Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/261850536

Chapter 3: Geological Sources of Metals in Coal and Coal Products

Chapter · January 2005

CITATIONS READS

2 591

3 authors:

Colin R. Ward Zhongsheng Li


UNSW Sydney The Commonwealth Scientific and Industrial Research Organisation
208 PUBLICATIONS   8,439 CITATIONS    67 PUBLICATIONS   1,226 CITATIONS   

SEE PROFILE SEE PROFILE

David H. French
UNSW Sydney
144 PUBLICATIONS   3,237 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Non-mineral inorganics occurrence in macerals of low-rank coals View project

Geochemistry of Oxide Minerals View project

All content following this page was uploaded by Zhongsheng Li on 25 April 2014.

The user has requested enhancement of the downloaded file.


Chapter 3: Geological Sources of Metals in Coal and Coal Products

3
Geological Sources of
Metals in Coal and
Coal Products
COLIN R. WARD1, ZHONGSHENG LI1
and DAVID FRENCH2
1
School of Biological, Earth and Environmental Sciences,
University of New South Wales, Sydney 2052, Australia
2
CSIRO Energy Technology, Lucas Heights Research Laboratories,
Menai 2234, Australia

INTRODUCTION
Coal is a major source of the world’s energy requirements, accounting for 37%
of the electricity generated and, through the use of metallurgical coke, some 70% of
world steel production. The International Energy Agency (IEA, 2001) has forecast
that the world energy requirements to 2020 will increase by 75%, with coal ac-
counting for approximately 45% of the primary energy mix. Given that coal is likely
to remain a significant source of energy for the foreseeable future and that its use
is actually likely to increase, it is important to understand the mode of occurrence
and behaviour of potentially harmful trace metals during coal utilisation in order to
minimise the environmental footprint associated with coal use.
For many purposes coal may be regarded as consisting of two classes of material,
the organic component (macerals) on one hand and the inorganic fraction (mineral
matter) on the other. It is the organic fraction that is most useful from the utilisation
viewpoint (Berkowitz, 1979), whereas the mineral matter mainly serves as a dilu-
ent. The mineral matter is the source of the ash produced during coal combustion
(Raask, 1985; Gupta et al., 1999), and contributes to the slag that is formed during
steel manufacture. It is also the main source of many of the potentially environmen-
tally harmful elements, such as mercury, which are becoming of increasing concern
to regulatory and health authorities (Swaine, 1990; Swaine and Goodarzi, 1995;
Finkelman et al., 2002).

NATURE AND ORIGIN OF COAL


Coal is a combustible sedimentary rock, composed essentially of lithified plant
debris (Diessel, 1992; Rapp et al., 1998; Taylor et al., 1998). The plant debris was

Ward, C.R., Li, Z. and French, D., 2005. Geological sources of metals in coal and coal products, in Metal
Contaminants in New Zealand, T.A. Moore, A. Black, J.A. Centeno, J.S. Harding, D.A. Trumm (Eds.),
resolutionz press, Christchurch, NZ, p. 49-79. © 2005 resolutionz press. All rights reserved.

49
METAL CONTAMINANTS I N N E W Z E A L A N D

originally deposited in a swampy depositional environment to form soft, spongy


sediment called peat. Physical and chemical processes brought about by compaction
and elevated temperatures with prolonged burial, however, changed the peat into
coal through a process of geological metamorphism referred to as coalification or
rank advance.
The properties of a given coal can be related to three independent variables, each
of which is determined by some aspect of the coal's origin:

Rank
Coal rank reflects the degree of metamorphism (or coalification) the original
mass of plant debris (peat) has experienced during its burial history. This depends on
the maximum temperature to which the coal has been subjected and the time it has
been held at that temperature, and hence reflects the depth of burial and geothermal
gradient prevailing for the sedimentary basin concerned.
The rank of a particular coal is indicated by a combination of properties (Fig.
1), including the carbon content of the mineral-free or dry, ash-free (d.a.f.) mate-
rial, the amount of heat energy released per unit mass when the mineral-free coal
is burned, and the proportion of light reflected from polished surfaces of particular
coal components (usually expressed by the reflectance of the contained vitrinite)
using special microscopic techniques (Taylor et al., 1998). Lower-rank coals include
materials described as lignite and sub-bituminous coal (Fig. 1), while higher-rank
coals include bituminous coal and anthracite. Both lignite and sub-bituminous coal
may be grouped together as brown coal.

Type
Coal type reflects the nature of the plant debris (peat) from which the coal was
derived, including the mixture of plant components (wood, spores, algae etc) in-
volved and the degree of oxidative degradation they suffered before burial.
The type properties of particular coals are reflected in the relative proportions of
the different plant-derived constituents, or macerals, that make up the organic frac-
tion. These are classified under three groups: vitrinite (also referred to as huminite in
lower-rank coals), representing well-preserved or gelified woody tissue; liptinite (also
called exinite), representing small waxy plant components such as spores and pol-
lens, algae, resin and leaf cuticles; and inertinite, which mainly represents oxidised
remains of woody material but may also include certain types of fungal remains
(Diessel and Smyth, 1995; Taylor et al., 1998; Scott, 2002). The chemical and other
properties of the individual macerals within these groups change progressively with
rank advance, and it is the combination of rank and type that fundamentally control
the overall nature of the organic matter in a particular coal seam.

Grade
The grade of a coal reflects the extent to which the accumulation of plant debris
has been kept free of contamination by inorganic material (mineral matter), both be-
fore and after burial and during rank advance. The mineral matter in coal produces
a non-combustible residue, or ash, when the coal is burned (Raask, 1985; Gupta et

50
Chapter 3: Geological Sources of Metals in Coal and Coal Products

�����������������
��������������
����������
��������

��������

���������
��������

������

������
������
��������

��������
�����
���
����

����� �������
���� ����
����������
����
��������
����������

������
����

��������
���
��������
���������������
����������
��� ��� ���� ����
����������

FIGURE 1: Variation of selected coal properties with rank advance (Ward, 1984).

al., 1999). A high-grade coal is therefore a coal with a low ash yield, while a low-
grade coal typically leaves a high proportion of ash after the combustion process.

Other Significant Coal Properties


A large number of other chemical and physical properties are also important
in evaluating the quality of coal for particular purposes (Rees, 1966; Karr, 1978;
Ward, 1984; Standards Australia, 1993; Thomas, 2002). These include the propor-
tion of moisture contained within the coal (either as-received or after air drying)
and the proportion of volatile compounds (volatile matter) released when the coal
is heated to high temperatures in the absence of air or oxygen. Both these proper-
ties, along with the hydrogen content of the organic matter, typically decrease with
rank advance.
The relative proportions of sulphur and nitrogen in the coal are also significant,
since these elements can form undesirable emissions (sulphur and nitrogen oxides)
when the coal is burned. The relative proportions of sulphur occurring in pyritic,
sulphate and organic forms may also be determined in some coal analyses. Sulphur,
along with phosphorus, is also undesirable in coals to be used for metallurgical ap-
plications, particularly iron and steel production. These elements are transferred from
the coal to the coke and from the coke to the iron and steel, where they are undesirable
contaminants that need to be removed from the metal in later production stages.
Other properties include the amount of heat energy liberated per unit mass when
the coal is burned (specific energy or calorific value) and the ease with which the coal
is ground to powder for particular types of combustion processes (grindability in-
dex). The porosity and strength of the coke produced from the coal, expressed by the
crucible swelling number and other types of tests, is also important in a coal to be

51
METAL CONTAMINANTS I N N E W Z E A L A N D

used for metallurgical applications (Ward, 1984). An indication of the significance of


these properties in terms of utilisation for different markets is given in Table 1. The
nature of the mineral matter, the composition of the coal ash, and the proportion of
particular trace elements, are also very significant for a variety of purposes; these are
discussed in more detail elsewhere in this Chapter.

WORLD COAL RESOURCES

Although carbonaceous beds occur in rock units as old as Precambrian in age,


coal was not formed as such until the evolution of land plants in the Silurian and
Devonian Periods. Major coal deposits occur in the Carboniferous strata of Europe,
North America and China, and in the Permian of Australia, India, South America
and the other landmasses that made up the former continent of Gondwana (Ward,

Table 1: Some coal quality parameters


for different markets (after Standards Australia, 1993)

Preferred Acceptable
Property Market Limits Limits Discussion
Ash Fuel 10 to 22 Max 35 High ash reduces specific energy,
% air dry increases CO2 output per GWh
and increases ash disposal
requirement
Coke 6 to 10 Max 10 Low ash reduces limestone
consumption and slag production
per tonne of steel
Cement 12 to 17 Max 30 High ash reduces nett specific
energy and increases CO2
production per tonne of cement
clinker
Total Fuel Max 1.0 Max 2.0 Dependent on pollution control
Sulphur requirements where the coal is to
be burned
% air dry
Coke Max 1.0 Max 1.2 The sulphur should be as low as
possible
Cement 0.7 to 3.5 Max 4.0 Sulphur in clinker is normally
required to be less than 1.3%
Nitrogen Fuel Max 1.8 Max 2.0 Low figures preferred by some
% daf users
Chlorine Fuel Max 0.1 Max 0.3 Should be low to reduce tendency
% air dry for fouling
Cement Low Max 0.1 Chlorine in clinker is expected to
be less than 0.03%
Ash Fusion Fuel High Min Dry bottom furnaces. Actual
(Deformation) (PF) 1100°C figures depend on power station
design
Note: daf = dry, ash-free basis

52
Chapter 3: Geological Sources of Metals in Coal and Coal Products

1984; Thomas, 2002). Coal of Mesozoic age also occurs in a number of areas, no-
tably the Jurassic of Australia and the Cretaceous of North America. There are also
significant resources of Tertiary age in different continents, including Europe, North
America, Asia, Australia and New Zealand.
Estimates of coal resources (coal that is or may in the future become economi-
cally viable), and also of the reserves of coal likely to be extracted for use under
different conditions, are mostly based on different limiting criteria in different coun-
tries. Table 2, which summarises these data, may therefore provide a better estimate
of the extent of the recoverable coal resources in the well-explored countries and
an underestimate of the amount in less intensively-explored regions. It can be seen
from the table that by far the greatest proportion of the world's recoverable coal is
located in three regions: the United States of America, the former Soviet Union and
the People's Republic of China. Australia, Poland, the United Kingdom, India, South
Africa and Germany, however, also have substantial economically recoverable or
potentially recoverable coal resources.

NEW ZEALAND COAL INDUSTRY

New Zealand is estimated to have up to 10 billion tonnes of recoverable coal,


including a large proportion of lignite deposits (Anon, 2003). The coal deposits of
New Zealand occur mainly in three regions: (1) the Late Cretaceous to Early Tertiary
Rewanui Coal Measures (bituminous) and the Late Eocene Brunner Coal Measures
(bituminous) in the West Coast region of the South Island; (2) the Eocene Waikato
Coal Measures (sub-bitiminous) in the North Island; and (3) the Late Oligocene to
Early Miocene Gore Lignite Measures in the Southland and Otago regions. Most
coal production in New Zealand comes from these three regions.
Cumulative coal production up to the end of 2003 has amounted to approxi-
mately 271 million tonnes. Production of coal in New Zealand increased sharply
from 3.3 million tonnes in 1998 to 3.9 million tonnes in 2001 (Anon, 2002), 4.4
million tonnes in 2002 (Anon, 2003), and 5.17 million tonnes in 2003, with most of
the increase being derived from the West Coast region. Of the 5.17 million tonnes
produced in 2003, 45% was bituminous coal from the West Coast region and 49%
was sub-bituminous coal, mainly from the Waikato region. The remaining 6% was
lignite from the Southland and Otago regions.

MINERAL MATTER IN COAL

As defined by Gary et al. (1972), mineral matter refers to the inorganic material
in coal. Mineral matter is more specifically defined by Standards Australia (1995)
as representing the sum of the minerals and inorganic matter in and associated with
coal. Under this definition, the material classed as mineral matter in coal embraces
three quite different types of constituents (Ward, 2002):
• Dissolved salts and other inorganic substances in the coal’s pore water;
• Inorganic elements incorporated into the organic compounds of the coal
macerals by absorption, adsorption or organo-complexation;
• Discrete inorganic particles (crystalline or non-crystalline) representing true
mineral components.

53
METAL CONTAMINANTS I N N E W Z E A L A N D

Table 2: World Coal Resources, 1996 (World Energy Council and BP-Amoco data).

Proved Recoverable Proved in Place Estimated


Reserves (a) Resources (b) Additional Resources (c)
Bitum. + Sub-bitum. Bitum. + Sub-bitum. Bitum. + Sub-bitum.
Country
Anthracite + Lignite Anthracite + Lignite Anthracite + Lignite
or Region
Mt Mt Mt Mt Mt Mt
Botswana 4,313 7,189 205,253
South Africa 55,333 121,218 5,000
Canada 4,509 4,114 6,435 12,905 26,045 31,990
USA 111,338 135,305 239,675 240,538 456,153 665,887
Brazil 11,950 17,072 15,319
Colombia 6,368 381
China 62,200 52,300
India 72,733 2,000
Indonesia 770 4,450 1,331 6,299
Japan 785 8,277
Kazakhstan 31,000 3,000
Pakistan 2,928
Thailand 2,000 2,315 3,000
Turkey 449 626 7,339 110
Uzbekistan 1,000 3,000
Czech 2,613 3,564 6,401 2,547 4,928 5,284
Repulic
Germany 24,000 43,000 44,000 78,000 186,000
Greece 2,874
Hungary 596 3,865 1,407 8,306 702
Poland 12,113 2,196 60,185 14,184
Russian Fed. 49,088 107,922 75,753 124,823 1,582,479 2,358,085
Serbia, 64 16,408
Montenegro
Ukraine 16,388 17,968 21,850 23,958 5,406 5,819
United 1,000 500
Kingdom
Australia 47,300 43,100 65,900 48,000 125,000 215,000
NewZealand 29 542
Other 5,502 10,727
Total World 509,491 474,720 Mt = million tonnes

a) Carefully measured and assessed as recoverable under present and expected economic condi-
tions and with existing technology; b) Proved amount in place available to be recovered (extracted
from the Earth in raw form) under present and expected conditions; c) Indicated and inferred
coal additional to the proved amount in place, including extensions to known deposits and as yet
undiscovered resources. Note: Not all countries report categories under notes (b) and (c), and not
all use the same limiting criteria.

54
Chapter 3: Geological Sources of Metals in Coal and Coal Products

The first two forms of mineral matter may also be described as non-mineral
inorganics. Such constituents are usually prominent in the mineral matter of lower-
rank coals, such as brown coals, lignites, and sub-bituminous materials (e.g., Given
and Spackman, 1978; Benson and Holm, 1985; Miller and Given, 1986; Ward,
1992), and contribute significantly to ash formation from combustion of lower-rank
coal deposits.
Expulsion of moisture and the associated materials in solution, together with
changes in the chemical structure of the organic matter, combine to remove most
of the non-mineral inorganics from coal with rank advance. Discrete inorganic or
mineral particles may occur in both lower-rank and higher-rank coals, and in the
absence of non-mineral inorganics commonly represent the dominant component of
the mineral matter in higher-rank coal seams (e.g., Rao and Gluskoter, 1973; Ward,
1978; Spears, 1987).
Coals extracted as mine products typically contain additional mineral constitu-
ents derived from bands and other concentrations of non-coal material within the
seam. They may also possibly contain some fragments of non-coal rock derived from
contamination of the mined product by roof or floor strata. Some of this mineral
matter (referred to as extraneous mineral matter) may be removed by cleaning proc-
esses in coal preparation plants. There is, nevertheless, usually a significant level of
mineral matter intimately associated with the macerals, commonly referred to as
inherent mineral matter, that forms an unavoidable part of even the cleanest coal
product, and must be taken into account along with the macerals in assessing the
coal’s behaviour during handling, storage, and use.

METHODS FOR MINERAL MATTER ANALYSIS

The ash produced from burning a coal represents the non-combustible residue
of the coal’s mineral matter. Many of the minerals occurring in coal undergo major
chemical changes at high temperature, involving destructive processes such as loss of
CO2 from carbonates and loss of structural water from clay minerals, as well as loss
of sulphur as SO2 from sulphide minerals such as pyrite (Rees, 1966; Raask, 1985;
Vassilev et al., 1995; Reifenstein et al., 1999). The non-mineral inorganics in the
coal macerals, left behind when the organic matter is burned, may react with some
of the liberated components to form mineral artefacts and retain otherwise volatile
elements (e.g., S fixed as sulphates) in the resultant ash residue.
As a result of losses of inorganic material during high temperature reactions, the
percentage of ash produced by coal combustion, determined as part of routine coal
analysis, is usually less than the percentage of mineral matter contained in the origi-
nal coal sample. The chemical composition of the ash, also commonly determined
in coal analysis programs, may in addition be somewhat different from the chemical
composition of the original mineral matter.
Because of the need to avoid the breakdown of mineral structures associated
with burning the coal, the percentage of mineral matter in a coal is more diffi-
cult to determine than the ash percentage. One of the most widely used methods
is based on destroying the organic matter at low temperature (around 120ºC)
by exposing the coal, under vacuum, to a reactive oxygen plasma produced by a
radio-frequency electromagnetic field (Gluskoter, 1965; Frazer and Belcher, 1973;

55
METAL CONTAMINANTS I N N E W Z E A L A N D

Standards Australia, 2000). The residue remaining after oxidation of the organic matter
consists of the essentially unaltered mineral components in the original coal, together in
some cases with additional mineral artefacts produced by reactions involving the non-
mineral inorganics.
Other methods of removing the organic matter without disrupting the mineral
structures include heating the coal to temperatures lower than those of actual com-
bustion, such as exposing the coal to air in a furnace at around 370ºC (e.g., Brown et
al., 1959; Ward et al., 2001), or treating the coal with a hot, concentrated solution of
hydrogen peroxide to oxidise the organic components (Nawalk and Friedel, 1972;
Ward, 1974). However, these methods are less effective than oxygen-plasma ashing,
and may also irreversibly break down some of the mineral components. Calculations
based on the ash percentage and ash composition, combined with other chemical
data, can also be used to provide an estimate of the mineral matter content in the
parent coal (e.g., Parr, 1928; King et al., 1936; Rees, 1966).

MINERALS FOUND IN COAL SEAMS

The minerals and mineral artefacts in the residues from low-temperature oxida-
tion may be identified from their crystal structures using X-ray diffraction methods
(e.g., Russell and Rimmer, 1978; Jenkins and Snyder, 1996). Minerals occurring in
coal samples may also be identified using optical or electron microscope techniques
(e.g., Kemezys and Taylor, 1964; Finkelman and Stanton, 1978; Stanton and Finkel-
man, 1979). Optical microscopy is typically applied to polished sections of coal, and
can also be used to identify the different macerals as well as the mineral components
(Taylor et al., 1998). Mineral identification using the electron microscope may be
aided by associated microchemical analysis, using X-ray fluorescence techniques. A
wide range of other methods may also be used to identify the minerals present in coal
samples, including thermal analysis techniques, Fourier-transform infra-red (FTIR)
spectrometry, and computation of mineral percentages from ash or whole-coal
chemical analysis data (Ward, 2002, Huggins, 2002; Vassilev and Tascon, 2002).
Some of the mineral matter in coal seams may occur as bands, lenticles, fracture
fillings, plant impregnations, mineral-rich nodules and other masses visible in hand
specimen within the organic matter. Inorganic elements dissolved in the pore water
may also be precipitated when the water evaporates from the coal in exposed out-
crops, drill cores or mine faces.
A large part of the crystalline mineral matter in higher-rank coals, however, oc-
curs in masses that can only be seen using optical or electron microscope techniques.
Such microscopically visible mineral matter includes fine layers and other concentra-
tions intimately intergrown with the maceral components, as well as discrete mineral
fragments or crystals and a range of nodules, lenticles, veins, pore infillings and
cell replacement structures (Fig. 2). The modes of mineral occurrence revealed by
microscopy often indicate how the mineral material may have formed, or how it
might respond to coal preparation and utilisation processes.
A list of the minerals that may be present in coal samples or low-temperature
ashing or oxidation residues is given in Table 3. The most abundant of these are
usually the clay minerals, although quartz, pyrite, siderite, calcite and dolomite or
ankerite, together in some cases with phosphate minerals such as apatite, may also

56
Chapter 3: Geological Sources of Metals in Coal and Coal Products

FIGURE 2: Photomicrographs of polished or thin sections showing modes of mineral occurrence


in coal: a) clay minerals of detrital origin (dark) interbedded with maceral components (field width
1.4 mm); b) calcium carbonate shell fragments (field width 1.5 mm); c) crystalline siderite nodule
enveloped by coal macerals (field width 1 mm); d) pyrite (light) infilling plant cells (field width
0.2 mm); e) apatite infilling plant cells (electron microscope image); f) calcite veins cross-cutting
vitrinite band (field width 1 mm).

be found as significant components of the mineral matter in many coal seams (e.g.,
O’Gorman and Walker, 1971; Rao and Gluskoter, 1973; Ward, 1978, 1989, 2002;
Vorres, 1986; Ward and Swaine, 1995; Ward et al., 2001; Vassilev and Tascon,
2002). Iron-bearing sulphate minerals, such as jarosite and coquimbite, may be
formed by oxidation of pyrite with exposure of the coal to the atmosphere; a proc-
ess that also liberates sulphuric acid in any associated runoff water. Many other
sulphates in oxidation residues, such as bassanite, glauberite and hexahydrite, usu-
ally represent mineral artefacts formed by interaction of non-mineral inorganics in
lower-rank coals during destruction of the organic matter.

57
METAL CONTAMINANTS I N N E W Z E A L A N D

Table 3: Principal minerals found in coal and oxidation residues (Ward, 2002)

SILICATES CARBONATES
Quartz SiO2 Calcite CaCO3
Chalcedony SiO2 Aragonite CaCO3
Clay minerals: Dolomite CaMg(CO3)2
Kaolinite Al2Si2O5(OH)4 Ankerite (Fe,Ca,Mg)CO3
Illite K1.5Al4(Si6.5Al1.5)O20(OH)4 Siderite FeCO3
Smectite Na0.33(Al1.67Mg0.33)Si4O10(OH)2 Dawsonite NaAlCO3(OH)2
Chlorite (MgFeAl)6(AlSi)4O10(OH)8 Strontianite SrCO3
Interstratified Witherite BaCO3
clay minerals Alstonite BaCa(CO3)2
Feldspar KAlSi3O8
NaAlSi3O8 SULPHATES
CaAl2Si2O8 Gypsum CaSO4.2H2O
Tourmaline Na(MgFeMn)3Al6B3Si6O27(OH)4 Bassanite CaSO4.½H2O
Analcime NaAlSi2O6.H2O Anhydrite CaSO4
Clinoptilolite (NaK)6(SiAl)36O72.20H2O Barite BaSO4
Heulandite CaAl2Si7O18.6H2O Coquimbite Fe2(SO4)3.9H2O
Rozenite FeSO4.4H2O
SULPHIDES Szomolnokite FeSO4.H2O
Pyrite FeS2 Natrojarosite NaFe3(SO4)2(OH)6
Marcasite FeS2 Thenardite Na2SO4
Pyrrhotite Fe(1-x)S Glauberite Na2Ca(SO4)2
Sphalerite ZnS Hexahydrite MgSO4.6H2O
Galena PbS Tschermigite NH4Al(SO4)2.12H2O
Stibnite SbS
Millerite NiS OTHERS
Anatase TiO2
PHOSPHATES Rutile TiO2
Apatite Ca5F(PO4)3 Boehmite Al.O.OH
Crandallite CaAl3(PO4)2(OH)5.H2O Goethite Fe(OH)3
Gorceixite BaAl3(PO4)2(OH)5.H2O Crocoite PbCrO4
Goyazite SrAl3(PO4)2(OH)5.H2O Chromite (Fe,Mg)Cr2O4
Monazite (Ce,La,Th,Nd)PO4 Clausthalite PbSe
Xenotime (Y,Er)PO4 Zircon ZrSiO4

FORMATION OF MINERAL MATTER IN COAL

As indicated in a recent review by Ward (2002), the minerals occurring in coal


may be formed by a range of different processes, including input of sedimentary par-
ticles (mineral fragments) into the original peat-forming environment, accumulation
of shells and other biogenic components within the peat deposit, and precipitation of
material from solution in the peat swamp or in the pores of the peat bed. Minerals
may also be formed by precipitation in pore spaces, joints and other fractures in the
coal by post-depositional processes. Non-mineral inorganics, which are also part of
the mineral matter, may be concentrated in the pore-filling water in different parts of
low-rank coal seams by post-depositional ion migration effects.

58
Chapter 3: Geological Sources of Metals in Coal and Coal Products

Detrital minerals

Some of the mineral matter in coal represents material washed or blown as de-
trital fragments into the accumulating peat deposit. Such material includes non-coal
sediment introduced from river water, airborne dust etc., as well as airborne debris
introduced to the peat by contemporaneous volcanic activity. Mineral matter formed
in this way is typically represented by small fragments of quartz and sometimes feld-
spar, along with fine, often irregular bands made up mainly of clay minerals, inti-
mately mixed with the organic matter (e.g., Davis et al., 1984; Ruppert et al., 1991).
It may also include thin but extensive bands of altered volcanic ash that are found
in some coal seams and referred to as tonstein deposits (e.g., Bohor and Triplehorn,
1993). Residual concentrations of mineral matter may also be developed with degra-
dation of the peat bed (e.g., exposure at low water levels) and removal of the organic
matter by oxidation, including the effect of fires on the accumulating plant debris.

Biogenic minerals

Many of the minerals found in coal may have resulted directly from biological
activity in the original peat swamp. The siliceous skeletons (frustules) of diatoms,
and also possibly siliceous sponge spicules, are abundant in a number of modern-day
peat deposits (Raymond and Andrejko, 1983). These particles are relatively soluble
in water, and may be corroded and partly dissolved in older peat accumulations.
Shells made up of calcium carbonate may also be present within or closely associ-
ated with coal beds (e.g., Ward, 1991; Kortenski, 1992). Shell-rich bands occurring
in some seams may represent drowning of the original swamp environment, with
replacement of the floral ecosystem by a deeper-water faunal accumulation.
Many of the plants forming modern-day peats contain accumulations of inor-
ganic silica, and possibly also a range of other elements, as phytoliths within the
vascular structure (e.g., Raymond and Andrejko, 1983). These mineral accumula-
tions may be preserved within the peat, or they may be released in solution with
plant decay, and either lost from the depositional system or reprecipitated in other
parts of the peat deposit.

Authigenic mineral precipitates

Minerals in coal may also be formed by crystallisation in place (a process re-


ferred to as authigenesis), either within the peat deposit at the time of its formation
(syngenetic mineralisation) or in the pores and fractures of the coal after compaction
and rank advance (epigenetic mineralisation). Syngenetic precipitates include siderite
nodules, microcrystalline aggregates of pyrite crystals referred to as framboids, and
a range of cell and pore infillings (typically consisting of kaolinite, quartz, phosphate
minerals, and pyrite). Epigenetic mineralisation, commonly occurring as joint or
fracture infillings, can include calcite, dolomite, ankerite and siderite, as well as
pyrite, marcasite, apatite, dawsonite, illite, and chlorite.
In the absence of significant detrital input, and where low proportions of pyrite
or carbonate minerals (see below) are present, authigenic quartz and clay minerals,
particularly well-ordered kaolinite, may represent the dominant form of mineral mat-
ter in the coal seam. These minerals commonly occur as petrifactions infilling the cell

59
METAL CONTAMINANTS I N N E W Z E A L A N D

cavities of plant fragments (e.g., Kemezys and Taylor, 1964; Lindqvist and Isaac, 1991;
Sykes and Lindqvist, 1993), and in the spaces between the individual maceral com-
ponents. Despite being dominated by silicates in both instances, the mineral matter in
such seams is usually quite different to the sediment that forms the non-coal rocks of
the immediately overlying and underlying (roof and floor) strata (Ward, 1989).
A range of phosphate minerals can also occur in coals, including apatite as well
as aluminophosphate minerals of the crandallite group. Although skeletal fragments
and possibly coprolite particles rich in phosphate may also be present, the phosphate
minerals in many coals occur as cell and pore infillings (Cook, 1962; Ward et al.,
1996; Rao and Walsh, 1999) and thus also represent syngenetic precipitates in the
original peat bed.
Pyrite is a common mineral in many coal seams, especially those associated with
sedimentary strata of marine origin. Much of the pyrite is intimately associated with
the organic matter, and was clearly formed during or shortly after peat accumula-
tion. Such early-formed pyrite can occur as small nodules or framboids (spherical
polycrystalline aggregates 1 to 100 µm across), as crystals infilling or replacing coal
macerals, and as more massive accumulations (e.g., Querol et al., 1989; Hower
and Pollock, 1989; Kortenski and Kostova, 1996). Early-formed pyrite in coal is
generally thought to have been precipitated by interaction of dissolved iron and H2S
within the peat bed (Williams and Keith, 1963). The H2S was derived from bacterial
reduction of sulphate ions in the reducing environment of the peat deposit, with the
sulphate introduced from the water, especially seawater, infiltrating into the swamp
or permeating through the peat bed.
Syngenetic accumulations of siderite may also be found in coal, typically as
nodules with a radiating crystal structure or as infillings and replacements of the
maceral components (Kortenski, 1992; Zodrow and Cleal, 1999). An abundance of
such syngenetic siderite is usually thought to indicate deposition of the coal mainly
under non-marine conditions, or at least under the influence of swamp or formation
waters with a low sulphate content (Ward, 1989). Larger masses of calcite and other
syngenetic minerals up to 300 mm or so in diameter, referred to as coal balls (e.g.,
Scott et al; 1996; Greb et al., 1999), may also be found in some coal seams.
In addition to syngenetic precipitates, sulphide, carbonate and clay minerals
may occur as epigenetic cleat and fracture infillings in coal seams (Cobb, 1985;
Kolker and Chou, 1994; Tarriba et al., 1995; Faraj et al., 1996; Hower et al., 2001).
Some of these occurrences may represent remobilisation of elements or minerals
within the coal, but others may be the result of external factors such as nearby
igneous intrusions, or produced by post-depositional fluid movement through the
coal-bearing succession.

Non-mineral inorganics

Particularly in lower-rank coals (sub-bituminous coals and lignites), a significant


part of the mineral matter may be represented by inorganic elements associated with
the organic constituents. Such occurrences include inorganic components dissolved
in the pore water of the coal, elements held in exchangeable relationships with
particular organic compounds (e.g., carboxylates), and inorganic elements forming
chelates and other organometallic complexes within the organic matter.

60
Chapter 3: Geological Sources of Metals in Coal and Coal Products

The elements occurring in different ways within the non-mineral inorganic frac-
tion may be removed from the coal by a series of selective leaching processes, and
determined by analysis of the different leachate fractions (Miller and Given, 1978;
1986; Benson and Holm, 1985; Ward, 1992). Soaking in water, for example, might
be used to remove constituents dissolved or potentially soluble in the pore water
(Fig. 3), soaking in ammonium acetate to remove exchangeable ions from compo-
nents such as carboxylates, and treatment with hydrochloric acid to remove any
acid-soluble organometallic complexes from the organic matter. The results may be
evaluated by calculating, as a percentage, the concentration of each element in the
respective leachates, in relation to the total concentration of the same element in the
coal, as determined from ash analysis of the untreated coal sample.
An example of the results obtained from such a selective leaching process is
given in Figure 4. A large proportion of the sodium in this particular instance, as
well as a significant proportion of the sulphur, appears to be dissolved or potentially
soluble in the pore water. Much of the Ca, Mg and Mn, together with all but a small
proportion of the remaining Na, appears to be held in an exchangeable relationship,
and removed by acetate treatment. Large proportions of the Fe and Al in the coal,
as well as much of the remaining Ca, Mg and Mn, appear to be in acid-soluble
organometallic components.
Examination of Figure 4 suggests that the proportion of some elements, particu-
larly sodium, released from the coal by the combination of leaching processes may
be greater that 100%. This is because some of the Na in the coal is volatile, and is
lost from the coal during combustion. The proportion of Na in the original coal
sample indicated by analysis of the coal’s ash in such cases would be less than the
total proportion of Na in the original raw coal that was subjected to the different
leaching processes. The percentages of each element leached by the different proc-
esses summarised in Figure 4 are based on the percentage of that element indicated
by ash analysis, recalculated as a fraction of the original coal by consideration also
of the ash percentage. Since more Na was actually present in the original coal than
the amount retained in the ash, the total percentage of Na leached from the coal by
the combination of treatments may in such cases exceed the proportion of Na in the
coal indicated by back-calculation from the ash analysis data.

FIGURE 3: Flowsheet for selective leaching of low-rank coals to evaluate non-mineral inorganic
components (after Ward, 1991).

61
METAL CONTAMINANTS I N N E W Z E A L A N D

FIGURE 4: Proportions of inorganic elements in a South Australian sub-bituminous coal removed


by water washing, ammonium acetate and hydrochloric acid during selective leaching experiments.
The values shown represent the total amount of each element leached from each coal sample as a
proportion of the total amount of that element in the same sample determined by raw (untreated)
coal ash analysis (after Ward, 2002).

The distribution of these non-mineral inorganics within the seam may be con-
trolled by post-depositional factors unrelated to the environments that formed the
different layers of the coal bed. Studies of the brown coals in the Latrobe Valley of
Australia, for example (Brockway and Borsaru, 1985), suggest that Na, Mg, Ca, Al,
Fe and Cl ions migrate with upward movement of water through the thick, porous,
coal seams, apparently as a result of a combination of hydrostatic pressure from the
underlying aquifers and loss by surface evaporation, and that the coal seam acts in
part as an ion exchange medium as groundwater moves through it. Ions that are
held more strongly to the organic matter appear to move more slowly, giving rise to
different concentrations of ions in solution or incorporated in the organic matter at
different levels in the coal bed.

METALS AND OTHER TRACE ELEMENTS IN COAL


Almost every element in the periodic table has been identified in coal (Swaine,
1990; Finkelman, 1994a), with the extent of knowledge increasing as more sensitive
analytical methods are developed. Most of these elements have also been found in
New Zealand coals at levels comparable to those reported for international coals
(Gray, 1986; Soong et al., 1984; Soong and Berrow, 1979; Sim, 1977; Sim and Lewin,
1975; Lynskey et al., 1984; Purchase, 1985; Gainsford, 1985; Newman, 1988; Li,
2002; Chapter 4). Elements in coal include the relatively common elements such as
the silicon, aluminium, iron, calcium and magnesium that make up the bulk of the
mineral matter discussed above, as well as a wide range of other elements that can-
not readily be related to any particular mineral component. Most of these elements
occur at comparably lower concentrations than the same elements in other rock and
soil materials, the only exceptions being selenium, boron, arsenic, antimony, which
are enriched in coal (with respect to crustal abundances) by factors of 1.5 to 16X
(Riley et al., 2002). Some of these elements may nevertheless give rise to potentially
toxic impacts in different ways when the coal is mined, processed or used.
The concentration of most trace elements in coal is determined from analysis
of the coal’s ash, using techniques such as ICP-MS, ICP-AES, NAA, XRF and AAS

62
Chapter 3: Geological Sources of Metals in Coal and Coal Products

(Davidson and Clarke, 1996; Huggins, 2002; Vassilev and Tascon, 2002), and recal-
culated as a fraction of the original coal sample. Any elements that are volatile and
escape during the combustion process will, however, be omitted in this type of ana-
lytical study. Methods that analyse the whole coal are in some respects preferable,
but detection limits may be lower and other difficulties may arise (e.g., difficulties in
getting elements into solution) as a result of dilution by the organic matter. Indica-
tive trace element concentrations in coals from several different countries are given
in Table 4. Similar data for New Zealand and US coals, together with enrichment/
depletion factors in relation to the averaged concentrations of the same elements in
the Earth’s crust, are given in Table 5 (see also Chapter 4).

Organic and inorganic affinity

Although the total concentration of particular trace elements can generally be


determined with a high degree of precision, an indication of the way in which the
various elements occur in the coal is typically based on little more than speculation.
This is because much of the mineral matter occurs either as fine grained, dispersed
material or as intimate aggregates of micron-sized grains, making it impossible to
obtain pure fractions for the application of conventional bulk trace element analyti-
cal techniques. Considerable effort has therefore been devoted to the investigation of
indirect methods of determining the different modes of trace element occurrence.
One such method is to analyse a series of finely powdered coal samples that have
been separated by float-sink processes based on immersion in liquids of different
densities. Elements that are mainly associated with the organic matter tend to have
higher concentrations in the light (low-density) or organic-rich fractions separated
in this way, whereas elements more strongly associated with the mineral matter typi-
cally have higher concentrations in the denser, mineral-rich fractions (Gluskoter et
al., 1977). Figure 5, based on studies of this type, indicates the degree of organic
or mineral association (organic or mineral affinity) for a number of elements in a
selection of Australian coal samples. The right-hand graph in each pair shows the
percentage of the element in the powdered coal occurring in each of the different
density fractions, and the left-hand graph shows the cumulative trend in element
concentration with increasing particle density. Most of the elements shown have an
affinity with the mineral matter, expressed by increasing concentrations in the denser
coal fractions; some, however, have an organic affinity, with similar if not higher
proportions occurring in the lower-density fractions.
As well as the ash percentage and the major inorganic elements such as Si, Al, Fe
and Ca, a large number of other elements, including Pb, Zn, Cu, Mo, Cr, Ba, Sr, Ti,
Co and As, typically display a mineral affinity in such evaluations (e.g., Gluskoter
et al., 1977). The concentration of these elements may reflect the presence of small
amounts of minerals in which the element is a dominant component, such as small
amounts of sphalerite (ZnS) giving rise to significant zinc concentrations (e.g., Hatch
et al., 1976), or larger amounts of minerals in which the element is a minor compo-
nent (e.g., rubidium occurring, along with potassium, in micas, feldspars and illitic
clay minerals; Ward et al., 1999). Combinations of different modes of occurrence
may also be developed in some coal samples.
Some elements, such as boron, gallium and in some coals germanium, appear to

63
METAL CONTAMINANTS I N N E W Z E A L A N D

Table 4: Trace element concentrations in some world coals (after Swaine, 1990;
Finkelman, 1994; Ren et al, 1999; Ward, 1984; Dale, 2003). Regional
USA data updated from Gluskoter et al. (1977), and Australian regional
data from Ward (1984); country data updated from Swaine (1990).

USA Australia
Element Illinois Western USA NSW & Australia EC/UK World
Appalachians
Basin Coals Qld
Lithium 20
Beryllium 1.7 1.3 0.46 2.0 0.87 1.5 3
Boron 110 42 56 60 19 30 80
Fluorine 67 89 62 114 150
Phosphorus 64 150 130 310
Chlorine 348 150
Scandium 2.7 5.1 1.8 3 4 5.2
Titanium 800 900 700 400
Vanadium 32 38 14 30 28 20 48 19
Chromium 18 20 9.0 20 9.6 6 26.2 8
Manganese 53 18 49 50 127 130 131.9 70
Cobalt 7.3 9.8 1.8 8 4.5 4 8.4 4
Nickel 21 15 5.0 20 5.8 15 41 13
Copper 14 18 10 17 15 15 14.6 7
Zinc 250 25 7.0 40 18 25 80.3 20
Gallium 3.2 5.7 2.5 6 4 4 6 2.3
Germanium 6.9 1.6 0.91 6 6 7
Arsenic 14 25 2.3 15 1.0 1.5 11.9 29
Selenium 2.2 4.0 1.4 2.7 0.53 0.8 2.4
Strontium 35 130 260 100 100 240
Zirconium 47 45 33 25 100 100 16 28
Molybdenum 8.1 4.6 2.1 3 0.85 1.5 2.8 4
Silver 0.03 0.02 0.03 <0.5
Cadmium 2.2 0.24 0.18 0.5 0.11 0.08 0.27
Tin 3.8 2.0 1.9 <3 2 1.3
Antimony 1.0 0.39 0.5 3.3
Barium 100 200 500 300 <100 70 120
Lanthanum 6.8 15 5.2 10
Tungsten 0.82 0.69 0.75 <10
Mercury 0.2 0.20 0.09 0.15 0.045 0.10 0.29
Lead 32 5.9 3.4 30 3.2 10 22 59
Thorium 2.1 4.5 2.3 4 2.8 2.7 3.9
Uranium 1.5 1.5 1.2 1.1 2 2.13

64
Table 5. Average trace element concentrations in Greymouth and US coals, in comparison to
concentrations of the same elements in the Earth’s crust. Values in ppm (except for Hg in ppb) on a coal basis.
Enrichment factors (EF) for the elements relative to the average composition of the crust are also shown.

Sample As Sb Hg* Se U Pb Cr Co Ni Mn S Cl Br Sc Cs Rb Sr
Greymouth coals1 1.5 0.16 21 0.77 0.3 5 7 30 24 27 0.35 450 2.2 1 2.6 8 45
US coals2 24 1.2 170 2.8 2.1 11 15 6 14 43 1.80 600 17 4.2 1.1 21 130
Earth's crust3 1.8 0.2 85 0.1 1.8 14 100 25 75 950 0.04 145 2.5 22 3 90 375
EF4Grey/crust 0.8 0.8 0.2 15.3 0.2 0.4 0.1 1.2 0.3 0.0 10.0 3.1 0.9 0.0 0.9 0.1 0.1

65
Sample Ba Zn La Ce Nd Sm Eu Gd Tb Tm Yb Lu Zr Hf Ta Th W
1
Greymouth coals 47 8 4 8 3.1 0.6 0.10 0.38 0.08 0.03 0.25 0.04 12 0.4 0.2 1.1 18.0
US coals2 170 53 12 21 9.5 1.7 0.4 6 0.3 0.2 1 0.1 27 0.7 0.2 3.2 1
Earth's crust3 425 70 30 60 28 6 1.2 6.2 0.9 0.5 3.4 0.5 165 3 2 7.2 1.5
EF4Grey/crust 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 12.0

Notes:
1: Mean concentrations of trace elements in 53 coal samples from Greymouth bituminous coals (Li, 2002)
Chapter 3: Geological Sources of Metals in Coal and Coal Products

2: After Bragg et al., 1997, U.S. Geological Survey's National Coal Resources Data System (NCRDS)
3: After Wedepohl, 1995
4: EFGrey/crust = Enr�
METAL CONTAMINANTS I N N E W Z E A L A N D

have a stronger affinity with the organic matter (e.g., Finkelman, 1994a; Ward et al.,
1999). These may represent elements incorporated into the organic compounds in
some way (e.g., as non-mineral inorganics), or they may possibly represent elements
that occur only in very fine mineral particles dispersed within otherwise ‘clean’
maceral components. Some elements may, of course, be partly associated with the
organic matter and partly with the minerals in the coal concerned.
Boron is of particular interest, partly because of its potential environmental
impact and partly because it is abundant in coals formed in a depositional setting
influenced by marine conditions (Swaine, 1990). Boron commonly shows a greater
affinity with the organic matter of coal (as in Fig. 5), possibly as a result of absorp-
tion of boron into the original plant tissues with immersion in seawater. Modern-day
seawater, for example, typically has a B concentration of around 5 ppm, whereas
river and lake waters typically have boron concentrations of around 0.015 ppm
(Bowen, 1979). However, boron in coal may also occur in clay minerals such as illite
(Bohor and Gluskoter, 1973), and possibly as discrete grains of tourmaline within
the mineral fraction.
Sixteen elements occurring in coal have been included in a list of potentially
hazardous air pollutants (HAPs) under the US Clean Air Act (Demir et al., 1997),
namely As, Be, Cd, Cl, Cr, Co, F, Hg, Mn, Ni, P, Pb, Sb, Se, Th and U. These may oc-
cur in a variety of ways (Davidson and Clarke, 1996; Kolker and Finkelman, 1998),
including organic and inorganic associations. Some may be reduced with removal of
mineral-rich material by coal cleaning processes in preparation plants (Davidson and
Clarke, 1996; Demir et al., 1997), but the effectiveness of such techniques depends
at least in part on the particle size of the host minerals and the extent to which those
minerals become separated (or liberated) from the organic matter with crushing of
the plant feed. Elements with an organic affinity are clearly more difficult to remove
in this way, since it is the organic fraction of the coal that is the desired product for
marketing and use. Removal of elements from the coal by such processes also has the
effect of concentrating them in the refuse or waste fraction, from which they may be
released into the environment by a different route when these wastes themselves are
emplaced in disposal systems.

Association of trace elements with particular mineral components

Knowledge of trace element occurrence in minerals and rocks generally, obtained


through use of techniques such as electron microprobe analysis, provides a basis
from which the mineral associations can be assessed of particular trace elements in
coal samples (e.g., Finkelman, 1982, 1994a). A more direct evaluation of mineral
associations for particular elements may also be obtained from correlations of el-
ement abundance with the abundance of particular mineral components (Fig. 6),
revealed through processes such as quantitative X-ray diffraction techniques (Ward
et al., 1999).
Elements such as As, Cd, Se, Tl, Hg, Pb, Sb and Zn, for example, sometimes
referred to as chalcophile elements, are generally thought to be associated with sul-
phide minerals such as pyrite, either as solid-solution constituents or as discrete
sulphide phases. Ward et al. (1999), for example, suggest that overall concentra-
tions of around 0.1% As are present in the pyrite in some Australian coals; electron

66
67
Chapter 3: Geological Sources of Metals in Coal and Coal Products

FIGURE 5: Concentration of selected trace elements in float-sink fractions of some Australian coal samples, showing organic (boron, germanium BALG)
and inorganic affinities (after Ward 1980; Ward and Swaine, 1995). Coal samples: ABXT = Greta seam, Aberdare, NSW; BALG = Balgownie seam,
Wollongong, NSW; DUDL = Dudley seam, Newcastle, NSW; GNTN = Great Northern seam, Newcastle, NSW; HBRN = Greta seam, Hebburn, NSW;
NP10 = Nipan 10 seam, Theodore, Qld.
METAL CONTAMINANTS I N N E W Z E A L A N D

FIGURE 6: Relation between concentrations of selected trace elements, based on whole-coal XRF,
and abundance of particular minerals in coals, based on quantitative X-ray diffraction analysis,
Gunnedah Basin, Australia (Ward et al., 1999): a) arsenic in relation to pyrite content; b) strontium
in relation to calcite and dolomite; c) rubidium in relation to potassium-bearing minerals, illite,
illite/smectite, mica and orthoclase; d) titanium in relation to kaolinite content.

microprobe data obtained by Kolker and Finkelman (1998) and Hower et al. (1997)
respectively indicate up to 1.2% and 3.5% As at individual points in some of the
pyrite particles in American coal samples. Belkin et al. (1997) and Finkelman et al.
(2002) indicate unusually high concentrations of As (up to 35,000 ppm or 3.5% on
a whole-coal basis) in some coals from SW China, occurring in a variety of forms
including post-depositional impregnations of the organic matter.
Elements such as Rb, Ti, Cr, Zr and Hf, often referred to as being among the
lithophile elements, are more probably associated with aluminosilicates such as mi-
cas, feldspars and the clay minerals. Chromium, for example, appears to be closely
associated with illite and illite/smectite in some US coals (Glick and Davis, 1987),
and also in some New Zealand coals (Li, 2002), and in others as discrete mineral
grains derived from adjacent basement rocks (Brownfield et al., 1995). Strontium
and barium are commonly associated with carbonate minerals, either as discrete
masses of relatively unusual minerals such as strontianite and witherite (Tarriba
et al., 1995), or as minor components of more common carbonates such as calcite
and dolomite (Ward et al., 1999). However, Sr and Ba may also occur as alumino-
phosphate minerals such as goyazite, gorceixite and crandallite (Ward et al., 1996),
intimately associated with the organic matter.
Small amounts, and in some cases individual particles, of rare or unusual miner-
als may also be noted in particular coal samples, especially during studies using
electron microscope and related techniques. Individual grains of minerals such as
crocoite (PbCrO4), clausthalite (PbSe) and xenotime (YPO4), for example, have
been identified in some coal samples (Li et al., 2001; Hower et al., 1999, 2000;

68
Chapter 3: Geological Sources of Metals in Coal and Coal Products

Hower and Robertson, 2003), along with more common but still relatively rare min-
erals such as zircon, chromite and tourmaline. Minerals such as these may in part be
associated with particular trace element occurrences. An indication of the possible
mode of occurrence for some of the elements identified as hazardous air pollutants,
derived from such studies, is given in Table 6.
Other indications of mineral associations can be obtained from sequential diges-
tion of different mineral components in a series of different reagents, coupled with
analysis of the major and trace elements liberated by the reagents concerned (Palmer
et al., 1993; Laban and Atkin, 1999). These provide a means of differentiating,
for example, the proportion of individual trace elements associated with carbonate
minerals, soluble in HCl, from the proportions associated with silicates (soluble in
HF) or sulphides (soluble in HNO3).
Most of the techniques described above for investigating the mode of occurrence
of trace elements are indirect and are based upon assumptions of mineral behaviour
that may not always be correct. Microbeam techniques (such as electron microprobe
analysis) hold great promise for directly determining the mode of occurrence of trace
elements in individual grains (Davidson and Clarke, 1996; Huggins, 2002; Galbreath
and Brekke, 1994; Katrinak and Benson, 1995; Querol and Chenery, 1995; Chenery
et al., 1995). Techniques such as nuclear microprobe analysis (proton induced gamma
ray emission (PIGE) and proton induced X-ray emission (PIXE)), secondary ion mass
spectrometry (SIMS) and laser ablation–inductively coupled mass spectrometry (LA-
ICP-MS) have the capability to perform micron scale non-destructive analysis at the
ppm level in individual mineral grains (Hickmott, 1993; Hickmott and Baldridge,
1995). Another potentially useful technique is synchrotron induced X-ray emission

Table 6: Indicative mode of occurrence of some potentially


hazardous trace elements in coal (Finkelman, 1994b)

Level of
Element Common mode of occurrence
confidence*
Antimony Pyrite and accessory sulphides 4
Arsenic Pyrite and accessory sulphides 8
Beryllium Organic association 4
Cadmium Solid solution in sphalerite 8
Chromium Organic and/or clay association 2
Cobalt Pyrite; some in accessory sulphides 4
Lead Galena 8
Manganese Carbonates, especially siderite and ankerite 8
Mercury Pyrite 6
Nickel Unclear; perhaps sulphides, organics or clay minerals 2
Organic association; pyrite and accessory sulphides;
Selenium 8
selenides
* Level of confidence: a number between 1 (low) and 10 (high) expressing the consistency and pre-
dictability of the element’s indicated common mode of occurrence in coal (see Finkelman, 1994b).

69
METAL CONTAMINANTS I N N E W Z E A L A N D

(micro SRIXE), which offers ppm sensitivity at sub micron resolution (Kolker and
Chou, 1994). The application of many of these techniques to coal mineral matter
analysis is in its infancy, and routine analytical techniques are yet to be developed.

MINERALS AND TRACE ELEMENTS IN COAL MINING AND USE

Knowledge of the form in which particular metals and other trace elements oc-
cur is important in assessing the likely release of any toxic material to the environ-
ment when the coal is mined, stored or used (Fig. 7). It may also be significant in
assessing the potential to reduce the concentration of particular elements in coal
products by selective mining and/or preparation processes, or the interaction with
the environment of the trace elements retained in fly ashes, slags and other utilisation
by-products (Davidson and Clarke, 1996). Sakulpitakphon et al. (2004) provide
one example of a case study, relating the occurrence of a particularly toxic element
(mercury) in different parts of a coal seam to the emission and capture of the same
element when the coal is used in a power station.

Mineral matter behaviour at high temperatures

Although reactions of the organic matter are fundamental to different aspects


of coal utilisation, the minerals and other inorganic constituents also react when the
coal is used in different ways, such as in combustion, gasification, coke production,
and iron and steel manufacture (Nankervis and Furlong, 1980; Raask, 1985; Burch-
ill et al., 1990; Clarke, 1993; Vassilev et al., 1995; Gupta et al., 1999; Reifenstein et
al., 1999; French et al., 2001). High-temperature processes involving mineral matter
include transformations in the production of fly ash and bottom ash, fusion and
crystallization to form slag deposits, vaporization and condensation associated with
fouling of furnace systems, interaction with internal boiler components to produce
corrosion, and abrasion of metals and other exposed materials in the combustion
plant by minerals or mineral-derived particles. The resistivity of the individual ash
particles, derived from the mineral matter, may also be significant in controlling the
collection of fly ashes in electrostatic precipitator systems.
As well as alteration of individual minerals brought about by exposure to high
temperatures, such as loss of CO2 from carbonates, S from sulphides, and crys-
tal lattice OH from clay minerals, these processes may include the formation of
new minerals such as tridymite, cristobalite, mullite, lime, periclase, haematite and
magnetite, more stable under the high temperatures of the furnace environment.
Interaction of minerals with each other may also take place, forming products such
as anorthite, sanidine, gehlenite and a range of spinel minerals (Hower et al., 1996;
Reifenstein et al., 1999; French et al., 2001). Interaction of minerals with carbon
may also occur, forming products such as silicon carbide, and, especially in low-rank
coals, interaction between the minerals and the non-mineral inorganic components.
Anhydrite, for example, may be formed by the interaction of Ca in mineral or or-
ganic association with organic or pyritic sulphur released during the combustion
process. Many of the minerals or their high-temperature products fuse at still higher
temperatures to form a liquid phase, which then cools in other parts of the fur-
nace to become a non-crystalline glass in the ash or slag residues. Some inorganic

70
Chapter 3: Geological Sources of Metals in Coal and Coal Products

FIGURE 7: Possible pathways for mobility of metals and other components associated with coal
mining, preparation, transport and use (Ward, 1984, after US Office of Technology Assessment).

components may vaporize, with the vapour condensing in cooler regions to form
different types of fouling accumulations.
A list of mineral species reported in different high-temperature products associ-
ated with coal utilization is given in Table 7. Non-crystalline or amorphous material,
effectively representing inorganic glass, may also be present, and indeed such glass
commonly forms a major component of fly ash and slag deposits (Hower et al.,
1996; Winburn et al., 2000; Ward and French, 2003). While some elements may oc-
cur within individual crystals such as magnetite, or as an inherent part of the glassy
constituents, many trace elements seem to occur as components adsorbed on to the
surfaces of individual fly ash particles (including unburnt carbon as well as mineral
and glassy phases), following condensation with particle cooling in the furnace at-
mosphere. Knowledge of the mineral and glassy phases in the ash, in relation to the
mineral matter characteristics of the original coal, provides a basis for understanding
the interactions that take place among the different mineral matter components at
the high temperatures associated with coal utilisation, and also the ways in which
the different trace elements may occur in combustion residues and by-products com-
pared to the original coal material. Such studies are also useful as a basis for assess-
ing the extent to which individual elements might be released from the ash when it is
exposed to weathering associated with disposal or use (Sajwan et al., 2003).

71
METAL CONTAMINANTS I N N E W Z E A L A N D

Table 7: Principal minerals identified in high-temperature phases


associated with coal utilization (Ward, 2002)

Mineral Composition Mineral Composition


Quartz SiO2 Periclase MgO
Cristobalite SiO2 Wuestite FeO
Tridymite SiO2 Hematite Fe2O3
Metakaolin Al2O3.2SiO2 Maghemite Fe2O3
Mullite Al6Si2O13 Magnetite Fe3O4
Albite NaAlSi3O8 Spinel MgAl2O4
Anorthite CaAl2Si2O8 Magnesioferrite MgFe2O4
Sanidine KAlSi3O8 Calcium ferrite CaFe2O4
Corundum Al2O3 Srebrodolskite Ca2Fe2O5
Pyrrhotite Fe(1-x)S Brownmillerite Ca4Al2Fe2O10
Oldhamite CaS Wollastonite CaSiO3
Anhydrite CaSO4 Gehlenite Ca2Al2SiO7
Aragonite CaCO3 Merwinite Ca3Mg(SiO4)2
Vaterite CaCO3 Melilite Ca4Al12MgSi3O14
Portlandite Ca(OH)2 Whitlockite Ca3(PO4)2
Lime CaO

For the reasons outlined in the discussion of mineral matter in coal, directly
determining the mode of occurrence of trace elements in coal utilisation by-products
such fly ash is still an analytical challenge. Indirect methods such as selective leach-
ing have been widely used, but many are based on assumptions that may be incorrect
(Davidson and Clark, 1996; Davidson, 2000). Microbeam techniques have been
applied, but the application of many of these is still in its infancy and routine proce-
dures have yet to be developed. The development of routine methods based on such
techniques is essential, however, if the environmental behaviour of trace elements in
such a heterogeneous material as fly ash is to be fully understood.

CONCLUSIONS

A wide range of metals and other trace elements are found in coal, with the
elements present ant their abundance depending in large part on the rank, type and
grade of the coal concerned and the nature of the minerals and other inorganics
associated with the maceral components. Some of the minerals were introduced to
the depositional environment when the original coal-forming peat was laid down,
either as sediment grains, skeletal particles or authigenic precipitates, and some were
introduced after deposition, and in many cases also after rank advance, from solu-
tions permeating the pore spaces and fractures of the sedimentary succession.
Despite several decades of research the mode of occurrence of many trace ele-
ments in coal and coal utilisation by-products, and their relation to the mineral or

72
Chapter 3: Geological Sources of Metals in Coal and Coal Products

organic components, is speculative or known only in general terms. This is in large


part a result of the fine-grained nature of coal mineral matter and coal utilisation
by-products, rendering the application of conventional bulk analytical techniques to
individual mineral phases extremely difficult. Indirect assessments such as selective
leaching, float-sink studies or relations between mineral and element abundances
have provided much useful information, with relevance to coal utilisation and envi-
ronmental studies. Microbeam techniques have considerable promise for the direct
determination of trace elements and their distribution in individual organic particles
and mineral grains, but have yet to find widespread application in evaluating the
mode of trace element occurrence in coals or their by-products.
The absence of definitive data on the modes of occurrence of particular elements
limits the capacity for understanding how those elements may respond in different
types of environmental situations. In view of the anticipated increases in coal use
over the coming decades, knowledge of the mode of occurrence and behaviour of
trace elements in coal and its by-products will nevertheless be essential in develop-
ing effective strategies for minimising the environmental footprint arising from this
long-standing but still effective energy source.

3 Key Points
• A wide range of metals and other trace elements are found
in coal, depending mainly on the rank, type and grade
of the coal concerned and the nature of the minerals and
other inorganics associated with the organic components.
• Some of the minerals were introduced when the origi-
nal coal-forming peat was laid down, either as sediment
grains, skeletal particles or authigenic precipitates, and
some were introduced after deposition, from solutions
permeating the pore spaces and fractures of the coal bear-
ing sedimentary succession.
• A number of different methods may be used to identify
the minerals, other inorganics and trace elements in coal,
including chemical analysis, X-ray diffraction, optical
and electron microscopy, and a range of sophisticated mi-
crobeam techniques.
• The nature of the minerals and the intimacy of their as-
sociation with the organic matter principally determine
how the metals and other elements are partitioned during
coal mining, preparation and use, and how the coal and
its products or waste materials may respond in different
types of environmental situations.

73
METAL CONTAMINANTS I N N E W Z E A L A N D

REFERENCES

Anon, 2002. New Zealand Mining. Crown Minerals, Ministry of Economic Development of
New Zealand, Vol 31, 40 pp.
Anon, 2003. Minerals and Coal Annual Report 2002. Crown Minerals, Ministry of Economic
Development of New Zealand, 16 pp.
Belkin, H.E., Zheng, B., Zhou, D., Finkelman, R.B., 1997. Preliminary results on the
geochemistry and mineralogy of arsenic in mineralized coals from endemic arsenosis areas
in Guizhou Province, P.R. China. Proceedings of 14th Annual Pittsburgh International Coal
Conference, September, 1997, Taiyuan, Shanxi, China, 20 pp (CD-ROM publication)
Benson, S.A., Holm, P.L., 1985. Composition of inorganic constituents in three low-rank
coals. Industrial and Engineering Chemistry, Product Research and Development 24, 145-
149.
Berkowitz, N., 1979. An Introduction to Coal Technology, Academic Press, New York, 345
pp.
Bohor, B.F., Gluskoter, H.J., 1973. Boron in illite as an indicator of palaeosalinity in Illinois
coals. Journal of Sedimentary Petrology 43, 945-956,
Bohor, B.F., Triplehorn, D.M., 1993. Tonsteins: altered volcanic-ash layers in coal-bearing
sequences. Geological Society of America Special Paper 285, 44 pp.
Bowen, H.J.M., 1979. Environmental Chemistry of the Elements. Academic Press. London,
316 pp.
Bragg, L.J., Oman, J.K., Tewalt, S.J., Oman, C.J., Rega, N.H., Washington, P.M., Finkelman,
R.B., 1997. National Coal Resources Data System -- U.S. coal quality database. U.S.
Geological Survey Open-File Report 97-134.
Brockway, D.J., Borsaru, R.M., 1985. Ion concentration profiles in Victorian brown coals.
Proceedings of International Conference on Coal Science, Sydney, Australia, October,
1985, Pergamon Press, pp. 593-596.
Brown, H.R., Durie, R.A., Schafer, H.N.S., 1959. The inorganic constituents of Australian
coals: I – the direct determination of the total mineral matter content. Fuel 38, 295-308.
Brownfield, M.E., Affolter, R.H., Stricker, G.D., Hildebrand, R.T., 1995. High chromium
contents in Tertiary coal deposits of northwestern Washington – a key to their depositional
history. International Journal of Coal Geology 27, 153-169.
Burchill, P., Richards, D.S., Warrington, S.B., 1990. A study of the reactions of coals and coal
minerals under combustion-related conditions by thermal analysis-mass spectrometry and
other techniques. Fuel 69, 950-956.
Chenery, S., Querol, X., Fernandez-Turiel, J.L., 1995. Quantitative determination of trace
element affinities in coal and combustion wastes by Laser Ablation Microprobe Inductively
Coupled Mass Spectrometry. Coal Science, Proceedings of the Eighth International
Conference on Coal Science 1, 327-330.
Clarke, L.B., 1993. The fate of trace elements during coal combustion and gasification: an
overview. Fuel 72, 731-736.
Cobb, J.C., 1985. Timing and development of mineralized veins during diagenesis in coal beds.
In: Cross, A.T. (editor), Compte Rundu of 9th International Conference on Carboniferous
Stratigraphy and Geology, Washington, D.C. and Champaign-Urbana, May 1979, Volume
4: Economic Geology – Coal, Oil and Gas. Southern Illinois University Press, Carbondale,
pp. 371-376.
Cook, A.C., 1962. Fluorapatite petrifactions in a Queensland coal seam. Australian Journal
of Science 25, 94.
Dale, L.S., 2003. Review of Trace Elements in coal. Project C11020, End of Grant Report.
CSIRO, Energy Technology Investigation Report 599, 59pp.
Davidson, R.M., 2000. Modes of occurrence of trace elements in coal. IEA Coal Research,
London, 36 pp.
Davidson, R.M., Clarke, L.B., 1996. Trace Elements in Coal. IEA Coal Research, London,
60 pp.

74
Chapter 3: Geological Sources of Metals in Coal and Coal Products

Davis, A., Russell, S.J., Rimmer, S.M., Yeakel, J.D., 1984. Some genetic implications of silica
and aluminosilicates in peat and coal. International Journal of Coal Geology 3, 293-314.
Demir, I., Ruch, R.R.,Damberger, H.H., Harvey, R.D., Steele, J.D., Ho, K.K., 1997.
Environmentally critical elements in channel and cleaned samples of Illinois coals. Fuel
77, 95-107.
Diessel, C.F.K., 1992. Coal-bearing Depositional Systems. Springer-Verlag, Berlin, 721pp.
Diessel, C.F.K, Smyth, M., 1995. Petrographic constituents of Australian coals. In: C.R.
Ward, H.J. Harrington, C.W. Mallett and J.W. Beeston (Editors), Geology of Australian
Coal Basins. Geological Society of Australia Coal Geology Group, Special Publication 1,
247-298.
Ellis, A.J., 1977. Concentration of some trace elements in New Zealand coals. New Zealand
Division of Scientific and Industrial Research Bulletin 218, 132-137.
Faraj, B.S.M., Fielding, C.R., Mackinnon, D.R., 1996. Cleat mineralisation of Upper Permian
Baralaba/Rangal Coal Measures, Bowen Basin, Australia. In: Gayer, R., Harris, I. (editors),
Coalbed Methane and Coal Geology, Geological Society Special Publication 109, 151-164.
Finkelman, R.B., 1982. Modes of occurrence of trace elements and minerals in coal: an
analytical approach. In: Filby, R.H., Carpenter, B.S., Ragrini, R.C. (editors), Atomic and
Nuclear Methods in Fossil Energy Research. Plenum Publishing Corporation, pp. 141-
149.
Finkelman, R.B., 1994a. Abundance, source and mode of occurrence of the inorganic
constituents in coal. In: Kural, O. (editor), Coal. Istanbul Technical University, pp. 115-
125.
Finkelman, R.B., 1994b. Modes of occurrence of potentially hazardous elements in coal:
levels of confidence. Fuel Processing Technology 39, 21-34.
Finkelman, R.B., Stanton, R.W., 1978. Identification and significance of accessory minerals
from a bituminous coal. Fuel 57, 763-768.
Finkelman, R.B., Orem, W., Castranova, V., Tatu, C.A., Belkin, H.E., Zheng, B., Lerch, H.E.,
Maharaj, S.V., Bates, A.L., 2002. Health impacts of coal and coal use: possible solutions.
International Journal of Coal Geology 50, 425-443.
Frazer, F.W., Belcher, C.B., 1973. Quantitative determination of the mineral matter content of
coal by a radio-frequency oxidation technique. Fuel 52, 41-46.
French, D., Dale, L., Matulis, C., Saxby, J., Chatfield, P., Hurst, H.J., 2001. Characterization
of mineral transformations in pulverized fuel combustion by dynamic high-temperature
X-ray diffraction analyser. Proceedings of 18th Pittsburgh International Coal Conference,
Newcastle, Australia, December, 2001, 7 pp.
Galbreath, K.C, Brekke, K.W., 1994. Feasibility of combined wavelength/energy-dispersive
computer-controlled scanning electron microscopy for determining trace metal distribution.
Fuel Processing Technology 39, 63-72.
Gary, M., McAfee, R., Wolf, C.L. (editors), 1972. Glossary of Geology. American Geological
Institute, Washington, DC, 805 pp.
Given, P.H., Spackman, W., 1978. Reporting of analyses of low rank coals on the dry, mineral-
matter-free basis. Fuel 57, 319.
Glick, D.C., Davis, A., 1987. Variability in the inorganic element content of U.S. coals
including results of cluster analysis. Organic Geochemistry 11, 331-342.
Gluskoter, H.J., 1965. Electronic low temperature ashing of bituminous coal. Fuel 44, 285-
291.
Gluskoter, H.J., Ruch, R.R., Miller, W.G., Cahill, R.A., Dreher, G.B., Kuhn, J.K., 1977. Trace
elements in coal: occurrence and distribution. Illinois State Geological Survey Circular 499,
154 pp.
Gray, V.R., 1986. The chemical properties and composition of Mokau coal. New Zealand
Journal of Geology and Geophysics 29, 447-461.
Greb, S.F., Eble, C.F., Chesnut, D.R., 1999. An in situ occurrence of coal balls in the Amburgy
Coal Bed, Pikeville Formation (Duckmantian), Central Appalachian Basin, USA. Palaios
14, 432-450.

75
METAL CONTAMINANTS I N N E W Z E A L A N D

Gupta, R., Wall, T.F., Baxter, L.A. (editors), 1999. The Impact of Mineral Impurities in Solid
Fuel Combustion, Plenum Publishers, New York, 768 pp.
Hatch, J.R., Gluskoter, H.G., Lindahl, P.C., 1976. Sphalerite in coals from the Illinois Basin.
Economic Geology 71, 613-624.
Hickmott, D.D., 1993. PIXE microanalysis of Fe-sulfides: comparison of Eastern and Western
US coals. Seventh International Conference on Coal Science, conference proceedings 1
– Poster proceedings.
Hickmott, D.D., Baldridge, W.S., 1995. Application of PIXE microanalysis to macerals and
sulfides from the Lower Kittanning coal of western Pennsylvania. Economic Geology 90,
246-254.
Hower, J.C., Calder, J.H., Eble, C.F., Scott, A.C., Robertson, J.D., Blanchard, L.J., 2000.
Metalliferous coals of the Westphalian A Joggins Formation, Cumberland Basin, Nova
Scotia, Canada: petrology, geochemistry and palynology. International Journal of Coal
Geology 42, 185-206.
Hower, J.C., Pollock, J.D., 1989. Petrology of the River Gem coal bed, Whitley County,
Kentucky. International Journal of Coal Geology 11, 227-245.
Hower, J.C., Robertson, J.D., 2003. Clausthalite in coal. International Journal of Coal
Geology 53, 219-225.
Hower, J.C., Robertson, J.D., Thomas, G.A., Wong, A.S., Schram, W.H., Graham, U.M.,
Rathbone, R.B., Robl, T.L., 1996. Characterization of fly ash from Kentucky power plants.
Fuel 75, 403-411.
Hower, J.C., Robertson, D.R., Wong, A.S., Eble, C.F. Ruppert, L.F., 1997. Arsenic and lead
concentrations in the Pond Creek and Fire Clay coal beds, eastern Kentucky coal field.
Applied Geochemistry 12, 281-289.
Hower, J.C., Ruppert, L.F., Eble, C.F., 1999. Lanthanide, yttrium, and zirconium anomalies
in the Fire Clay coal bed, Eastern Kentucky. International Journal of Coal Geology 39,
141-153.
Hower, J.C., Williams, D.A., Eble, C.F., Sakulpitakphon, T., Moecher, D.P., 2001. Brecciated
and mineralized coals in Union County, Western Kentucky coal field. International Journal
of Coal Geology 47, 223-234.
Huggins, F.E., 2002. Overview of analytical methods for inorganic constituents in coal.
International Journal of Coal Geology, 50, 169-214.
International Energy Agency, 2001. World Energy Outlook - 2000. International Energy
Agency. Paris, 460 pp
Jenkins, R., Snyder, R.L., 1996. Introduction to X-ray Powder Diffractometry. John Wiley
and Sons, New York, 403 pp.
Karr, C. (editor), 1978. Analytical Methods for Coal and Coal Products (Volumes 1-3),
Academic Press, New York.
Katrinak, K.A., Benson, S.A., 1995. Trace metal content of coal and ash as determined
using scanning electron microscopy with wavelength-dispersive spectrometry. American
Chemical Society, Division of Fuel Chemistry, Preprints 40, 71-79.
Kemezys, M., Taylor, G.H., 1964. Occurrence and distribution of minerals in some Australian
coals. Journal of the Institute of Fuel 37, 389-397.
King, J.G., Maries, M.B., Crossley, H.E., 1936. Formulas for the calculation of coal analyses
to a basis of coal substance free from mineral matter, Journal of the Society of Chemical
Industry 55, 277-281.
Kolker, A., Chou, C.-L., 1994. Cleat-filling calcite in Illinois Basin coals: trace element evidence
for meteoric fluid migration in a coal basin. Journal of Geology 102, 111-116.
Kolker, A., Finkelman, R.B., 1998. Potentially hazardous elements in coal: modes of occurrence
and summary of concentration data for coal components. Coal Preparation 19, 133-157.
Kortenski, J., 1992. Carbonate minerals in Bulgarian coals with different degrees of
coalification. International Journal of Coal Geology 20, 225-242.
Kortenski, J., Kostova, I., 1996. Occurrence and morphology of pyrite in Bulgarian coals.
International Journal of Coal Geology 29, 273-290.

76
Chapter 3: Geological Sources of Metals in Coal and Coal Products

Laban, K.L., Atkin, B.P., 1999. The determination of minor and trace element associations
in coal using a sequential microwave digestion procedure. International Journal of Coal
Geology 41, 351-369.
Li, Z., 2002. Mineralogy and trace element of the Cretaceous Greymouth coals and their
combustion products, University of Canterbury, unpublished Ph.D thesis, Volume II,
179 pp.
Li, Z., Moore, T.A., Weaver, S.D., Finkelman, R.B., 2001. Crocoite: an unusual mode of
occurrence for lead in coal. International Journal of Coal Geology 45, 289-293.
Lindqvist, J.K., Isaaac, M.J., 1991. Silicified conifer forests and potential mining problems in
seam M2 of the Gore Lignite Measures (Miocene), Southland, New Zealand. International
Journal of Coal Geology 17, 149-169.
Miller, R.N. and Given, P.H., 1978. A geochemical study of the inorganic constituents of some
low-rank coals. Technical Report 1, U.S. Department of Energy Report FE-2494-TR-1,
314 pp.
Miller, R.N., Given, P.H., 1986. The association of major, minor and trace elements with
lignites: 1 - Experimental approach and study of a North Dakota lignite. Geochimica et
Cosmochimica Acta 50, 2033-2043.
Nankervis, J.C., Furlong, R.B., 1980. Phase changes in mineral matter of North Dakota
lignites caused by heating to 1200ºC. Fuel 55, 4-8.
Nawalk, A.J., Friedel, R.A., 1972. Concentration of minerals in coals by peroxidation of the
organic matter. Proceedings of Symposium on Chemicals and Oil from Coal, Central Fuel
Research Institute of India (1969), pp. 526-530.
O’Gorman, J.V., Walker, P.L., 1971. Mineral matter characteristics of some American coals.
Fuel 50, 135-151.
Palmer, C.A., Krasnow, M.R., Finkelman, R.B., d’Angelo, W.M., 1993. An evaluation of
leaching to determine modes of occurrence of trace elements in coal. In: Chiang, S.H.
(editor), Proceedings of 10th International Pittsburgh Coal Conference, School of
Engineering, University of Pittsburgh, pp. 1062-1067.
Parr, S.W., 1928. The classification of coal. University of Illinois Engineering Experimental
Station, Bulletin 180, 62 pp.
Querol, X., Chenery, S. 1995. Determination of trace element affinities in coal by laser
ablation microprobe inductively coupled plasma mass spectrometry. In: M.K.G. Whateley
and D.A. Spears (Editors), European Coal Geology, Geological Society Special Publication
82, 147-155.
Querol, X., Chinenon, S., Lopez-Soler, A., 1989. Iron sulphide precipitation sequence in
Albian coals from the Maestrazgo basin, southeastern Iberian Range, northeastern Spain.
International Journal of Coal Geology 11, 171-189.
Raask, E., 1985. Mineral Impurities in Coal Combustion: behaviour problems and remedial
measures. Hemisphere Publishing Corporation, New York, 484 pp.
Rao, C.P., Gluskoter, H.J., 1973. Occurrence and distribution of minerals in Illinois coals.
Illinois State Geological Survey Circular 476, 56 pp.
Rao, P.D., Walsh, D.E., 1999. Influence of environments of coal deposition on phosphorus
accumulation in a high latitude, northern Alaska, coal seam. International Journal of Coal
Geology 38, 261-284.
Rapp, A.R, Hower, J.C., Peters, D.C., 1998. “Atlas of Coal Geology”. American Association
of Petroleum Geologists Studies in Geology 45 (CD-ROM).
Raymond, R., Andrejeko, M.J. (eds), 1983. Mineral Matter in Peat: its Occurrence, Form and
Distribution. Los Alamos National Laboratory Report LA9907 OBES, Los Alamos, New
Mexico, 242 pp.
Rees, O.W., 1966. Chemistry, uses and limitations of coal analyses. Illinois State Geological
Survey Report of Investigations 220, 55 pp.
Reifenstein, A.P., Kahraman, H., Coin, C.D.A., Calos, N.J., Miller, G., Uwins, P., 1999.
Behaviour of selected minerals in an improved ash fusion test: quartz, potassium feldspar,
sodium feldspar, kaolinite, illite, calcite, dolomite, siderite, pyrite and apatite. Fuel 78,
1449-1461.

77
METAL CONTAMINANTS I N N E W Z E A L A N D

Ren, D., Zhao, F., Wang, Y., Yang, S., 1999. Distributions of minor and trace elements in
Chinese coals. International Journal of Coal Geology 40, 109-118.
Riley, K.W., French, D., Nurukawa, T., 2002. The significance of trace elements in power
generation. Japan-Australia Coal Research Workshop, Tokyo, November 11-13, 2002
(unpublished).
Ruppert, L.F., Stanton, R.W., Cecil, C.B., Eble, C.F., Dulong, F.T., 1991. Effects of detrital
influx in the Pennsylvanian Upper Freeport peat swamp. International Journal of Coal
Geology 17, 95-116.
Russell, S.J., Rimmer, S.M., 1978. Analysis of mineral matter in coal, coal gasificaiton ash,
and coal liquefaction residues by scanning electron microscopy and X-ray diffraction. In:
Karr, C. (editor), Analytical Methods for Coal and Coal Products, Academic Press, New
York, 3, pp. 133-162.
Sajwan, K.S., Alva, A.K., Keefer, R.F. (editors), 2003. Chemistry of Trace Elements in Fly Ash.
Kluwer Academic, New York, 346 pp.
Sakulpitakphon, T., Hower, J.C., Schram, W.H. and Ward, C.R., 2004. Tracking mercury
from the mine to the power plant: geochemistry of the Manchester Coal Bed, Clay County,
Kentucky. International Journal of Coal Geology 57, 127-141.
Scott, A.C., 2002. Coal petrology and the origin of coal macerals: the way ahead? International
Journal of Coal Geology 50, 119-134.
Scott, A.C., Mattey, D.P., Howard, R., 1996. New data on the formation of Carboniferous
coal balls. Review of Palaeobotany and Palynology 93, 317-331
Sim, P.G., Lewin, J., 1975. Potentially toxic trace metals in New Zealand coals. New Zealand
Journal of Science 18, 635-641.
Soong, R., Berrow, M.L., 1979. Mineral matter in some New Zealand coals; 2, Major and
trace elements in some New Zealand coal ashes. New Zealand Journal of Science 22,
229-233.
Soong, R., Godbeer, W.C., Swaine, D.J., 1984. The determination of fluorine in some New
Zealand coals by the fluorine ion-selective electrode method. New Zealand Journal of
Science 27, 151-154.
Spears, D.A., 1987. Mineral matter in coal with special reference to the Pennine coalfields. In:
A.C. Scott (Editor), Coal and Coal-bearing Strata – Recent Advances. Geological Society
Special Publication 32, 171-185.
Standards Australia, 1993. Guide to the technical evaluation of hard coal deposits, Australian
Standard 2519, 60pp.
Standards Australia, 1995. Coal and coke – glossary of terms. Australian Standard 2418,
180 pp.
Standards Australia, 2000. Higher rank coal - mineral matter and water of constitution.
Australian Standard 1038 Part 22, 20 pp.
Stanton, R.W., Finkelman, R.B., 1979. Petrographic analysis of bituminous coal: optical and
SEM identification of constituents. Scanning Electron Microscopy 1, 465-471.
Swaine, D.J., 1990. Trace Elements in Coal. Butterworths, London, 278 pp.
Swaine, D.J., Goodarzi, F. (editors), 1995. Environmental Aspects of Trace Elements in Coal.
Kluwer Academic Publishers, Dordrecht.
Sykes, R., Lindqvist, J.K., 1993. Diagenetic quartz and amorphous silica in New Zealand
coals. Organic Geochemistry 20(6), 855-866.
Tarriba, P.J., Gamson, P.D., Warren, J.K., 1995. Secondary mineralisation in coal seams,
Hunter Valley Coalfield, New South Wales: unique mode of occurrence for Sr-Ba-Ca
carbonates. In: Boyd, R.L. and McKenzie, G.A. (eds), Proceedings of the 29th Newcastle
Symposium, "Advances in the Study of the Sydney Basin", Department of Geology,
University of Newcastle, Newcastle, New South Wales, pp. 87-93.
Taylor, G.H., Teichmuller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998. Organic
Petrology. Gebruder Borntraeger, Berlin, 704 pp
Thomas. L., 2002. Coal Geology. John Wiley and Sons, London, 384 pp.
Vassilev, S.V., Kitano, K., Takeda, S., Tsurure, T., 1995. Influence of mineral and chemical
composition of coal ashes on their fusibility. Fuel Processing Technology 45, 27-51.

78
Chapter 3: Geological Sources of Metals in Coal and Coal Products

Vassilev, S.V., Tascon, J.M.D., 2002. Methods for characterisation of inorganic and mineral
matter in coal: a critical overview. Energy and Fuels 17, 271-281.
Vorres, K.S. (editor), 1986. Mineral Matter and Ash in Coal. American Chemical Society
Symposium Series 301, 537 pp.
Ward, C.R., 1974. Isolation of mineral matter from Australian bituminous coals using
hydrogen peroxide. Fuel 53, 220-221.
Ward, C.R., 1978. Mineral matter in Australian bituminous coals. Proceedings, Australasian
Institute of Mining and Metallurgy 267, 7-25.
Ward, C.R., 1980. Mode of occurrence of trace elements in some Australian coals. Australian
Coal Geology 2, 77-98.
Ward, C.R. (editor), 1984. Coal Geology and Coal Technology. Blackwell Scientific
Publications, Oxford, 345 pp.
Ward, C.R., 1989. Minerals in bituminous coals of the Sydney Basin (Australia) and the
Illinois Basin (U.S.A.). International Journal of Coal Geology 13, 455-479.
Ward, C.R., 1991. Mineral matter in low-rank coals and associated strata of the Mae Moh
Basin, northern Thailand. International Journal of Coal Geology 17, 69-93.
Ward, C.R., 1992. Mineral matter in Triassic and Tertiary low-rank coals from South
Australia. International Journal of Coal Geology 20, 185-208.
Ward, C.R., 2002. Analysis and significance of mineral matter in coal seams. International
Journal of Coal Geology 50, 135-168.
Ward, C.R., Corcoran, J.F., Saxby, J.D., Read, H.W., 1996. Occurrence of phosphorus
minerals in Australian coal seams. International Journal of Coal Geology 31, 185-210.
Ward, C.R., French, D., 2003. Determination of glass content and estimation of glass
composition in fly ash using quantitative X-ray diffractometry. Proceedings of 12th
International Conference on Coal Science, Cairns, Queensland, 2-6 November, 11 pp (CD
publication).
Ward, C.R., Matulis, C.E., Taylor, J.C., Dale, L.S., 2001. Quantification of mineral matter in
Argonne Premium Coals using interactive Rietveld-based X-ray diffraction. International
Journal of Coal Geology 46, 67-82.
Ward, C.R., Spears, D.A., Booth, C.A., Staton, I., Gurba. L.W., 1999. Mineral matter and
trace elements in coals of the Gunnedah Basin, New South Wales, Australia. International
Journal of Coal Geology 40, 281-308.
Ward, C.R., Swaine, D.J., 1995. Minerals and inorganic constituents. In “Geology of
Australian Coal Basins” (ed. C.R. Ward, H.J. Harrington, C.W. Mallett and J.W. Beeston),
Geological Society of Australia Coal Geology Group, Special Publication 1, 93-109.
Wedepohl, K.H., 1995. The composition of the continental crust. Geochimica et Cosmochimica
Acta 59, 1217-1232.
Williams, E.G., Keith, L., 1963. Relationship between sulphur in coals and the occurrence of
marine roof beds. Economic Geology 58, 720-729.
Winburn, R.S., Grier, D.G., McCarthy, G.J., Peterson, R.B., 2000. Rietveld quantitative X-ray
diffraction analysis of NIST fly ash standard reference materials. Powder Diffraction 15,
163-172.
Zodrow, E.L., Cleal, C.J., 1999. Anatomically preserved plants in siderite concretions in the
shale split of the Foord seam: mineralogy, geochemistry, genesis (Upper Carboniferous,
Canada). International Journal of Coal Geology 41, 371-393.

79
METAL CONTAMINANTS I N N E W Z E A L A N D

80

View publication stats

You might also like