excitation
mainly by the mercury band at 365.5 nm, since a full
fluorescence spectrum cannot be obtained with the blue
light filter combination. Spectral analysis of the
fluorescence emitted by the object in the range 400-700 nm
is made by a motor-driven continuous interference filter or
grating monochrometer synchronized with a chart recorder.
The barrier filter is withdrawn in this case while the
measurements are taken. The photomultiplier used should
also have an adequate response through the relevant spectral
range.
It is the shape of the fluorescence spectra that is currently
used in coalification studies, not the absolute intensities
involved. The parameters measured include the peak
wavelength, the red/green quotient, and the alteration,
which is an increase or decrease (fading) in intensity after
specified periods of irradiation. Ottenjahn et al (1974) have
shown that the peak wavelength and the red/green quotient
obtained on sporinite increase with increasing rank up to
medium volatile bituminous coal, which is the highest rank
in which the fluorescence phenomenon is normally
encountered.
3.3.5

Macerals in low rank coals

Lignites and sub-bituminous coals have physical and


chemical properties which seem to set them apart from coals
of higher rank. Likewise, when they are examined under the
microscope, they appear more complex, showing greater
variability in the macrai materials. Many petrographers

therefore believe they can better characterize these low rank


coals with a

special macrai classification rather than constrain (b) Humodetrinite


themselves with the system described above that is
The humodetrinite sub-group contains the macrais attrinite
traditionally used for hard coals.
and densinite. Both of these consist of fine, microscopically
The macrais derived by humification of lignodiscernable fragments, mostly less than 10 (im (Fig. 3.5d).
cellulosic tissues show the greatest changes as rank
Densinite differs from attrinite in glification, the particles
progresses. It is therefore in the terminology for these
macrais that the greatest differences between the two that it appears to have undergone a greater extent of
classifications are found. This group of materials in low having been cemented together with some loss of detail of
rank coals is referred to in this chapter as the huminite the particulate structure.
macrai group, and is regarded as equivalent to and the
precursor of the vitrinite macrais found in higher rank (c) Humocollinite
coals. Table 3.6 summarizes the classification of huminite
macrais, and gives details of their supposed origin and The third sub-group, humocollinite, includes the two
equivalents in the hard-coal classification system. The ' macerals gelinite and corpohuminite. Gelinite consists of
amorphous humic gels (Fig. 3.5c) and corpohuminite
group contains six different macrais, which are
consists principally of the coalified products of from
disrincmished from each other on the basis of ffrain
0
------------ - -. . .
piiiuuapiitiiv^a, nmui cue piiiiicuy een ww.iv.uuuo uuivcu
size and degree of glification. They are organized into three
sub-groups, namely humotelinite, humodetrinite and tannins. Corpohuminite typically has an elliptical or a rodhumocollinite, which represent a series of coarse-grained, like form imposed by the surrounding cell walls. It is
fine-grained and colloidal- size particles of humic materials, resistant to weathering, and isolated individual bodies or
respectively.
groups of bodies may become concentrated as a result of
destruction of the cell tissues. Corpohuminite often has a
(a) Humotelinite
reflectance higher than that of other huminite macerals (Fig.
3.5e). It is especially abundant in corky tissues and in
The humotelinite sub-group contains the macrais textinite lignites derived from conifers.
and ulminite, both of which have maintained a recognizable
The I.C.C.P. International Handbook of Coal
plant cell structure. In the case of textinite, the cell outlines Petrography (1971) gives many other details of these
appear sharp because they are ungelified. Ulminite may macerals, and also documents many of their technological
exist in various stages of glification, but, however properties in processes including briquetting, low- and highindistinct, a cell structure is still discernable (Fig. 3.5c).
temperature carbonization and extraction, and their
Although textinite is a common component of soft behaviour in weathering processes.
brown coals, such as those mined in the Miocene deposits of
West Germany, it has been observed only rarely in the
higher rank lignites of North America. Presumably it has 3.3.6 The chemistry of coal macerals
been transformed into ulminite as a result of geochemical Figure 3.6 shows the differences in elemental chemistry of
glification. Textinite is preferentially formed from the cell some important macerals in the same coal seams. It can be
walls of resistant plant tissues, notably those of conifers.
seen from this that the overall chemical composition of a
----------------------

<_,

coal sample to some extent


TABLE 3.6 Classification of the huminite macerals of low rank coals. (Modified from I.C.C.P. 1971.)
Equivalent in hard coals
Macrai group

' Macrai subgroup

Macrai

Source

huminite

humotelinite

textinite

ungelified cell wall material

ulminite

humodetrinite

humocollinite

attrinite
densinite
gelinite
corpohuminite

telinite/telocollinite
telinite/telocollinite

gelified plant tissues with


recognizable cell structure
humic detritus

formless humic gels


secondary colloidal cell
excretions, and primary cell
infillings (tannins)

* desmocollinite

gelocollinite
corpocollinite

Fig. 3.5 Reflected-light photomicrographs of coal (under oil immersion), (a) Dark oval alginite with crenulated margin in centre of field. Sporinite and
liptodetrinite appear somewhat lighter. Carboniferous boghead-cannel coal from West Virginia (x480). (b) Alginite derived from colonies of Reinschia.
Boghead coal, Kentucky, blue-light illumination (xl300).
(c) Ulminite consisting of highly gelified cell walls and with cell infillings of granular gelinite and more highly reflecting corpohuminite. Palaeocene subbituminous coal from Wyoming (x480). (d) The groundmass of this layer, from the same coal as Fig. 3.5(c), consists mostly of humodetrinite particles.
The white fragments are inertodetrinite and the dark bodies are mainly sporinite (x480). (e) The cell fillings of vesiculated corpohuminite are higher in
reflectance than the ulminite which encloses them. Palaeocene lignite, Montana (x480). (f) Particle of weathered medium volatile bituminous coal
showing microfractures and discolouration (x480).

Fig. 3.6 Carbon and hydrogen contents of macerals. (From Murchison 1964.) The smaller hatched lines connect points of the same rank, o Resinites; E
exinites; V vitrinites; M micrinites.

reflects the mixture of macerals that it contains. The liptinite


macerals at a given rank are richer in hydrogen than the
corresponding vitrinite, and this in turn has higher values
than the corresponding inertinite components.
The liptinite macerals contain the most strongly aliphatic
organic components, whereas the inertinite macerals contain
the most aromatics. This has been shown by infrared
spectroscopy, X-ray diffraction, physical constitution
analysis and broadline *H nuclear magnetic resonance
(N.M.R.) spectrometry (Dormans et al 1957; Cartz & Hirsch
1960; van Krevelen 1961; Tschamler & de Ruiter 1966).
More recently, Retcofsky and VanderHart (1978) concluded
from 13C cross-polarization (CP) N.M.R. that the fusinite
from a high volatile A bituminous coal was more aromatic,
and the liptinite in that coal less aromatic, than the vitrinite.
They also estimated that the number of rings per mean
structural unit of vitrinite of high volatile A bituminous rank
was 3-4, whereas fusinite had the largest polynuclear
aromatic ring system with five rings. The aromaticities of
macerals increase with increasing rank (Dormans et al 1957;
Davis 1978; Retcofsky & VanderHart 1978).

3.4 THE NATURE AND APPEARANCE OF


MICROLITHOTYPES

While a knowledge of the macerals present in a coal is


essential in most applications of coal petrology, there are
also a number of areas where the manner , of distribution
of these macerals through the coal and the way in which
the different macrai groups are associated with each other
may be highly significant.
A coal in which the vitrinite occurs predominantly as
relatively thick bands, for example, is likely to have
different breakage characteristics, and probably different
carbonization properties, to one in which the same amount
of vitrinite is finely disseminated throughout the seam.
The vegetation and the original swamp environments that
gave rise to the two types of materials may also have been
quite different in each case.
Associations of macerals, as determined
microscopically, are called microlithotypes. Just as the
macerals themselves are often regarded as an equivalent to
the minerals in other rocks, the microlithotypes may be
considered as equivalents to the discrete beds, lenticles or
laminae, made up of different mineral combinations, that
3.3.7
The microscopic appearance of coal
are also fundamental components of many clastic and nonminerals
clastic sediments.
The three macrai groups, vitrinite, liptinite and
The minerals occurring in coal are discussed in Section 2.9,
inertinite,
can be associated as shown in Table 3.7 to form
and the scope of this chapter does not include further
a
total
of
seven possible combinations. Three of these
coverage of these or other inorganic constituents.
combinations
are made up of one single macrai group
Nevertheless, the identification and
(monomaceral
microlithotypes), three contain members of
characteristization of mineral species under the
two macrai groups (bimaceral microlithotypes) and the
microscope is an important aspect of coal petrology.
last contains a representative of all three groups
Kemezys and Taylor (1964), and Mackowsky (1968a)
(trimaceral microlithotypes or trimacerites).
describe, in some detail, the appearance of the principal
According to established convention (e.g. I.C.C.R 1971),
coal minerals in microscopic studies.
the association must have a minimum band . width of 50 on
before it can be classed as a microlithotype. In addition,

constituents that make up less than 5% of the association are 5%, while a trimacerite must contain at least 5% of each of
normally disregarded. Thus, a band of vitrinite with a small the three macrai groups.
amount of (say) liptinite would not be classed as a clarite
Low concentrations of mineral matter are usually
unless the liptinite was present in greater abundance than ignored in the determination of microlithotypes. If

TABLE 3.7 Microlithotypes1 and carbominerites.


Microlithotypes
vitrite
clarite
trimacerite

f duroclarite
vitrinertoliptite
^ clarodurite
all on V and I
all on I and E
all on I
all on E

all on vitrinite (V) all on


V and exinite (E)
on Vs inertinite (I) and E, but with V > I and E on V,
I and E, but with E > V and I on V, I and E, but with
I > V and E
vitrinertite
durite
inertite
liptite

1attritus. In the same manner, fusain


bands less than 37 fim thick are
arbitrarily assigned as a constituent of
opaque attritus.Another problem, and
one of greater practical importance, is
the fact that the system does not lend
itself nearly as well to studies of all
ranks of coal as does the Stopes-Heerlen
system, simply because of the difficulty,
and in some cases the impossibility, of
preparing thin sections of high rank
coals. Also, the C l i t cin/^T7 nrdicVlPfl surfaces has become
JLUUJ V/*

a quantifiable technique through


reflectance measurement, while thin
section examinations retain the problem
of variation in optical properties with
section thickness. The Thiessen-Bureau
of Mines system is now obsolete in
practice, but as a large amount of
descriptive work on U.S. coal was done
by Thiessen and his colleagues, it is still
widely used for review purposes.
3.5.2
The genetic classification of
the U.S.S.R.
Academy of Sciences
The lithotypes and microlithotypes of
the Stopes- Heerlen system are
recognized on the basis of their
physical appearance and macrai
composition, respectively, and no
systematic palaeo-environmental

20-60% (by volume) clay mineral; remainder maceral 20~60%


(by volume) carbonate mineral, remainder maceral 5~20% (by
volume) pyrite, remainder maceral 20-60% (by volume)
quartz, remainder maceral 5-60% (by volume) of various
minerals > 60% clay, quartz, carbonate, > 20% pyrite

Associations of microlithotypes
impurities
(carbominerites)
carbankerite
carbopyrite
carbopolyminerite dirt, pyrite

the amount of
mineral matter is
significant, but the
relative density of
the microlithotype
is less than 1.5, the
abundance
and
type of mineral
matter can be
described by a
qualifying
adjective,
using
terms such as
'argillaceous
durite5 (Stach et al
1975). However,
where the mineral
matter is more
abundant, and the
relative density of
the association lies
between 1.5 and
2.0, the material is
referred to as a
carbominerite.
The
types
of
carbominerites
normally
recognized, and the
volumetric
percentages
of
mineral
species
that correspond to
the
required
density range, are
also given in Table
3.7. The names of
both
microlithotypes
and
carbominerites
both have the
suffix ending ite,
as for example, in
vitrite.
The
methods
of
microlithotype
analysis
are
discussed
more
fully

with mineral
carbargilite
carbosilicite

3.5 OTHER
CLASSIFI
CATION
SYSTEMS
FOR
COAL
MICROCO
MPONENT
S
The
StopesHeerlen system for
identification and
nomenclature
of
coal constituents,
as described in
Section 3.3, is the
principal system of
classification used
throughout
the
world
at
the
present
time.
Because it is based
on three maceral
groups, vitrinite,
liptinite
and
inertinite,
analytical results
can be plotted
readily in simple
representations
such as triangular
diagrams,
yet
where
greater
detail is required,
data can be readily
extended
to
encompass
the
individual
macerals or submacerals of each
group. In this
section, however,
some
other
systems used to
classify
the
microcomponents
of
coal
are
considered.

3.5.1
T
he
ThiessenBureau of
Mines
system of
coal
classificat
ion
Following
very
comprehensive
studies of coals in
thin section at the
U.S. Bureau of
Mines, Reinhardt
Thiessen (Thiessen
1920; Parks &
ODonnell 1956;
I.C.C.P.
1963)
developed
a
system
of
description for the
microscopically
recognizable
ingredients of coal.
The three major
components of
banded bituminous
coal
in
this
classification can
be identified at
either
the
macroscopic
or
microscopic level.
These
are
anthraxylon,
equivalent to the
bright
vitrain
bands of coal,
fusain, which is
much the same as
defined in the
Stopes-Heerlen
system,
and
attritus, which is
represented
by
those bands of coal
with
a
dull,
prarmlar
armearance
and
consisting of a
micro-

c*---------- ~r --------------------------- --------------------o - --------- -

fragmental mixture
of varied entities.
Microscopic
examination
of
thin
sections
enables
the
constituents
of
attritus
to
be
distinguished
as
either translucent
attritus or opaque
attritus.
Translucent attritus
includes
spores,
cuticles, resins etc.,
and opaque attritus
includes granular
opaque
matter
(micrinite),
sclerotia etc.
Table
3.8
summarizes
the
Thiessen-Bureau
of
Mines
nomenclature and
classification, and
correlates
the
terms used with
those
of
the
Stopes-Heerlen
system. A feature
of the ThiessenBureau of Mines
system is that
arbitrary thickness
limits were set for
some
of
the
components
and
constituents.
Anthraxylon, for
example, includes
only those vitrain
bands greater than
14 xin thick, and
any vitrinite with a
lesser
band
thickness would be
described
as
translucent humic
degradation
matter,
a
constituent
of
translucent

Co
al
Pet
rol
ogy
and
Pet
rog
rap
hic
An
aly
sis
TABLE 3.8 Correlation of the Thiessen-Bureau of
Mines and Stopes-Heerlen classifications. (Modified
Transmitted light
Thiessen-Bureau of Mines System
Banded
components
Constituents of attritus
Anthraxylon
(translucent)

Translucent
attritus

Attritus

Opaque
attritus

Macrais

Macrai
group

vitrinite more than 14 fim in width

Vitrinite

translucent humic matter

vitrinite less than 14 un in width

spores, pollen, cuticles, algae


cuticles, algae
resinous and waxy substances

sporinite, cutinite, alginite


resinite

Liptinite

brown matter (semitranslucent)

weakly reflecting semifusinite


weakly reflecting macrinite

Inertinite

rpfl^rtirta cWotiniie

granular opaque matter

WwaiMj ivi*wu**p ------------------------------------

micrinite
amorphous (massive) opaque matter,
finely divided fusain, sclerotia

Fusain
(opaque)
after I.C.C.P. 1963.)
fusinite less than 37 m
in width strongly
reflecting macrinite
strongly reflecting
sclerotinite
fusinite and semifusinite
more than 37 fim in width

interpretation
is
implied. Indeed, as
can be seen from
Section 3.8, there
is
often
no
adequate
consensus
of
opinion
among

Reflected light StopesHeerlen System

coal
petrologists
regarding the environmental
conditions
that
gave rise to many
of the major coal
lithotypes.
However, at the

Institute
of
Geology, Academy
of Sciences of the
U.S.S.R., Moscow
a
genetic
classification
of
microcomponents
of humic coals was
developed
following detailed
study
of
the
majority of coal
deposits and basins
in the U.S.S.R.,
representing a wide
range of tectonic
and environmental
settings (Timoveev
&
Bogoliubova
1965;
I.C.C.P.
1971).
Within
this
system, coals are
classified
according to the
material
composition of the
coal (class and
subclass), and the
degree of structural
preservation
or
degradation
(group).
The
horizontal rows in
Table 3.9 represent

the six classes of


materials. In thin
section, the classes
gelinitic,
semigelinitic,
semi- gelifusinitic,
gelifusinitic,
quasigelifusinitic,
and
fusinitic
contain materials
which
vary
progressively from
red, through brown
to black. This
progression
reflects
increasingly
aerobic conditions
in the peat bog,
due in turn to the
degree of flooding
and
water
movement.
The
processes by which
the original lignocellulosic
plant
tissues
were
transformed into
the
microcomponents
characteristic
of
these coal classes
are seen from the
table
to
be
glification for
gelinitic

and semigelinitic coals, fusinization for fusinitic coals,


and a two-stage process of glification followed by
fusinization for the other three classes. A characteristic
feature of quasigelifusinitic coals is that they contain
large amounts of detrital quartz and clays, as a result of
deposition under flooded, running- water conditions.
The vertical columns in Table 3.9 are the genetic
groups of coal, namely telinitic, posttelinitic,
precollinitic, collinitic and leiptinitic. These five groups
represent progressively greater physical and biochemical
degradation that developed in response to increasing
tectonic stability of the area of peat accumulation. In a
tectonically stable area of peat accumulation, for
example, extensive decomposition of vegetal material
would have led to the formation of the collinitic group of
coals. Within a tectonically mobile area, by contrast, a
more rapid rate of subsidence would have provided
greater opportunity for the preservation of plant
materials, giving rise to the peats from which the telinitic
group of coals would form.
Therefore, in order to classify a coal using this system
it is necessary to characterize both the type of substance
(class), and the structure (group) of the microcomponents
present. Table 3.9 shows how each group is sub-divided
into sub-groups based on the class of microcomponents,
each of which represents a definable environmental
setting. Thus, gelinite- posttelinitic coal, for example, is
genetic sub-group representing the product of a heavily
flooded, stagnant peat bog within a relatively tectonically
mobile area.
Although some of the terms are similar to those in the
Stopes-Heerlen system, they are not used in the same
way. For example, the term fusinite-telinitic coal does
not imply the presence of the vitrinite group macrai
telinite. Rather it is a coal containing fusinite with a
distinct cellular structure. Not only is the U.S.S.R.
Academy of Sciences nomenclature used for strictly
genetic purposes, but also for the industrial evaluation of
coals in that country.
The above description provides only a summary of
some of the principal features of the Russian genetic
classification, since each of the sub-groups may be
further differentiated into several genetic types. The
system has not been used extensively outside the
U.S.S.R., possibly because of its complexity, while
western petrographers are also deterred by the
descriptions being based on thin-section examination.
Another significant drawback at the present time is that
correlations between the genetic coal types and original
coal swamp facies have not been confirmed for coal
basins outside the U.S.S.R. Nevertheless, the approach
and findings of the Russian petrologists should be given
careful con-

>
q

a a

8 'c
T3

.5 V
5 11
<2 '3

O .3
<j e

I . is
II
&
U) 3

aI

4j %)

a
y
1*

I i
II

Si
kS

;a g.
I|iJ

1i

u&

<5 -| 3
It

'K 55.

-s <2
G U3 W

tji A

J
a

3
a

a
|||
O MH

t- s i
3 2 # 8
g?
a -i

H
a2 H
G

the conditions of illumination. This property is related


to the aromaticity of the organic compounds in the
coal concerned, and increases progressively for all
macerals as the rank of coal increases.
3.5.3
Spackman terminology and
Precision measurement of the reflectance of
classification
individual macerals, particularly vitrinite, is widely
Spackman (1958) noted that each conventional used as an index of coal rank. The technique has
maceral term really represents a suite of materials the advantage over chemical parameters, such as
with greatly varying physical and chemical those outlined in Section 2.10, that it is applied to a
properties. He therefore considers that is is single selected petrographic constituent, and is
important to designate macerals by the extent of therefore not
the metamorphic changes which have occurred.
The ending ^inoid5 in, for example, vitrinoid,
according to this classification implies a maceral of
the Vitrinite Suite,5 but one with distinctive
properties rather than a range such as exists for
sidration in any interpretations
environments of coal formation.

of

___!____ ____n*->A fiicinAI/^O

flip

De various nntunuiua cum luoiiivi^o nuiuu .x.w


Tnertinite Suite5, and exinoids and resinoids within their
own respective suites.
Subsequent usage of the Spackman system seems to
have diverged somewhat from this original concept.
Schapiro and Gray (1960) established 22 vitrinoids, each
with a designation derived from the vitrinite reflectance
distribution. Thus, vitrinoid V22 is an
vitrinite. Likewise, he proposed that there
anthracitic vitrinoid with a maximum reflectance
in the range 2.20-2.29%. Such categories are
sometimes referred to as vitrinoid types or Vtypes5 (Section 3.7.6). Although it may be useful in
aspects of coal utilization (Section 4.3.6), this is a
somewhat unfortunate development because it
means that vitrinites are not being sub-divided on
the basis of the kinds of material involved, but on
quite arbitrary divisions of the normal distribution
of vitrinite reflectance readings. These divisions
may group different materials together, or
unnecessarily subdivide portions of the same basic
material into different
V-types.
3.6 REFLECTANCE OF VITRINITE AND
OTHER MACERALS
The reflectance5 of a maceral or other particle in a
coal is the proportion of directly incident light,
usually expressed as a percentage, that is reflected
from a plane polished surface under specified

influenced by the relative proportions of the


different macerals within the coal. Considered in
conjunction with an index of coal type, such as
the actual percentage of vitrinite present, the rank
of the coal determined in this way provides an
ideal basis for petrographic classification (e.g.
Bennett & Taylor 1970).
3.6.1
Theoretical basis of
reflectance measurement
Reflectance measurements made with the
precision required for coal rank determination are
generally carried out by comparing the amount of
light reflected from the maceral concerned to the
amount of light reflected from a standard
substance under the same illumination
conditions. Such measurements can be carried
out with dry objectives or with water- or oilimmersion lenses, but most studies employ an
oil-immersion technique, using an oil with a
refractive index of 1.518 at 23C and an incident
light wavelength of 546 nm. The amount of light
reflected from the surface of the maceral is
determined from the electrical output of a
photomultiplier system (Section 3.7.6), and the
reflectance can be calculated from the following
formula:

where R = reflectance of the coal maceral; Rs =


reflectance of a calibration standard; A =
deflection of galvanometer, chart recorder etc. of
photomultiplier system for the maceral; and As =
deflection of galvanometer, chart recorder etc.
for the calibration standard. In most cases,
however, the petrographer sets the reflectance of
the standard as the galvonometer reading, so that
the reflectance is read directly.
The relationship between reflectance (R) and
the other optical properties of both the reflecting
material and the immersion medium can be
expressed by Beers equation as follows:
(n - nxf + w2k2 ^
(n + nx)2 + n2 ke
where n = refractive index of reflecting material;
k = absorption index of reflecting material; and
nx = refractive index of medium in which the
measurement is made.
The refractive index of immersion oil
changes with temperature, and this in turn

influences the value obtained in reflectance


measurement. In order to

obtain a value of the relectance in oil (RJ corrected


to a standard temperature (23 or 25 C), it is
necessary, in theory, to measure the reflectance in
two media (air is convenient as the second) and
calculate n and k from the following equations:

Vitrinite
in coal
often
displays
a three-

n=
2 -

1 /..2

i a+ *0)/(i - Ro) -(! + R ) / ( i - R )


and
k2 = jg, ( n + l)2 - ( n - 1 f
2 (1 - R )
where i? = measured reflectance in oil; R =
o

ya

measured reflectance in air; and n 1 = refractive


index of oil at temperature of measurement.
Substituting the calculated values for n and k
and the known value of n x at 23 C (usually
1.518) into Beers equation above allows the
calculation of R o (23 C). Note that the
reflectances in these equations are expressed to
unit base, rather than as percentages.
In practice, such computations are seldom performed. Instead of correcting for the actual oil
temperatures, the R o of glass calibration standards
is calculated from their refractive indices using the
Fresnel equation, with a value of 1.518 assumed
for r
R=
S

(s - X 100
(2 + 1 f

where R s is the reflectance of the standard


expressed as a percentage, n 2 is the refractive
index of the glass, and n 1 is that of the immersion
medium.
Obviously, if the reflectances of the maceral
being measured and the standard used for
calibration are very close, any variation in
refractive index of the oil due to a change in
temperature from 23 C will cause both to vary by
similar amounts. However, if the coal and the
standard have widely differing reflectances, any
variation in refractive index of the oil will
introduce significant errors. Consequently, it is
customary to calibrate reflectometer systems with
standards whose reflectances have been calculated
using n0 = 1.518, and which are close to those of
the macerals under study.
3.6.2

Optical anisotropy of reflectance

Fig. 3.7 The anisotropic character of coal. 7?and R are


the maximum and minimum reflectances of vitrinite,
respectively. R is an apparent minimum reflectance,
intermediate between R and R

dimensional variation in reflectance that is similar


to the variation in optical properties exhibited by
uniaxial negativesubstances. The optic axis of the
indicatrix in these circumstances is approximately
normal to the bedding plane. The vertical axis is
shorter than the two horizontal axes, mainly in
response to the vertical stresses imposed by the
weight of superincumbent strata (Fig. 3.7).
When viewed under plane polarized light, all
planes or sections through a given layer of
vitrinite, according to this model, should display
the maximum reflectance of the material in at least
one orientation on the microscope stage. A section
cut perpendicular to bedding should display the
maximum reflectance when the plane of
polarization of the light is along the bedding trace
and the minimum reflectance of the material when
it is at right angles to this direction. A section cut
parallel to bedding, on the other hand, theoretically
should exhibit the maximum reflectance in any
orientation, while one that is oblique to bedding
should display the maximum reflectance when the
light is polarized along the bedding trace and a
value intermediate between the minimum and the
maximum (i.e. an apparent minimum) when it is at
right angles to the bedding.
In a grain mount made up of crushed fragments
of the coal embedded with random orientation in a
plastic binder, every vitrinite particle examined
should exhibit the maximum reflectance in at least

one position during rotation of the microscope


stage when the field is illuminated with vertically
incident
plane polarized light. The value of this maximum
reflectance can be recorded for a large number of
particles and the mean maximum reflectance5
(Rmax) calculated to give a widely used rank
parameter.
If the reflectance is measured in non-polarized
light, on the other hand, the reflections from all
directions on the vitrinite surface will be integrated
to give a random reflectance5. Figure 3.8
compares the influence of the reflectance
anisotropy of a surface upon reflectances obtained
in polarized and non-polarized light. A statistical
mean of random (non-polarized) readings taken on
many vitrinite particles in a polished grain mount
of a coal theoretically gives the same value for the
random reflectance as would be obtained if the
random readings were taken in polarized light
(without stage rotation). However, the range of
theoretically possible values from a sample is
much less for non-polarized iigbt [ Rm t0 (Rmax + R
min)/2 1 than for polarized light (Rmax to Rmin)- The
statistical relationship between random, maximum
and minimum reflectances (Hevia & Virgos 1977)
is also shown in Fig. 3.8. Davis (1978c) has
(iy/z/, none anu ^OOK

Not all coals display uniaxial reflectance


characteristics. Hevia and Virgos (1977), Cook et
al (I98la) and Levine and Davis (1984) have all
reported the existence of coals that demonstrate a
biaxial anisotropy. An important implication of this
observation is that not all coal sections need
necessarily display the maximum reflectance
value. Most of the above reports describe biaxial
negativecoals, but Levine and Davis (1984) have
also reported a sample of low volatile bituminous
coal from Pennsylvania that had the characteristics
of a biaxial positive material.
3.6.3

Vitrinite reflectance and coal rank

The reflectance of vitrinite, as well as that of other


macerals, particularly those of the liptinite group
(see below), increases progressively with the rank
of the coal in which it occurs. Figure 3.9 illustrates
the relation between the maximum reflectance of
vitrinite and the carbon content of the same
material (another rank indicator), arid Table 3.10
gives an indication ofthe ranges of mean maximum
vitrinite reflectance (max) that correspond to the
principal A.S.T.M. rank designations, based on
other parameters as outlined in Section 2.10.2. The
rate of change of reflectance is not uniform with

nuwer anu i^avis

summarized the relative merits of maximum and


/1

. .. .

..J

random reflectance
petrographic studies.

1_

/1 A rjA\ TT_____________________ _________I T"\_________* _

determinations

in

coal respect to many of these other indices, and, in fact,


reflectance is a sensitive indicator of rank change
in higher, rather than lower rank deposits.
Figure 3.9 also shows the variation in minimum
reflectance, and hence, from the difference
between maximum and minimum reflectance, the
bi-reflectance5 of the vitrinite. This latter property
also increases significantly with rank, although, as
discussed further
in
uiv.iv, io uvn nv,vv,ooai relation between
rank and anisotropy in some individual coalfields.

n
iimi
' ........

a uiiuui in

3.6.4

4.01-

Reflectance of other macerals

0 > ! _________I __ _I
___________1 __________l 1 _
70

Carbon (%)

75

.80

85

90

95

Fig. 3.9 The relationship between carbon content and


maximum () and minimum (o) reflectances of
vitrinite. (From Davis 1978.)

TABLE 3.10 Vitrinite reflectance limits (in oil) and A.S.T.M.


coal rank classes.
Rank

t i l l I

The coalification track of sporinite displays a


sharp increase in reflectance, known as the
coalification jump (Stach & Michels 1955), at a
rank corresponding to a vitrinite reflectance of
about 1.2%. The reflectances of these two macerals
then merge at a vitrinite reflectance of about 1.5%.

Maximum reflectance (%)

sub-bituminous
high volatile bituminous C
high volatile bituminous B
high volatile bituminous A
medium volatile bituminous

<0.47
0.47-0.57
0.57-0.71
0.71-1.10
1.10-1.50

low volatile bituminous


semianthracite
anthracite

1.50-2.05
2.05-3.00 (approx.)
>3.00 (approx.)

semifusinite, micrinite and sporinite relative to that


of vitrinite in a number of coals. The reflectance of
fusinite approaches that of vitrinite as rank
increases, and in the case of a peranthracite
(Alpern & Lemos de Sousa 1970) the fusinite may
actually have a lower reflectance than the vitrinite
in the same coal.

The reflectance of resi- nite is similar to that of


sporinite in the same coal, while that of cutinite is
somewhat higher. Cutinite also exhibits a strong
reflectance anisotropy. Hower (1978) has
described liptinite macerals (sporinite, resi- nite
Figure 3.10 shows the random reflectances
of fusinite,
(b)

(a)

Fig. 3.8 Random reflectance in polarized and non-polarized light. In polarized light, the random reflectance obtained on a single particle would
vary between R and R' - , the apparent minimum reflectance, depending upon the orientation of the particle. In non-polarized light,
the reflectance in all directions will be integrated into the random reflectance reading, regardless of orientation. The relationships
shown are from Hevia and Virgos (1977). (a) Polarized light, (b) Non-polarized light. _
_
R^rcpraaite the average of a number of readings. Rnild ^ = Rrand = (2Rmax + Rmin)/3;
^rand non-pol ^ av ^max

^ min

Fig. 3.10 Reflectances of different macerals through a range of coal rank. (From Hoover & Davis 1980.) Vitrinite; sporinite; A
fusinite; A semifusinite; o micrinite.

and cutinite) in an anthracite sample, all with


similar random reflectances of about 4.9%. The
bireflectances of the liptinite macerals, especially 3.6.5 Reflectance at different wavelengths
the cutinite, in this case were greater than that of Figure 3.11 shows that, for bituminous coals, the
the vitrinite.
reflectances in air and oil, and the refractive index
of vitrinite, decrease with increasing wavelength of
the incident light. With increasing rank, the
dispersion curves for air reflectance and refractive
index tend to flatten, and, in the case of
anthracites, there is a reversal in slope, so that
values increase towards the red end of the
spectrum.
From the visible region of the spectrum into the
ultraviolet, the reflectances and refractive indices
of vitrinites have been observed to decrease. The
absorption coefficients, however, tend to peak
within the ultraviolet region near 280 nm,
extending into the visible range in the case of
anthracites (Gilbert 1962; McCartney et al 1965).
This strong absorption in the ultraviolet has been
interpreted as being due to the presence of
significant proportions of aromatic structures. The
different slopes of the dispersion curves for
vitrinites of varying rank, relative to the slope for a
calibration standard, provide a compelling reason
why reflectance measurements should be made in
monochromatic light.

1.701----------------1------------------1-----------------1----------------*
600
550
500
450
400
Wavelength (nm)

Fig. 3.11 Dispersion of the optical properties of vitrinites (From Davis 1978.) (a) Air reflectance, (b) Oil reflectance.
(c) Refractive index, o 82.6-83.4% carbon; A84.0% carbon; A87.7~88.1% carbon; 91.4% carbon; 92.8% carbon;
92.8%~93.3% carbon.

Murchison and Jones (1964) and Hevia (1974)


have shown that the shapes of the dispersion
curves for liptinite macerals are fairly similar to
those of vitrinite from the same coal, although, of
course, displaced to lower values. The curves for
inertinite macerals are less steep than those of the
vitrinite; those of higher reflecting inertinite have
very low slopes, and some display minima in the
range 480-550 nm (Hevia 1974).
i 7 DPTDnnu APuir ANAT VK
M.

AJ 1

111V lllUllilk/lU

Micropetrographic study of coal may be used


either as a technique of quality evaluation in the
testing of a coals economic value, or as an aid to
understanding the geologic history of the material.
Though the selection of samples to be studied in
each case may be somewhat different, both types
of study generally involve measuring the relative
proportions of the various macerals and/or
microlithotypes that may be present, and assessing
the rank of the coal by means of such properties as
vitrinite reflectance.
The actual techniques involved in such studies
are similar in many ways to petrographic methods
used in other branches of geology, although a
number of refinements have been introduced to
deal specifically with coal and related materials.
As with other aspects of coal analysis, several
international and national standards are available
that describe the most appropriate procedures in

some degree of detail. In other cases, such as with has set. Once set, the moulded resin and coal pellet
the introduction of automated microscopy to coal is ground and
petrography, much of the necessary information
can only be found in the discussions of various
research investigations.
3.7.1

Sample preparation

As indicated in Section 3.3, almost all petrographic


studies of coals are now carried out by means of
polished section techniques. Polished sections of
coal may be prepared from single lumps, broken pr
sawn from a hand specimen, a section of drill core
or an exposed coal face, or they may be prepared
from a representative sample of the seam or seam
sub-section, crushed to a granular or coarse
powder form. Individual lump specimens have the
advantage that they can preserve the geometric
relationships of the various bands and other masses
in the coal to a greater extent than crushed coal
samples, but have the disadvantage that they only
represent a limited sub-section of the seam in
question. A relatively small amount of a wellprepared, representative crushed sample, on the
other hand, can provide data on the abundance and
optical characteristics of the macerals or
microlithotypes present in a much larger mass of
in situ or mine product material.
Lump specimens for polished blocks are
usually air-dried and impregnated with an
appropriate resin if necessary to provide strength.
The specimen is then trimmed to expose the. face
to be studied (usually one perpendicular to the
bedding planes), and mounted in an appropriate
mould with a cold-setting resin material. Once set,
the embedded specimen can be ground or cut along
the desired plane, and polished with successively
finer abrasive powders.
Crushed coal specimens for petrographic study
are usually prepared to a coarser particle size than
are samples for many other analyses. For most
purposes, a maximum particle size of 1 mm is
required, compared to 200 fim for (say) proximate
analysis, but it is also necessary to produce a
minimum amount of fine material in the crushing
process. These requirements should be taken into
account when designing the sample preparation
sequence for a coal analysis programme in which
petrographic studies are to be included at some
stage.
The crushed material is well mixed with a coldsetting resin to form a paste or slurry, and the
mixture poured into a mould of appropriate shape
to harden. Care should taken to avoid the solid
particles settling and segregating before the binder

polished for petrographic examination. If any


segregation of panicles has occurred, the pellet
should be cut in the vertical plane and one half
polished. An alternative procedure is to employ a
hydraulic press to prepare cylindrical moulds in
which close packing prevents segregation.
Grinding and polishing are carried out with a senes
of abrasives on wet laps of low-nap cloth and/or
silk. A high quality of relieffree polish is required
for reflectance determinations.
The techniques of polished section preparation
are described in some detail by the I.C.C.P. (1963),
Stach et al (1975), A.S.T.M. (1981a) and the
International Organization for Standardization
(I.S.O.) (in preparation, a). Although much less
widely used, the methods of preparing coal thin
sections for microscopic examination are given by
Thiessen (1920), van Krevelen (1961) and Francis
(1961).
3.7.2

maceral
and
also
activate
the
stage

Maceral analysis

The relative proportions of the various macerals in


a coal, at least on a volumetric basis, can be
determined from either polished blocks of lump
coal or grain mounts by the techniques of modal
analysis. In the past, this has included lineal
Mg. 3.12 Microscope equipped with Swift automatic point counter.
analysis by means of the integrating stage (van
Krevelen 1961; Galehouse 1971)
advance mechanism each time a counter is
vYiuvix iiiuv^u Liiv linage L/v.ixc-aLii LIIV
uy
a. depressed, moving it on to the next grid point.
otiltd
According to Hilliard and Cahn (1961), a twoof micrometer-graduated screw spindles. Each
dimensional
grid is the most efficient method of
spindle was moved only when a particular maceral
m r \ T - t - T - i c it*v> nrro KaManf-U tUa atTamaan K.r n nam/vi

was under the cross-hair in the field of view. At the


conclusion of the examination, the amount of each
maceral present was determined from the
contribution that each spindle had made to the total
traverse length.
Today, however, modal analysis is more
commonly based on the technique of point
counting (Glagolev 1934; Chayes 1949, 1956), in
which the microscope stage is moved in a series of
fixed increments and the identity of the maceral
falling beneath the cross-hairs after each advance
is recorded. The volumetric abundance of each
maceral is determined from the total number of
points at which it was encountered in relation to
the total number of points recorded in the traverse
network.
Stage movement can be accomplished
manually with a mechanical stage equipped with
check stops, or even by visually adjusting the stage
vernier scale. However, there are automatic point
counters such as that displayed in Fig. 3.12, which
keep a tally of the number of counts made on each

E95 = 2

1/2

analysing volume proportions, with a grid


spacing such that the majority of structural
features should be intersected by no more than
one grid point. The I.S.O. (in preparation, b)
recommends that maceral analyses should be
performed on minus 1 mm coal particles, and,
with the I.C.C.P. (1963), that interpoint and
inter-line distances should be approximately 0.5
mm. The probable error, at the 95% confidence
level, involved in counting individual
components (E95) is given by the equation

(loo - p )

where p is the percentage of the individual


component, and n is the total number of points
counted.
For components present in proportions of 10,
50 and 90%, the probable errors involved in a
maceral analysis of 500 points are 2.7, 4.5 and
2.7%, respectively, assuming of course that no
errors in identification are made during the
analysis. The repeatabilities of such results
(2F/V2) are 3.8, 6.4 and 3.8, respectively. The
reproducibility, that is the difference between
results of two different operators

analysing different sub-samples of the same coal,


has been found to be about 1.5-2 times the
theoretical repeatability (I.S.O., in preparation, b).
Thus, maceral percentages should be reported to
the nearest integer and not, as is frequently done,
to the first decimal place.
The total number of points counted in maceral
analysis can vary according to the standard
procedure followed, the purpose of the analysis,
and the accuracy desired. Galehouse (1971)
provides useful nomograms and tables for
determining quickly the number of points that must
be counted in order to achieve desired levels of
accuracy. The I.C.C.P. (1963) recommends
counting 500 or 1000 points to give an accuracy of
about 2-3%. I.S.O. (in preparation, b) requires
that at least 500 points be counted. The A.S.T.M.
(1981b) standard deals with precision in maceral
analysis by requiring that two separate analyses of
1000 points be performed on each of two grain
mounts. For the values to be acceptable, the mean
difference between maceral percentages of the two
grain mounts should not exceed 2%.
The magnification employed for maceral
analysis should be such as to permit resolution of
most maceral occurrences, i.e. at least x200
overall. However, A.S.T.M. (1981b) recommends
that it should be greater than X400. Oil immersion
objectives are normally used to provide best
identification conditions.
Selection of the macerals or maceral groups
and, if desired, the mineral species to be counted
depends upon the purpose for which the analysis is
intended and the predilections of the petrographer.
Some laboratories count only the three maceral
groups, whereas the A.S.T.M. standard specifies
the six macerals that should be counted.
Various conventions can be adopted to help the
netrwranher maintain consistency in those
instances
I ------O - -1 ----- --------------- ------------ - J

- - -

- -

Fluorescence microscopy can enable distinctions to


be made between liptinite macerals which would
otherwise appear uniformly dark in reflected white
light. Also, in petrographic analyses performed in
blue light, the percentages of the fluorescent
liptinite macerals often appear greater than when
conventional white light is used. Identification is
made more readily by observing the fluorescence
colours on a dark background than by observing
the dark reflections on a medium background.
Table 3.11 provides a comparison of selected
analyses carried out both in white and combined
white and blue light, showing the larger
proportions of liptinite macerals which were
recorded in the combination analysis (Davis 1975).
In the combination analysis, the percentages of
liptinite macerals were determined in blue light,
and the balance, due to nonfluorescing vitrinite and
inertinite group macerals, apportioned according to
the results of the white-light analysis.
3.7.4

Bases for reporting maceral analyses

Petrographic analyses are reported on either a


mineralfree or mineral-containing basis. For the
latter, the amount of mineral matter may be
determined together with macerals during point
counting, or calculated using an empirical formula.
Direct determination using techniques such as
radio-frequency oxidation is also possible (Section
2.9.2).
Some difficulties are associated with the
determination of mineral percentages by modal
analysis. Certain minerals, including quartz and the
volumetrically important clay minerals, often
appear dark, almost black, under oil immersion. It
is therefore possible to mistake them for voids or
for the resin binder, esneciallv where there is no
surrounding matrix of
--r----------J

- - - - - - -

- -

- -

where the cross-hair lies over the perimeter of -------------(_j higher reflecting coal macerals to provide an
particles or on boundaries between macerals (Stach
optical contrast. Other minerals, including pyrite
et al 1975; I.S.O., in preparation, b).
grains, may be plucked out during polishing, so
that their proportions may tend to be
3.7.3
Combined white- and blueunderestimated. In some instances, the minerals
light petrographic analysis
are so finely divided that it is impossible to resolve
them under the microscope, while much of the
inorganic content of lower rank coals also exists as
ion-exchangeable cations within the organic
molecular structure rather than as discrete mineral
species.
In spite of these difficulties, it is often useful to
include total minerals as a category in the pointcount analysis, or even to count individual classes
of minerals, such as clays, carbonates, pyrite,

quartz etc. Several formulae have been developed


to calculate the volume percentage, rather than
mass percentage, of mineral matter in coals. The
A.S.T.M. (1981b) equation below is based on the
Parr formula (Section 2.9.2), and average densities
of 2.80 and 1.35 have been assumed for the
mineral and organic contents, respectively, to
convert the mass of mineral matter to a volume
percentage.
_ ________________100
[(1.08A
+
Q.55S)/2.8]
__________________
(VOl)
[100 - (1.08A + 0.55S)]/1.35 + (1.08A - 0.55S)/2.8

when A and S are the ash and sulphur values of the


coal, respectively, expressed on a dry basis.

TABLE 3.11 Comparison of white-light and combined white- and blue-light maceral analyses (PSOC-123; lithotype of No. 5. Block Seam, West
Virginia; Ro = 0.71%).
Analysis
White light (%)
Combination
(white/blue) (%)

Vitrinite

Sporinite

37

24

29

33

Cutinite

Resinite

Alginite

Fusinite

Semifusinite

Macrinite

15

11

The I.S.O. (in preparation, b) has simplified


this to:
Although the technique is not as widely used as
maceral analysis, the relative proportions of the
various microlithotypes and, if necessary, of

MM, =
MMD
(vol) _________________________p_____________
carbominerites (coal-mineral associations) in a
2.7 - (0.011 MMp)
coal sample can also be determined by modal
where MMp, the weight percentage mineral matter, analysis methods. In the most common of the two
is derived through the Parr equation and equals analysis techniques, the microscope eyepiece is
fitted with a 20 point cross-line reticule (Fig. 3.13).
1.08A + 0.55S.
The magnification characteristics of the
An empirical equation has also been given by
microscope are matched to the actual size of the
I.S.O. (in preparation, b) as follows:
grid to delineate a 50pm squarein the field of view,
normally with a x25 objective being used. Each
MM, n = 0.61A - 0.21
band or maceral association falling within this
(vol)
square can be evaluated in terms of size (i.e. if it is
Benedict et al (1968) have used the equation greater than 50 pm in thickness) and also with
below as a simple method for estimating the respect to the relative proportion of the various
volumetric mineral content of coals:
constituents. For example, if all 20 points on the
reticule fall on vitrinite, the microlithotype has less
MM(voI) = 0.6 (A + S)
than 5% of any other maceral group and is
therefore classed as a vitrite. If one or more of the
where A and S are as defined above.
20 points coincide with inertinite, however, the
Occasionally, the results of maceral analysis material would be classed as a vitrinertite (Table
are reported on a weight percentage basis. To do
3.7).
this one must assume a value for the relative
Each observation on a group of 20 grid line
density of the macerals at the particular rank level
intersections is regarded as a point, even though
of the subject coal. The I.C.C.P. (1963) lists the area covered may in fact cross the boundary
densities for five macerals at 19 rank levels, based
between two different naturally occurring
on the percentage of total carbon. Table 3.12 gives microlithotype bands. At least 500 points should be
a comparison of a single maceral analysis reported
counted to achieve a satisfactory result, with care
on a volume percentage mineral-free basis, a being taken to ensure that the whole of the
volume percentage mineral- containing basis, and a
specimens surface is covered to avoid any bias
weight percentage mineral- containing basis.
due to particle segregation. The I.C.C.P. (1963)
and Stach et al (1975) describe conventions that
should be followed when the reticule does not fully
3.7.5 Microlithotype analysis
cover the image of a coal particle, and lies partly
on the binding material of the crushed coal
specimen.
The
second
method
of
performing
microlithotype analysis is termed selon la ligne
(according to the line). A simple eyepiece
measuring reticule is used, with divisions spaced
equally such that at least 20 fall within 50 pm. In
the case of particulate samples,
X

Micrinite
13

10

TABLE 3.12 Maceral analysis reported to different bases [high volatile B bituminous coal (PSOC-68)].
Vitrinite
Fusinite
Semifusinite
Macrinite Micrinite
Volume (%), mineral-free

81

Volume (%), mineral-containing


Weight (%), mineral-containing

79
75

8
9

4
4

1
1

1
1

Sporinite Resinite
3
3
3

Mineral

1
1

Fig. 3.13 Twenty point eyepiece reticule for microlithotype analysis.

the reticule is rotated so that the line lies perpendicular to


the bedding displayed in any particle. Microlithotype
identification is made by considering the natural band5 of
50 pm in width containing the mid-point of the reticule.
The 5% limit is determined by estimating the proportion of
macerals or minerals across the width of the band under
consideration. Where stratification is uncertain, the reticule
should
t__

_______J :

________i______________________________*.u~

uc maiiuanicu m me vcmuai piuMiiuii aiiu uic

estimation made on a circle of radius 50 pm.


The selon la ligne5 method is also particularly suitable
for obtaining microlithotype profiles of column samples of
coal. A traverse can be made down the entire column of
coal, a reticule length at a time.
Analysis by either method on the same coal will give
different results, because the 20 point ocular method,
which does not recognize natural bands5 in the sample,
tends to produce more bimaceral and trimaceral categories.
J.W. Hunt (personal communication) has, in comparing the
two methods on the same sample, measured the amount of
vitrite + duroclarite + clarodurite by the selon la ligne 5
method at 30%, compared to 45% by the 20 point ocular
method.
3.7.6

Reflectance measurement

The light source must be stabilized to ensure that no


variations in photomultiplier output arise due to minor
fluctuations in light intensity with surgesin the power
supply. Because the amount of light reflected from the coal
is rather small (usually less than 2%) the power rating and
intensity of the light source should be high to increase the
signal-to-noise ratio of the system. A 100 W quartz-halogen
lamp is most commonly used for this purpose.
If maximum5 rather than random5 reflectance is to be
measured, a polarizer is placed into the
<4
1
T
+ *>
4 /\
f/ \
f U A
U1G1UC1U
llglll
UCcUll.
iCll C^ LCU Ui l
IU

*/ %

J_tlglll
lli^

Id

specimen surface by means of a vertical illuminator. In


microscopes that use a Berek prism as the vertical
illuminator, the polarizer should be set in the 45 position
(Berek 1937; Broadbent & Shaw 1955).
The reflected light from the surface of the coal passes
back through the objective to the vertical illuminator.
Between the vertical illuminator and the photomultiplier
head are a filter and a limiting aperture. The filter is usually
a band interference filter of 546 nm, which is the
wavelength of the mercury line within the green region of
light. The aperture, which may be of fixed or variable
dimensions, limits the area of the field on which the
reflectance measurement is actually taken. A square of
about 4-5 ^m on the side is most commonly used. In many
modem photometric systems, an image of the aperture can
be seen superimposed upon the actual area of measurement
within the field of view.
The measuring device used in most reflectance systems
is a photomultiplier tube, selected to give

A number of methods, both direct and indirect, have been


employed to measure the reflectance of coal macerals, and
these are reviewed by Davis (1978c). Of these techniques,
however, the single-beam comparative method is now used
Ct illgll itopuiiov UL
11111. l-'lglLai ^HVLWll VVSMlllW
by most, if not all, coal petrologists throughout the world.
The single-beam comparative method is based on the (Gray Todd & Drexler 1979) and photodiodes (Otte &
use of a reflected-light microscope photometer, calibrated Pfisterer, in preparation) have also been used as satisfactory alternatives for this
by means of glass, mineral or synthetic standards, which
purpose. A stablized power
are then replaced by the vitrinite in the coal sample for a
supply is required for the
series of reflectance readings. Detailed descriptions of the
photometer. A chart, meter
equipment and procedures used in the process are given by
I
the I.C.C.P. (1971), Galopin and Henry (1972), Piller
(1977), A.S.T.M. (1981c) and I.S.O (in press, c). The
components of a reflectance photometer are illustrated in
Quartz
halogen
Fig. 3.14.
n

mnnAnna nt

n m T^irritvll nVlAtAn AA11 n tpl*c

lamp

Photomultiplier cathode ^
Interference filter ezzza
Magnification system * with limiting aperture
Polarizer

'

Dl'ild \
Vertical illuminator
Objective-^--^
Specimen

*rr77-rr

Fig. 3.14 Optical components of a microscope photometer.

Coal Petrology and Petrographic Analysis

39

or computer is needed to observe and/or record the


reflectance readings, and a data processor is often an
integral part of the reflectometer system.
The scale on which the reflectance measurements are
read is established from zero to the reflectance of the
primary standard, preferably one which has a reflectance
just slightly above that of the subject vitrinite (Juckes
1973). The reflectance of glass standards (R) can be
calculated from the Fresnel equation (Section 3.6.1), but,
because there is some indication that the surface properties
of materials may also influence their measured optical
properties (Piller 1977), they may be calibrated by the
supplier. It is possible to obtain or prepare single mounts
which contain a number of such standards if required.
Once the upper limit of the scale has been calibrated the
zero should be corrected for any photomultiplier dark
current and back reflectance (primary glare) from the back
lens of the objective (Kotter 1960; Jones 1962; Galopin &
Henry 1972; Piller 1977). At least one additional standard
should be checked against the scale and the calculated (or
calibrated) and determined values should not differ by
more than about 0.02% actual reflectance.
Suitable vitrinite particles are selected for measurement
from a grain mount which has been carefully polished to
give a level, scratch-free and relief-free surface (A.S.T.M.
1981c). A recommended practice is to select the particles
from a grid pattern covering most of the grain mount. All
of the various vitrinite macerals and sub-macerals might be
included for applications where coal properties are being
related to technological behaviour, such as carbonization.
However, for precise measurements of the degree of
metamorphism it may be best to restrict the measurements
to bands of homogeneous telocollinite, which tends to be
slightly more highly reflecting, rather than the more
heterogeneous desmocollinite. In the case of lignites, the
preferred vitrinitic, or huminite maceral for reflectance
measurement is ulminite.
The number of readings that must be averaged to obtain
the mean reflectance may vary, depending on the
application, between about 20 and 100. Obtaining a
reflectance distribution chart on a blend of two or more
coals, however, may require 500 or more readings.
Checking and adjusting the calibration of the
microphotometer should be done at intervals during the
analysis, and also at its conclusion. The standard deviation
of the mean for 100 readings on a single seam should be
about 0.01-0.02% (I.C.C.P. 1971; I.S.O., in press, c). The
reproducibility between two different operators analysing
sub-samples of the same coal might be 0.08%.
In addition to reporting mean reflectance, it is also
common for the reflectance distribution to be reported as

V-types (vitrinoid types), or V-types, which represent


ranges 0.1 and 0.05%, respectively. For example, V-type 7
covers the reflectance range 0.70-0.79%, and % V-type
7.25 covers the range 0.70-0.74%. Figure 3.15 is a
histogram plot of a lh V-type distribution. Reference to
Table 3.10 (Section 3.6.3) indicates that this particular

sample is a blend of medium, high and low volatile


bituminous components. The use of such data in
establishing the coking characteristics of a coal or blend is
discussed further in Section 3.9.3.
3.7.7

Automated microscopy of coal

A number of methods are under investigation to provide


automatic microscopic measurement of the petrographic
properties of coal. The principal goal of research with these
systems is to develop the capacity of performing rapid and
accurate maceral analyses, based on the characteristic
differences in reflectance of the various components in a
single coal. These techniques have the potential to provide
improved reproducibility due to the removal of operatorinduced sources of variation, and also enable a greatly
increased number of sample points to be investigated ner
unit time.

---*------C?- *"

1*

Davis (1978c) has reviewed many of the early attempts


to automate coal petrographic analysis, including the
clockwork stage and recording device of Vendl (1934) and
the AMEDA, which incorporated ten electronic counters
and a belt-drive stage mechanism (Bomberger & Duel
1964; Bayer et al 1968). Most of these, however, were
limited by the data-handling capacity of the instruments
used. A

Fig. 3.15 1/2 V-type reflectance histogram for a blend of


high, medium and low volatile bituminous coals. Sample: coal blend. i?max = 1.245%.

'

system developed by the U.S. Bureau of Mines


(McCartney & Ergun 1969; McCartney et al 1971) also
employed continuous scanning, and recorded reflectance
data on magnetic tape for subsequent computer processing.
Three generations of automated microscopes, using both
continuous and stepping stage movement with data
processing in real time, are described by Davis and Vastola
(1977), Kuehn and Davis (1979) and Hoover and Davis
(1980).
Although the instruments used in modern automated
microscopy vary considerably, they can be classified into
two types, image-plane or specimen- plane scanners, on the
basis of the type of scanning technique employed (Weibel
et al 1972; Kuehn & Davis 1979). In image-plane scanners,
the scanning is performed by a sensor on the video
projection of the microscopic image, and for this reason,
techniques of this type are known as image-analysis
systems. Many thousands of points are analysed for each
field, but reproducibility of results is also a function of the
number of fields analysed during a scan. Such systems
have the advantage of being able to analyse the size and
shape of the particles under investigation. Various
applications of the technique to coal and coke petrography
are described by B.C.R.A (1975), Harris (1977), Harris et
al\ (1977), Zeiss (1979) and Chao et al (1979).

Coal Petrology and Petrographic Analysis

41
With specirnen-plane scanners, on the other hand,the
light source and detection equipment are aligned along the
optic axis of the microscope, and the sample is moved
across the field in a plane perpendicular to that axis. The
detector is usually a photomultiplier tube, and is capable of
resolving reflectance values more precisely than the
sensors in image-plane systems. A single reflectance
reading is taken over a small area (usually about 1-4 /im
across) after each increment of stage movement.
Specimen-plane analysis is performed on a linear or
raster pattern across the exposed polished surface with the
individual reflectance readings either immediately adjacent
to each other or separated by an interval of no more than
several tens of micrometres. In addition to accumulating
some thousands of individual data points, the analysis can
also, if adjacent readings are taken, provide information on
particle sizes based on intercept or chord-length
distributions. An instrument of this type, developed at the
Pennsylvania State University, is illustrated in Fig. 3.16.
The principal output of most automated microscope
systems is a reflectogram or histogram of the frequency of
readings at all reflectance levels (Fig. 3.17). Most
reflectograms have two prominent peaks, one at the low
reflectance end representing the binder used to pelletize the
coal particles, and the other representing the contribution
for the preponderant coal component, vitrinite. Because the
systems use non-

Fig. 3.16 Automated reflectance microscope consisting of Leitz Orthoplan microscope equipped with MPV 2 photometer system and stepping stage, and
Digital Equipment MINC-11 computer. (From Davis et al 1983.)

Attempts to derive actual maceral analyses from such


polarized light, the reflectance of the vitrinite mode is the
random reflectance. Hoover and Davis (1980) have shown data have involved curve-stripping techniques to overcome
how random reflectance can be converted to maximum the problems of overlapping reflectance distributions of
macerals and of boundaries (McCartney et al 1971; Kojima
reflectance by the following equation:
et al 1974; Hoover & Davis 1980; Kojima & Sakui 1980).
1.061 R where R and R , are Although themodels employed tend to oversimplify the
R max
rand
the
complex interactions of particle composition, size and
mean maximum
texture, some agreement with visually-derived analyses has
max
rand
reflectance and the mean random reflectance respectively. been obtained.
A major problem of automated techniques in
The data supplied to the computer in automated analysis
with visual microscopy
at present
is
are only a series of reflectance readings, and interpretation comparison
fAn
1 mriil
P
11 /rUf nn f
of which macerals these represent is usually made by
sole means of discrimination between coal constituents
interpreting the shape of the reflec- togram. The three
since, in visual analysis, the shape of entities, and even
reflectograms in Fig. 3.17 for example, are of
their modes of association with other materials may also be
petrographically very dissimilar lithotypes from a single
taken into account. Another problem has been the difficulty
seam section. The profiles clearly reflect the maceral group
in interpreting edge effects, the spurious readings
compositions as determined by conventional visual
obtained on the edges of particles and the boundaries
analysis and depicted in the accompanying bar graphs.
between macerals.
Automated reflectance microscopy is an ideal tool for
quality control evaluation of coals in industrial situations
(Davis & Vastola 1977). The reflectogram profile is
sensitive to changes in the composition of coal feedstocks,
especially in blends, and can be used to monitor quality in
various areas of coal utilization. Figure 3.18 gives an
illustration of the kind of advantage this approach has over
chemical analysis. The coal represented was marketed as a
midvolatile material, but the reflectogram shows it to be
mainly a blend of high and low volatile bituminous
components with only a minor amount of medium
HIV,

A
SJll Llll, I t V L l SJL luittltu ilgilL

Lilt

Fig. 3.17 Comparison of automated microscopy reflectograms with visual petrographic


analyses. (From Davis & Vastola 1977.)
*Vertical scale based on the most frequent reflectance reading, with a xlO factor for
reflectances greater than 0.12%. Ex., exinite; Vit., vitrinite; In., inertinite.

Coal Petrology and Petrographic Analysis

'

43

the kind of detail that can be reproduced in this way.

Average random reflectance


{%)

Fig. 3.18 Automated microscopy reflectogram of a blend of high,


low and medium volatile bituminous coals. (From
Hoover & Davis 1980.)

component, with only a minor amount of medium 3.8 ORIGIN OF MACERALS AND COAL LITHOTYPES
volatile coal.
The factors that influence the abundance and structure of
ViitrVi
r\f* nvritp ha also aided the
J
the constituents in coals of varying type are so interrelated
automatic measurement of the amount and size that it is difficult to identify a completely independent set
distribution of this mineral in coals (Bayer et al 1968; of controls. Factors that can have a bearing on the
McCartney & Ergun 1969; Davis & Vastola 1977; Kuehn constitution of a peat deposit include the following:
& Davis, 1979). Kuehn et al (1980), for example, have
(a) The nature of the plant community. The type of
used a specimen-scanning method to monitor changes in plants and the relative abundance of each form is an
pyrite content and petrographic composition with particle obvious control, depending in turn on the geologic age of
size and density in laboratory beneficiation experiments. the deposit, the physiographic setting, the
A new application for automated reflectance rlimafp anH tbp snnnlv nf nutrients:
w.-------------------------------------r r ~ j ~ ~ --------------------------- >
microscopy (Davis et al 1983) has been prompted by a
(b) The climate prevailing in and around the
need to map the petrographic variation across coal depositional site. The temperature and humidity, as well as
surfaces. Reflectance values are stored on disc in a any seasonal fluctuations in these factors, influence the
matrix that enables the proper spatial arrangement of nature of the plant community, the extent of peat build-up
values to be recreated as a map by an image processor and decay, and the rate and products of weathering in the
and colour camera. The maps depict lithotype variation, swamp hinterland;
and the distribution of mineral occurrences, cracks and
(c) The extent of plant decomposition. This in turn
cleat in the coal. They can be matched by computer depends on the nature of the plants themselves, on the
against images derived by other techniques. Figure 3.19 climate and on the Eh and pH conditions in the swamp
compares the reflectance map with a conventional waters;
photograph, and illustrates
(d) The tectonic setting of the deposit. This is a major
1 UV lilgli X VllVVtUilVV V*
---

* A

factor in controlling subsidence rates and in determining


the physiographic setting, the rate of nutrient supply and
the extent of plant decay;

(e) The physiographic setting, palaeogeography or


depositional milieu of the deposit. These represent a
combination of factors including the depth and movement
of waters, the shape and areal extent of the peat swamp and
the chemistry, including Eh and pH, of the swamp waters.

Fig. 3.19 Comparison of image derived from automated microscope reflectance mapping (b) with photograph (a) of a polished coal surface.
(From Davis et al 1983.)

'

3.8.1
Origin of vitrinite and bright coal
lithotypes
Vitrinite has long been regarded as the result of deposition
of ligno-cellulosic tissues (wood, bark etc.) in stagnant,
highly toxic waters that protected the organic material from
extensive biochemical decay (White 1933; Tasch 1960).
Raistrick and Marshall (1939) have noted that the great
majority of vitrain sheets in Carboniferous coals, which are
usually about 6 mm thick, represent the bark shells of
lycopods, although the wood of gymnosperms and cycads
was another important source of this type of material.
On the basis of detailed petrographic and palynologic
profiles through a number of coal seams, Smith (1968)
suggests that the vitrinite-rich layers in humic coals were
most likely to have been deposited in areas that underwent
greater subsidence than the surrounding regions. In studies
of Australian coal measures, Shibaoka and Smyth (1975)
have shown that coals deposited in the thicker sections of
troughs (i.e. areas of greater relative subsidence) are rich in
bright lithotypes, and that these are more likely to be
associated with mineral partings than the duller, vitrinitepoor seams of the more stable shelf areas. Shibaoka and
Smyth (1975) and Cook (1975) noted that the lower seams
of the Australian Newcastle Coal Measures and the Triassic
Ipswich Coal Measures, which are deep basin type
deposits, are rich in vitrinite, whereas the Illawarra Coal
Measures, deposited with steady, slow subsidence,
contained seams typically very poor in vitrinite. Other
shallow- basin type coals deposited on stable shelves or
basement rocks, like the Blair Athol, Leigh Creek and
Callide coals, also are vitrinite poor.
The fine-grained type of vitrinite, desmocollinite, such
as is commonly found in clarain, is generally assumed to
have originated either from smaller plant organs, such as
leaves, or from finer fragments of larger tissues that have
been partly degraded. Desmocollinite usually has a lower
reflectance than the telocollinite (i.e. coarse vitrinite) in the
same coal, and this has been attributed to either admixture
with fine liptinitic detritus (Taylor 1966) or to the adsorption of bitumen, possibly lipoid substances derived in part
from liptinite, into the macerals structure (Teichmuller
1974b; Spackman et al 1976). Some desmocollinite also
exhibits fluorescence characteristics, probably also
reflecting the presence of bitumen impregnations. A
similar explanation would account for the unusual
fluorescence, reflectance and chemical characteristics of
some jet, vitrain lenses derived from coniferous wood and
found in marine sediments with a high lipid content
(Traverse & Kolvoord 1968; Given et al 1975; Spackman
et al 1976; Davis 1978b).
Another hypothesis for the higher reflectance of
telocollinite (or vitrinite A, pseudo-vitrinite, band
vitrinite) is that the plant material has undergone oxidation,
probably during the early stages of its development
(Benedict et al 1968). Koch (1970), however, suggests that
it is due to the nature of the original plant material
involved, noting that, in the Jurassic coals of Afghanistan,
this maceral type had the characteristics of coniferous

Coal Petrology and Petrographic Analysis

tissues. Cook (personal communciation) has suggested 45


that
any chemical differences between band vitrinite and
vitrinite in attrital layers may have been imposed by
differences in the chemistry of the original plant materials,
with the tissues that gave rise to the band vitrinite being
hydrogen deficient and therefore oxygen rich. Such
material would have decomposed more rapidly than the
hydrogen-rich tissues, and thus band vitrinite would have
been preserved only under rapid burial conditions.
3.8.2
Origin of inertinite and dull coal
lithotypes
The inertinite macerals are generally regarded as having
been derived from the same types of plant debris as
vitrinite, but have undergone chemical changes due to
processes such as charring, oxidation, mouldering and
fungal attack at an early stage of their depositional history.
Fusinite, for example, is thought to have originated in
many cases as plant fragments that have suffered partial
combustion in forest fires. Such material, known as
pyrofusinite (Teichmuller 1950), usually has a high
reflectance, appearing yellow or white under oil
immersion, and with distinct cell wall margins. Another
form, called degradofusinite, is characterized by
indistinct cell walls and a lower, semifusinitic reflectance.
This material is thought by Teichmuller to result from
dehydration, oxidation and biochemical alteration of plant
debris. However, Given (in Given et al 1980) has
challenged the possibility that degradofusinite might be the
product of biochemical decay. It is -a characteristically
abundant constituent of the Permian coals of
Gondwanaland. Other fusinite may be of primary origin,
derived directly from certain pigmented plant tissues.
Semifiisinite, in general, is regarded as representing
ligno-cellulosic plant debris that has suffered a lesser
degree of degradation than fusinite. Fragments of both
fusinite and semifusinite, often forming discrete
inertodetrinite particles, can also be transported some
distance from their place of formation, accumulating in
quiet, sub-aquatic conditions with other particles such as
spores or algae and possibly fine mineral matter.
Sclerotioids (secretion sclerotinite or carbonized resin
bodies) are also thought to result from oxidizing conditions
in the peat swarnp, although such materials may also be
transported from an oxidizing environment nearby. True
sclerotinite represents fungal material and is also formed in
relatively dry oxygenated peat layers (Stach et al 1975).
_ Micrinite may possibly be formed in two ways. It may
represent a by-product of the coalification of liptinite
macerals, including bituminite (TeichmiiUer 1974b) and
exudatinite (Shibaoka 1978). The suggestion is supported
by its frequent association with spo- rinite. Although
Teichmuller indicates that micrinite is not normally found
in peats or brown coals, others (e.g. Spackman &
Barghoorn 1966) have reported it in such materials.
Together with the fact that it frequently occurs as cell

fillings in higher rank coal, this suggests that some


micrinite may also be formed by the breakdown of cell
wall substances.
The origin of macrinite, a relatively rare component of
most coals, is not well understood. At least some of it may,
however, result from the oxidation (or fusinization) of
organic gels formed during peat accumulation.
Because they are made up mainly of fine inertinite
debris and have a relative abundance of spore material, the
black durains of many humic coals are classically regarded
as sub-aquatic ooze deposits, forming in deeper water
environments than the bright, vitrain- rich lithotypes
(White 1933; Tasch 1960). Hacquebard et al (1967) have
generally indicated support for these concepts, interpreting
the high telinite (bright coal) contents of some Canadian
seams as representing forest moor environments, typical of
the central part of a flood plain. The duller intervals of
those seams, with relatively high liptinite and mineral
contents, are thought to be the product of an open moor
environment with a generally higher water table. As an
alternative, however, Smith (1968) has suggested that black
durains, rich in macrinite and thick-walled crassispores,
were formed in a raised, well-drained peat swamp
environment. Tenuidurains, in which thin- walled spore
types are typically associated with semifiisinite and a high
proportion of mineral matter, are interpreted by Stach et al
(1975) as representing a horizon of peat oxidation above
the water table. Smith (1968) on the other hand interprets
them as representing a sub-aquatic incursion phase due to
flooding of the peat swamp.
The presence of a large amount of mineral matter
in a coal may also reflect the development of deep-water
conditions. However, some workers suggest that such
concentrations could also result from sub-aerial ablation
of peat, or from excessive peat decay in swamp waters of
pH greater than 4.5 (Cecil et al 1979).
3.8.3
Origin of liptinite macerals and
sapropelic coals
The organic components that give rise to the various
liptinite macerals are described in some detail in Section
3.3.3. Unlike the ligno-cellulosic materials that make up
vitrinite, the liptinite precursors are largely resistant to
degradation and to the glification or similar processes
generally associated with peat formation. They do,
however, exhibit a number of significant changes during
coalification or rank advance, and these are discussed
more fully in Section 3.6.4.
Cannel coals are believed to be derived from the more
resistant components of decayed plants, especially spore
remains, that were washed into ponds, lakes and lagoons
to accumulate as fine grained, anaerobic bottom muds.
Boghead coals or torbanites, on the other hand, are
thought to represent circumstances in which clear,
aerated surface waters, free of humic matter, allowed

algal colonies to flourish (Moore 1968). After death, the


remains of these colonies accumulated in euxinic bottom
oozes, eventually to become masses of alginite or
bituminite.
3.8.4
Temporal and climatic influences on
peat accumulation
The coals of different geologic periods were derived
from different plant assemblages, and under differing
climatic conditions. These variations have clearly
produced some diversity in the petrographic character of
coals of different ages.
The plants of the Carboniferous, including such groups
as the seed ferns, lycopods and sphenopsids, were
commonly prolific producers of thick-walled megaspores
and microspores. After the Carboniferous, however, the
importance of these plants declined, with many becoming
extinct, and the gymnosperms . (which include the seed
ferns) became the dominant members of the plant
community. These plants produced smaller, thinner-walled
palynomorphs than did the Carboniferous plants, and
consequently contributed less sporopollenin to the resulting
peats than the larger, thicker-walled spores common in the
Carboniferous. More importantly the number of
megaspores produced was reduced to four, with three of
these aborting and then being retained in the
megasporangium without developing an exine of liptinitic
source material.
The angiosperms have been important members of
swamp and marsh communities from the Tertiary to the
present. All of these plants retained the megaspore in the
ovule with no production of sporopollenin, and with a
reduction in total pollen production compared to many
gymnosperms. There has therefore been a progressive
decrease in the contribution of spores and pollen to the
coal-forming swamps from the Carboniferous to the
Tertiary, and a change from large, thick-walled
palynomorphs to small, thin-walled palynomorphs in the
liptinites of the coals produced through time (Waddell et al
1978).
Gondwana coals, formed in the southern hemisphere
mainly during the Permian Period, were derived from the
well-known Glossopteris flora. This assemblage was
dominated by gymnospermous plants and more restricted
in species than the flora which inhabited the Carboniferous
forests. The climate for these deposits is believed to have
been mainly cool temperate, in contrast to the sub-tropical
conditions that probably existed during much of the
European Carboniferous.
Some of the characteristics of the Gondwana coals
include a much lower sporinite content than is found in the
Carboniferous deposits, and a higher inertinite content,
made up mainly of semifusinite that formed under
oxidizing conditions. Cook (1975) has suggested that a dry
climate may have been responsible for the abundance of

'

inertinite-rich coals in the late Permian coal-bearing


sequences of Australia.
As stated in Section 3.3.3, thick cuticles may be
indicative of a dry, warm climate. Cutinite is rare in
Australian Permian coals, but is quite common in the
Triassic Ipswich Coal Measures and Jurassic Walloon Coal
Measures of Queensland, especially in the Jurassic
Rosewood and Darling Downs deposits. Resinite is found
in larger quantities in Cretaceous and Tertiary coals than in
those of the Carboniferous because of contributions from
the resins of conifers in the later-formed materials.
Suberinite is a component of Mesozoic and, particularly,
Tertiary coals. True fungal sclerotinite is also a ubiquitous
component of Tertiary coals, but is only rare in Carboniferous deposits.
3.9 APPLICATIONS OF COAL PETROLOGY
Coal petrology represents one of the main avenues for
gaining an understanding of the origin of coal deposits, and
also provides invaluable information in establishing the
most appropriate means of using the various coals formed
by these processes. Petrologic data, collected on either a
megascopic or a microscopic scale, can be used to help
solve problems with coal seam correlation, and may be
considered in conjunction with other geologic factors to
assist in interpreting the location, extent and quality of
economic coal resources. Systematic data on coal rank,
compiled from features such as vitrinite reflectance, can
also be used to establish the tectonic history of the region,
and, as a by-product, assess the likelihood that petroleum
hydrocarbons might have been generated at some stage of
basin development as well.

Coal Petrology and Petrographic Analysis

47
3.9.1 Seam correlation and other uses of petrographic
profiles
Correlations may be drawn from petrologic data wherever
a coal seam laterally maintains a unique characteristic, such
as a particular vertical sequence of lithotype variation.
Hacquebard (1952) has described how the distinct and
uniform petrographic composition and seam profile of the
Tracy Seam of the Sydney coalfield, Nova Scotia could be
used for correlation over distances of at least 12 km, while
Smyth (1967) has suggested some correlations between
seams of the Newcastle and Illawarra Coal Measures of
New South Wales on the basis of their petrographic
profiles. Davis (1968) has also noted that two vertically
adjacent seams at Moura, Queensland, had such distinct
profiles (Fig. 3.20) that, even from megascopic data, it was
possible to identify which of the two seams was present in
a situation where the stratigraphic sequence had been
complicated by faulting.
In some instances, although the profile itself may vary,
the presence of persistent horizons within the seam can be
a considerable aid in correlation. Hacquebard et al (1965)
and Cameron (1971) were able to trace individual durain
layers in the Harbour seam of Nova Scotia across the width
of the Sydney coalfield, a distance of over 32 km, even
though the durains themselves changed laterally. A fusain
layer in the No. 12 seam of Western Kentucky also extends
over an area greater than 130 km2 (Austin & Davis 1979).
Inorganic bands and partings within seams are
commonly useful marker horizons as well. The prominent
blue band, a claystone parting of the Herrin (No. 6) seam
of the Illinois Basin, has been traced over thousands of
square kilometres in Illinois (Willman et al 1975) and
western Kentucky, while kaolinite clayrocks or tonsteins
(Section 5.3.1) have been employed as markers both within
and between individual coalfields. Scheere (1956) has
demonstrated the correlation of such horizons from France,
through

200 150 100

50

^====^=======i cm

200

Vertical scale

Fig. 3.20 Petrographic profiles of two coal seams from Moura, Bowen Basin, Queensland. (From Davis 1968.) ^ Vitrinite; S semi-inertinite;
inertinite; exinite; mineral.

Belgium and Holland and into Germany, covering a


distance of about 400 km, and in Queensland, Australia,
such horizons have been traced in individual seams
(Davis 1972, 1973; Carr & Davis 1973; Beeston 1974).
Six carbonaceous shale partings also occur over an area
greater than 1500 km2 within the No. 12 seam of western
Kentucky referred to above (Austin & Davis 1979).
The vertical succession of lithotypes in seams has
been used to study the variation in environmental
conditions during coal formation. For example,
Hacquebard et al (1967) noted the lack of petrographic
variability in coals of limnic sequences in contrast to the
greater variability of those in paralic coal measures. The
predominant seam profile in Australian coals is one in
which the abundance of vitrinite decreases from floor to
roof, and this is interpreted by Smyth (1970) as
representing a rise in the depositional surface as a result
of the development of progressively drier conditions in
the peat swamp.
3.9.2 Applications of reflectance measurements

As indicated in Section 3.6.3, the principal use of


reflectance data is as a measure of coal rank. It has long
been established that rank increases with depthin
stratigraphic successions or borehole profiles (a
phenomonen described traditionally by Hilts Law), and
this is thought to reflect the effects of rising temperature
as the thickness of overburden increases.
(a) Reflectance profiles
Reflectance profiles in boreholes have been used to study
the tectonic histories of coal-bearing sequences. Dow
(1977) has shown the effects of such processes as loss of
cover, the development of unconformities, igneous
intrusion and faulting on the profiles of organic
maturation in sedimentary sequences. The temperature
and time required for coalification have been related to
the rank or level of metamorphism by authors such as
Karweil (1956), Lopatin and Bostick (1973), Hood et al
(1975) and Bostick et al (1979). Figure 3.21 is a
nomograph, modified after Hood et al (1975), that
enables palaeotemperatures to be derived from
reflectances, provided the effective heating time can be
estimated. The effective heating time is the time during
which the sediments were exposed to temperatures within
15 C of the maximum attained.
In conjunction with the depth of the seams in
question, such data can be used to calculate
palaeogeothermal gradients and hence allow the

Fig. 3.21 Relation of vitrinite reflectance to maximum temperature (Tmax) and effective
heating time. (Modified from Hood et al 1975.)

thermal and tectonic histories of various coal-bearing


sedimentary basins to be progressively reconstructed.
Rank patterns and changes in geothermal gradient have
been related to depth of burial, depth of basement (Cook
& Kanstler 1980), basement lithology (Koppe &
Anderson 1974; Anderson & Koppe 1976) and the
thermal properties of the rocks in question (Teichmuller
& Teichmuller 1968; Facer et al 1980). They have also
been related to thermal events such as the initiation of
crustal rifting (Cook & Kanstler 1980) and to differences
in heat flow between adjacent crustal blocks (Hower &
Davis 1981b).
(b) Isoreflectance maps
Lateral variations in the rank of a particular seam or
reference horizon across an area may be expressed by
contoured isoreflectance maps (Section 6.7.2).
Bartenstein et al (1971), for example, have mapped
changes in rank due to thermal metamorphism of
Carboniferous and Cretaceous sediments across the
Middle Cretaceous pluton of the Bramsche Massif in
north-western Germany, and Beeston (1977) has mapped
vitrinite reflectance trends along the 75 km outcrop belt
of the German Creek seam in Queensland, Australia. In
the latter case the rank of the coal varies from subbituminous to semi-anthracite, and it was possible, by
consideration of the palaeogeothermal gradient, to
calculate the thickness of overburden that had been
removed from the area of highest rank material.
Hower and Davis (1981a) have published a quadratic
trend surface map showing the reflectance

of coals, mainly of Allegheny age, in western


Pennsylvania, while Senftle and Davis (1982) have
mapped the reflectance of the Lower Kittanning seam in
the same area. Facer et al (1980) have produced a third
degree trend surface map of the lateral variation in
reflectance for seams in the Illawarra Coal Measures of
New South Wales. Isoreflectance maps can be used in the
planning of mining operations to optimize the supply of
coal blend components for metallurgical coke production
(Benedict & Thompson 1978).
(c) Interpretation of stress patterns
In another type of study, the orientation of the principal
reflectance directions in coal specimens has been used to
interpret the directions of tectonic stress in several
coalfield areas. Hower and Davis (1981a) and Levine and
Davis (1984), for example, found that the maximum
retlectance of tightly folded biaxial coals of low volatile
bituminous and anthracitic rank in Pennsylvania was
oriented sub-parallel to the fold axes, while Stone and
Cook (1979) observed that the biaxial reflectance
anisotropy of samples from the proximity of normal
faults in the Southern Coalfield of New South Wales
could be related to probable local stress patterns
associated with the faulting. Studies of this type can be
applied to a number of problems with the stability of
underground mining operations, such as those discussed
in Section 7.6.
Although reflectance anisotropy generally increases
with increasing rank (Section 3.6.2), values of
bireflectance obtained from vitrinite in the Bowen Basin
of Queensland show fairly uniform values

throughout much of the rank range in that area (Davis


1971). Hower and Davis (1981b) found that another
index of anisotropy,
R ~~ R

max

R~

rand

remained more or less constant along traverses from west


to east in western Pennsylvania, whereas vitrinite
reflectance generally showed marked increases. They
interpreted this to indicate that the past depth of burial
was also uniform across the area.
(d) Generation of petroleum hydrocarbons
In 1915, White observed that the presence, amount and
type of petroleum in sedimentary rocks could be related
to the rank of any associated coals in the sequence. Rank
at that time was defined solely by chemical parameters,
and the resulting theory has therefore become known as
the carbon ratio theory*. It has been investigated in
many parts of the world, and several of the findings of
these studies have been compiled by Bostick and
Damberger (1971).
Briefly, the theory suggests that a certain level of
maturation (i.e. rank) is necessary for organic matter in
the source rocks to generate hydrocarbon liquids. As the
rank increases, however, the liquids become lighter and
ultimately only natural gas is produced under the
prevailing temperature conditions. Figure 3.22 gives a
summary of correlations between coal rank (vitrinite
reflectance) and the zones of petroleum generation, and
also shows the death line5, where the rank becomes too
high for liquid hydrocarbons to form.
As well as the rank, however, the composition of the
organic matter (kerogen) in the sediments is also
important in determining the type and amount of
petroleum that might have been generated. Kerogen
derived from algae and marine organic matter will
generate oil and gas in relatively large quantities, whereas
humic kerogen, mostly derived from terrestrial plants, is
thought by some authors to generate only minor amounts
of gases (Dow 1977; Tissot & Welte 1978). In addition,
the source rocks must contain sufficient kerogen (at least
about 0.5% organic matter) to provide for the possible
accumulation of petroleum in the first place.
3.9.3
The petrographic prediction of coke
properties

In the 1960s, the work of Ammosov et al (1957) and


Schapiro et al (1961) gave a new stimulus to thepractice
of coal petrology through the development of procedures
whereby the quantitative strength characteristics of
metallurgical cokes from single or blended coals could be
predicted. Such procedures are described more fully in
Section 4.3.6. Although these empirical methods have
generally been successful, some of the concepts
incorporated in the original mathematical models are not
accepted universally, and some care should be taken in
their use.
Depending on the rank of the coal, the models are
based on the assumption that vitrinite, liptinite and
possibly some of the semifusinite react in the process of
coke formation and bind the other inertinite components
and mineral matter into a porous fused solid. The model
devised by Schapiro et al (1961) predicts coke strength
on the basis of the relative abundance of the reactive and
non-reactive constituents, and also on the abundance of
vitrinite in reflectance classes that correspond to the
vitrinoid types (V-types) discussed in Section 3.7.6.
However, the model makes an arbitrary subdivision of the
semifusinite, regarding one-third of that material as
reactive and two-thirds as non-reactive, a procedure that
is not in accord with observations of Taylor et al (1967),
which suggest that this maceral does not become plastic
during coking. It is also not always desirable to treat
liptinite in the coal
_______^1__ ..
as
a
reactive
maceral in the same sense as the viLimiic, since me upimiLe
is mosiiy uevoiaimzeu in the carbonization process and
contributes little to the final product. The breakdown of
vitrinite reflectance data into V-types, although it
represents a practical way of dealing with the coking
properties of the vitri- nites of different rank from the
various coals in a blend, also makes arbitrary and
unproved distinctions between and within the vitrinite
types that might be present in a single coal.
Other methods of coke strength prediction include
those of Benedict et al (1968), which allows for the
poorer caking characteristics of pseudovitrinite 5
compared to reactive5 vitrinite, and Mackowsky and
Simonis (1969), which incorporates parameters of coking
conditions into the prediction. Gray et al (1979) have
described how results using the method of Schapiro et al
(1961) may be modified to allow for the coking
conditions.
Petrographic methods also have been established to
predict the pressures and expansion or contraction
developed during carbonization, and the reactivities of
cokes (Benedict & Thompson 1976; Thompson &
Benedict 1976; Gray Goscinsky & Shoenberger 1979).
J

3.9.4
The microscopic detection of coal
oxidation

The oxidation of coals, caused by weathering in

Fig. 3.22

Correlation of coal rank and vitrinite reflectance with the zones of petroleum generation and destruction. The relative importance
of the zone of petroleum generation depends upon the composition of the original kerogen. (Modified from

outcrop or stockpile, can result in a deterioration of the


coking properties of bituminous coals, and in detrimental
effects on the specific energy and preparation
characteristics of coals (Section 4.6). Consequently, the
detection of oxidation is an important aspect of the
evaluation of coals for industrial usage. There are certain
microscopic features of weathered coals which are easily
recognized. These are the discolouration of particles
(especially around the peripheries), the presence of
microfractures and unusual relief within the weathered
macerals

(Benedict & Berry 1964; Crelling et al 1979). Figure 3.5f


is a photomicrograph of weathered coal which displays
these features.
Gray et al (1976) have described a technique for
staining coals with an alkaline solution of the red stain
Safranin 0. Oxidized particles appear green under the
microscope. The test, as described, is not reliable for
coals higher in rank than high volatile bituminous;
however, a modification which overcomes this limitation
has recently been developed (Gray, personal
communication).

* The descriptions of the microlithotypes refer to their analysis by the 20 point ocular method (Section 3.7.5).

You might also like