Anda di halaman 1dari 45

Coal Mineral

Transformations
– Effects on
Boiler Ash
Behaviour

Report No.
COAL R278
DTI/Pub
URN 05/659

March 2005
by

F. Wigley and J. Williamson

Department of Materials
Imperial College of Science, Technology and Medicine,
London,
SW7 2AZ

Tel: 020 7594 6747


Email: jim.williamson@imperial.ac.uk

The work described in this report was carried out under contract as
part of the DTI Cleaner Coal Research and Development
Programme. The programme is managed by Mott MacDonald Ltd.
The views and judgements expressed in this report are those of the
contractor and do not necessarily reflect those of the DTI or Mott
MacDonald Ltd

First published 2005


© Imperial College copyright 2005

iii
iv
Contents

Page

Executive Summary 4

1. Introduction 5

2. Project Overview - Aims and Objectives – Activities 6

3. Coals and Minerals Chosen for the Project 8

4. Coal Combustion Characterisation 10

5. Generation and Characterisation of Ashes from Coal and 13


Mineral Combinations

6. Mineral Transformations and Ash Formation Processes 20

7. Ash Particle Interactions and Transformations 23

8. Implications of Findings for Power Station Performance 27

9. Ash Sales and Ash Disposal 33

10. Conclusions 37

11. Further Work 38

12. References 39

Acknowledgements 40

Appendices 41

3
Executive Summary

The UK now imports more than 50% of the coal that is used for coal-fired power
generation. UK generators are offered an increasingly wider range of world-traded
coals for burning in boilers that were designed to burn a relatively narrow range of
indigenous coals. This project was undertaken to provide UK boiler designers and
operators with an improved knowledge of the combustion characteristics of coals for
which they had little combustion experience. The study placed particular emphasis
on the effects that a wider range of coal minerals and mineral matter distributions
might have on the many aspects of boiler operation. These ranged from coal grinding
for pulverised coal combustion, to combustion behaviour, levels of unburned carbon
in ash, precipitator performance, gaseous and particulate emissions, and the
slagging and fouling characteristics of the ash.

The coals were selected to reflect the wide range of world-traded coals that are now
on offer and came from North and South America, Australia, South Africa, Indonesia,
China, Russia and India. The coals were chosen on the basis of the ash content and
ash chemistry that UK utilities might encounter. As a consequence of the varied
geographical origins of the coals and the range of ash chemistry, the nature and
distribution of the mineral matter in the coals was found to be significantly different
from that of indigenous coals.

Coal and mineral matter characterisation was carried by Nottingham University and
Imperial College London. Combustion studies were undertaken by E.ON, using the
Combustion Test Facility (CTF) at the Ratcliffe Power Technology Centre and by
Imperial College, using a high temperature Entrained Flow Reactor (EFR). In addition
the EFR was used to study the mineral transformations of the minerals found in the
suite of coals. The combustion facilities generated a range of samples for analysis
and characterisation, including combustion ash and unburned char, cyclone ashes
and deposits collected on ceramic probes and a slag panel. Characterisation of the
samples enabled the combustion performance and slagging propensity of the coals
to be assessed and ranked against that of a typical UK bituminous coal (Harworth).

Some of the coals would be unsuitable for UK boilers. Two coals from the US
Powder River Basin had a high slagging and fouling potential, a high ash coal from
India could give potential ash handling problems unless blended with a low ash coal,
and a South African coal gave high NOx and high levels of unburned carbon. The
remaining coals would be expected to give few operational problems.

The implications of burning a wider range of imported coals have been considered.
Sales of boiler ashes to the construction market are an important consideration in the
overall economics of coal-fired power generation. Several of the ashes with a high
calcium content would be unlikely to meet current and anticipated specifications for
use with cements and concrete.

Existing methods of coal and ash characterisation were found to be generally


satisfactory in predicting the combustion performance of the coals burned at rig
scale. The more advanced coal and ash characterisation techniques were found
valuable in understanding the mineral transformations, the ash formation and ash
deposition mechanisms.

4
1. Introduction
It is widely acknowledged that pulverised coal combustion will continue to be a major
source of electrical power generation for the foreseeable future. Existing power
stations will almost certainly seek to use a wider range of coals, while tighter
emissions limits and greater demand for plant availability will challenge the power
station operators; similar constraints will no doubt apply to newly built plant. The
inorganic components of coal – the mineral matter – impact directly on each of these
constraints.

During combustion, the coal minerals transform to ash particles, some of which may
stick to the boiler walls and interact to form deposits. Coal ash deposition is one of
the factors affecting routine boiler operation, and occasionally causes severe
restrictions on plant availability. Coal ash deposition has been investigated during a
previous DTI-funded project, but to date there has been no reliable method for
predicting ash deposition that meets the needs of both coal purchasers and boiler
designers. Expensive test runs on rig-scale combustion facilities are currently used to
predict boiler performance. While computational fluid dynamic (CFD) ability and
computing power have greatly improved recently, computer modelling of deposition
in full-scale boilers is restricted by a lack of fundamental knowledge about the nature
and formation of fly ash particles.

The wide range of coals being considered for import to the UK, or for use in overseas
power stations, introduces new coal minerals and expands the variation in mineral
abundance and distribution in the pulverised coal. For example, in Southern
Hemisphere coals the major iron-bearing mineral is usually siderite [Fe(CO3)], while
Northern Hemisphere coals typically contain pyrite [FeS2]. There are also distinctive
differences in the nature of the association between the clay minerals and the coal
organic material. The transformation of siderite during combustion is not understood;
the same situation applies to calcite [Ca(CO3)], a common mineral in many coals.
Although the transformation processes that convert pyrite into iron-rich ash particles
are known, the kinetics established from laboratory studies do not agree with
observations made at power stations.

Most of the ash particles leave the boiler and are removed from the gas stream by
electrostatic precipitators, but a small fraction of the fine particles remain in the flue
gas and are emitted from the stack. The nature and amount of char in the ash has a
significant effect on precipitator performance. The nature of the fly ash has a
significant impact on ash disposal, with sales as a cement replacement material
being preferable to incurring landfill taxes. A more detailed understanding of coal ash
formation processes and the nature of the fly ash, especially the char and finest ash
particles, would allow a better appreciation of the options available to meet the fine
particle emission limits and to improve the economic value of the ash.

Thus an improved understanding of the fundamental issues of ash formation


processes and the nature of the fly ash produced in pulverised coal combustion
should be beneficial to UK companies purchasing coal and running power stations in
the UK or abroad, and would also benefit companies involved in designing new
power stations. The DTI-supported CoMiT project aims to provide a more complete
and quantitative description of the coal combustion characteristics, mineral matter
transformations and the ash formation process during pulverised coal combustion.

5
A suite of coals has been chosen for this project by the collaborators, primarily E.ON
and TXU Europe Power. The suite includes coals from North and South America,
Australia, South Africa, Indonesia, China, Russia and India. The mineral behaviour of
these coals has been studied using the E.ON Combustion Test Facility (CTF) and the
Imperial College Entrained Flow Reactor (EFR). The mineral matter in the coals and
the nature of the deposits generated during the CTF and the EFR trials have been
characterised at Imperial College.

2. Project overview
The project was a collaboration between Imperial College London, E.ON (formerly
Powergen), Rio Tinto Technology, TXU Energy and the University of Nottingham.
Imperial College London acted as the main contractor, and the University of
Nottingham was a sub-contractor. TXU Energy was bought by Powergen during the
course of the project; Mr PR Cooper continued his involvement with the project as an
independent consultant. The personnel involved and contact details of each
collaborating organisation are listed in Appendix 1. Just under half of the project
costs were provided by the DTI under the Cleaner Coal R&D Programme, the
remainder being provided by E.ON, Rio Tinto, TXU Energy and Imperial College
London. In-kind contributions were made by E.ON, Rio Tinto, TXU Energy and
Imperial College London.

2.1 Aims and Objectives


The aim of the project was to provide a more complete and quantitative description of
coal combustion characteristics, mineral transformations, and of the nature of the ash
produced during pulverised coal combustion. The data would provide an improved
basis for computer modelling of ash formation. Specific objectives were to:
• Improve the understanding of the transformations undergone by major coal
minerals during pulverised coal combustion.
• Provide quantitative data on the rates of coal mineral transformations.
• Provide quantitative descriptions of boiler ashes, for CFD modelling of boilers
and precipitators, especially for the char and finer ash particles.
• Improve the understanding of the effect of coal mineral transformations on the
processes of boiler ash formation, deposition, emission, ash usage and
disposal.
The project aimed to deliver:
• Quantitative data describing boiler ashes and their parent coal minerals,
suitable for CFD modelling of ash behaviour in pulverised coal fired boilers
and precipitators. A wide range of UK and foreign coals to be included in the
study.
• Improved the understanding of the transformations undergone by major coal
minerals during pulverised coal combustion, including quantitative data on the
rates of coal mineral transformations.
• Improved understanding of the effect of coal mineral transformations on the
processes of boiler ash formation and deposition, together with an
assessment of the implications for wider aspects of power station operation
including coal selection, preparation, emission and usage.

6
• Quantitative data on the combustion performance of UK and overseas coals
and, in particular, a description of the properties of char that affect ash
precipitation and usage.

The work programme consisted of six main activities.


2.2 Activity A – Coal selection and mineral characterisation
The nature, distributions and associations of the minerals in a wide range of power
station coals have been characterised using CCSEM and other techniques. Rio
Tinto, TXU Energy and E.ON all helped with the selection and acquisition of the most
appropriate coals for the programme. The chemical composition and size distribution
of coal minerals is complex, and CCSEM was used as the technique to provide the
necessary information in a quantitative manner. The coal minerals form the starting
point for the transformation to ash in pulverised coal combustion. Samples of coal
minerals were acquired by Imperial College London.

2.3 Activity B – Determination of combustion characteristics


Basic combustion characteristics of the test coals were determined at Nottingham
University using the ‘Reactivity Assessment Programme’ and standard rank and
maceral analysis techniques. Coal char samples were prepared using the drop tube
furnace, and char samples from the rig-scale trials on the CTF have been examined.
Nottingham personnel worked closely with the Imperial College team. This part of the
programme was aimed at determining the affect of ash properties on the behaviour in
electrostatic precipitators.

2.4 Activity C – Generation of ashes from coals and mineral combinations


Pulverised coal samples and naturally occurring minerals have been transformed
under conditions representing pulverised coal combustion in the EFR and ash and
deposit samples collected on deposition probes. Some coals, especially those with
the higher iron contents, were also transformed under lower oxygen conditions to
investigate the effects of the reduced oxygen levels present in low-NOx burners.
Some of the same coals were also be burnt in the Powergen CTF, producing ash and
deposit samples to compliment the EFR studies. The minerals were fired singly and
in combination to investigate the nature of interactions between minerals, including
such processes as the transfer of calcium from calcite to aluminosilicate ash
particles.

2.5 Activity D – Investigation of ash formation processes


The ash samples from the EFR and the CTF have been characterised, using CCSEM
and other techniques, to provide a quantitative description of the combined size and
chemical distribution of the ash particles. Particular emphasis has been placed on the
characterisation of the smallest particles, which contribute preferentially to particulate
emissions. The resulting data has been used to determine and quantify the effects of
temperature, residence time, oxygen concentration and association with coal organic
material on the transformations of major coal minerals.

2.6 Activity E – Investigation of ash particle interactions following deposition


The deposit samples from the EFR and the CTF have been characterised to
investigate the interaction between ash particles following deposition. The effect of
the degree of mineral transformation on the tendency to deposit and on the
7
properties of the deposit has been determined. The interaction between ash particles
following deposition was studied to establish the effects of the physical and chemical
nature of ash particles on the rates of deposit sintering.

2.7 Activity F – Implications of coal ash mineral transformations on power


station operation
The implications of the findings from the project on the wider aspects of the operation
of pulverised coal fired power boilers have been considered, including such issues
as:
• The range of acceptable coals for UK pulverised coal combustion
• Coal cleaning, blending and milling
• Deposition behaviour
• The collection efficiency and environmental impact of char and fine ash
particles
• The marketing potential of the precipitator ash.

2.8 Activity G – Project management and reporting


Progress meetings took place at regular intervals throughout the project. Imperial
College London provided financial and administrative organisation for the project.
Reports were produced on each project activity.

3. Coals and minerals chosen for the project

Coals were selected to cover the wide range of ash chemistries of coals
currently imported or under consideration for pulverised combustion at UK
utilities. Table 3.1 shows the suite of coals selected for combustion trials on
the E.ON Power Technology’s Combustion Test Facility (CTF) and on the
Imperial College Entrained Flow Reactor (EFR). Harworth, a Nottinghamshire
coal, was chosen as the baseline UK coal against which the performance of the
other coals could be assessed.

Several tonnes of each coal were acquired and ground in a Lopulco vertical spindle
mill (subcontractor James Durrans and Sons) to a nominal specification of 65 ±3%
<75µm and <1% >300µm. Previous experience with Durrans has shown that this
specification results in a size distribution that is closer to that obtained from full-sized
milling plant, rather than using the normal PF specification.

The proximate, ultimate and ash analyses for each of the coals in the above table is
given in Appendix 2.

A further selection of fourteen coals, for which only small quantities were available
(kg’s), were chosen for mineral characterisation and combustion studies on the EFR.
The coals chosen for this part of the study are listed below in Table 3.2

A selection of minerals commonly found as major minerals in coal were acquired for
studying mineral transformation and mineral interaction processes. The minerals
chosen are shown in Table 3.3.

8
Coal Origin Characteristics
Harworth UK Baseline UK coal for
comparison
Bailey USA Typical Eastern US bituminous
coal, relatively high Fe
PRB 1 USA High Na Powder River Basin
coal
PRB 2 USA Typical low ash PRB coal
Bowen Basin Australian, NQ Rio Tinto reference coal
Hunter Valley Australian High quartz coal, high Al: Si
Prodeco Colombian Typical South American coal,
relatively high Fe
Goedehoop South Africa High Ca ash (lignitic ash), with
included clays
Talcher India High ash coal
Blends
PRB 1/ PRB 2 USA High Na PRB, blended with
(50:50) typical PRB (low ash)
PRB 1/Bailey USA High Na PRB, blended with
(50:50) Eastern USA coal (high ash)

Table 3.1 Coals selected for CTF and EFR combustion trials

Coal Origins Characteristics


Binungan Indonesia Low rank, very low ash, variable
Na and Si:Al
Blacksville USA Similar to Bailey
Callide Australia High siderite content reported
Daliuta China Very high Ca
Fila Mistra Venezuela Low rank, very high Fe and Ca
Lati A Indonesia As Binungan
Lati B Indonesia As Binungan
Lati C Indonesia As Binungan
Roto B Indonesia As Binungan
Kusbass Russia Little known about this coal,
reported variable quality
Park Slip UK Opencast, Welsh coal, high Fe
Pinang Indonesia Low rank, very low ash, high Na
Koornfontein South Africa High Ca
Spitzburgen Spitzburgen High Ca and Fe

9
Table 3.2 Additional coals selected for mineral characterisation and EFR
combustion trials

Mineral Nominal composition


Quartz SiO2
Kaolinite Al2Si2O5(OH)4
Illite (K,Na,Ca,Mg,Fe,Al,Si)14O20(OH)4
Pyrite FeS2
Calcite CaCO3
Siderite A FeCO3
Siderite B FeCO3
Dolomite CaMg(CO3)2
Apatite Ca5(PO4)3OH

Table 3.3 Minerals selected for transformation studies

4 Coal combustion characterisation


4.1 Petrographic Analysis of Coals and Chars
Petrographic analysis of the coals and chars was performed by the University of
Nottingham Fuel and Energy Centre to assist in the prediction of the gaseous
emissions and burnout performance of the coals. The information obtained was used
to compare the data obtained during the CTF trials with that predicted by
petrographic analysis.

Standard methods of maceral analysis were combined with novel techniques


developed by Nottingham University as part of the project. These include a Reactivity
Assessment Program (RAP) and an Automated Char Analysis (ACA) procedure. All
the coals used for the CTF trials were characterised, and in addition nine of the
additional coals that were used for the EFR trials were characterised. The RAP
technique involves the determination of the reflectance and fluorescence of blue and
white light from a mounted and polished specimen under carefully controlled
conditions. A key feature of the RAP program is the determination of the point on the
grey scale axis that accurately divides the macerals into what may be termed
‘reactive’ and ‘unreactive’ categories. The automated char analysis used an
automated image analysis system to remove the subjectivity present in a manual
characterisation. Developments in the char analysis procedures included the use of
coloured resins to help to distinguish char particles from the mounting resins, and
programs with the ability to separate touching particles. These, and other advances,
have been described by Lester et al (2002).

4.2 Predictions Based on Coal and Char Characterisation


In general, the correlation between predicted burnout of the coals and that realised in
rig combustion with 15% OFA and 1% excess oxygen was good, see Table 4.1.

Maceral analysis correctly identified the high inertinite containing Southern


hemisphere coals, Goedehoop, Bowen Basin and Callide, and by similar means the
highly reactive coals from Indonesia were found to contain more than 90% of vitrinite.

10
Reflectance measurements suggested the Goedehoop, Callide and Park Slip were
all blended coals.

Coal ’Unreactives’ Predicted Measured


(%) combustible combustible
loss (%) loss (%)
Harworth 5.2 0.75 0.74
Predoco 3.7 0.63 1.16
Goedehoop 21.3 2.01 1.98
Bowen Basin 6.8 0.88 0.86
Bailey (B) 5.6 0.78 0.78
Hunter Valley 9.8 1.11 1.06
Talcher 2.6 0.55 0.67
PRB 1 1.3 0.45 0.47
PRB 2 1.2 0.44 0.27
SC/B blend 3.4 0.61 0.52
SC/JR blend 1.5 0.46 0.16

Table 4.1 Comparison of measured and predicted combustible losses based on


petrographic analysis of the coals. Measured combustion loss were obtained
with15% OFA and 1% excess oxygen conditions on the CTF.

Based on the proportion of ‘unreactives’ as determined by the RAP program, the 20


coals were divided into four groups as shown in Table 4.2 Predeco (and several
other South American coals) have been the subject of further investigations in order
to explain their less than expected combustion performance. The presence of
inertodetrinite and micrinite macerals, macerals classified as relatively unreactive,
may interfere with the normal vesculation processes that produce thin walled chars
that can readily burn out [Cloke et al (2002)].

‘Unreactives’ Predicted burnout performance


<5% Burnout - good
Bailey, Blacksville, Fila Mistra, Binungan, Lati A, B and C,
Roto B, Talcher PRB 1, PRB 2.
Predoco (burnout performance was less than that
predicted)
5 - 10% Burnout - acceptable
Harworth, Hunter Valley, Bowen Basin, Callide, Kuzbass
10 - 15% Burnout - borderline/fairly poor
Park Slip
15 - 25% Burnout - poor
Goedehoop
Table 4.2 Predicted burnout performance of the coals
11
4.3 CTF Test Procedures
The standard methodology developed by E.ON for coal characterisation using the
CTF was followed for these trials. The tests were carried out using a standard low-
NOx PF burner with a tertiary/secondary air ratio of 3.5:1. Air staging using 15%
overfire air (OFA) to ports 204, 205 and 206 was used, although runs without OFA
were also undertaken for comparison with the effect of OFA on NOx production. The
majority of tests were conducted at a thermal input of 1MW, although deposition
samples were additionally taken at 1.2MW. A schematic diagram of the CTF,
showing the position of the sample collection ports, is shown in Appendix 3.

4.4 Pulverised Coal Size Analysis


Particle size analysis was made using a Malvern Size Analyser and showed that the -
75µm fractions were similar for all the samples, with the exception of the Bowen
Basin and Talcher coals. These coals were much finer, containing approximately
80% < 75µm, despite a specification for 65% < 75µm. It is known that Bowen Basin
has a Hardgrove index of (HGI) of around 60, and that Talcher contains much fine
quartz dispersed throughout the coal. These properties are discussed further in
Section 8.4, but could explain the finer size distribution produced by milling these
coals.

4.5 Gaseous Emissions


NO emissions were measured for each coal and then corrected to 3% O2. NO
emissions showed a linear reduction with decreasing O2 for all the coals tested
and, as expected, were noticeably lower with OFA in service, being up to 150
ppm lower at 4%O2. As might also have been expected from the coal analyses,
the Goedehoop and Bowen Basin coals gave the highest NO emissions. NO
emissions from both these coals were approximately 100ppm higher than
those from all the other coals, with and without OFA. PRB 1 and PRB 2 gave
the lowest NO emissions, approximately 100 ppm lower than any of the other
coals. Although OFA has a beneficial effect in reducing NO emissions, it is
important to assess any detrimental effect on the overall combustion efficiency
of the process. CO emissions were measured at the beginning and the end of
the 15% OFA condition and once without any OFA. Generally the OFA
appeared to have little effect on the position of the oxygen breakpoint for the
majority of the coals, the breakpoint lying between 0.2 – 0.3% for most coals,
the one exception being the Talcher coal with a significantly higher value at
high thermal inputs.

4.6 Combustion efficiency


Char samples were taken for characterisation at Nottingham University from the
lower part of the CTF furnace at 1% and 3% excess O2. Fly ash samples were taken
at a point equivalent to ultimate burnout at full scale and tested for LOI. Fly ash
samples were taken from a point equivalent to the entry to the convective pass, and
at the same point at which the deposition probe was inserted. Fly ashes were
characterised using CCSEM and other techniques at Imperial College.

12
5 Generation and characterisation of ashes from coal and mineral
combinations

5.1 CTF and EFR test procedures


The ash slagging propensities of the coals was studied by inserting deposition
probes into both the CTF and the EFR to collect deposits for subsequent
examination. Ceramic (mullite) coupons on an air cooled support were inserted into
the CTF at the beginning of the convective zone were the gas velocities were about
10m/s and the temperatures ranged from 1100°C to 1275°C, depending on the
thermal input. In the EFR an uncooled mullite probe was inserted into the reactor at a
point were the gas temperatures were about 1250°C and the residence times were 1
to 2 seconds. A furnace wall slagging panel was inserted into the CTF at a point
slightly above and facing the burner.

The nature of the CTF deposits and the ‘deposit types’ was assessed on a scale from
0 to 10, where 0 is loose dust and 10 is a fully fused deposit, see Table 5.1. Now
known as the Jones index, this index was first described by Jackson and Jones
(1989). Deposits were produced at temperatures ranging 1015° to 1240°C and the
data for each deposit extrapolated to give a value at 1200°C so that ashes/deposits
could be compared with one another.

Jones Descriptions Compressive strength of


Index deposit (kNm-2)
0 Dust 0
1 Mostly dust, some very light <50
sintering
2 Mostly lightly sintered ash 50-100
3 Coherent light scattering 100-150
4 Light/medium scattering 150-400
5 Coherent medium scattering 400-750
6 Medium/strong scattering 750-1000
7 Strongly sintered ash 1000-1500
8 Strongly sintered ash, some 1500-2500
fusion
9 Mostly fused ash particles 2500->3000
10 Hard, vitreous slag >>3000

Table 5.1 Jones Index - nature of deposits [Jackson and Jones (1989)]

Experience has shown that a visual assessment is at least as reliable as the more
difficult compressive strength measurements. The CCSEM characterisation provided
much more microstructural and chemical detail and could be used to explain
apparent anomalies in the visual predictions.

13
Samples of minerals and coals were subject to the time-temperature history of a
large pulverised coal boiler using the Entrained Flow Reactor at Imperial College.
This 5m vertical down-flow reactor has a peak temperature of 1600°C, dropping to
1200°C over a residence time of about two seconds. The EFR has been described
more fully by Hutchings et al [1996)].

5.2 EFR Test Procedures


The Entrained Flow Reactor was used to subject the coal and mineral particles to the
same time-temperature history that PF particles experience in a large pulverised coal
boiler. This 5m vertical down-flow reactor has a peak temperature of 1600°C,
dropping to 1200°C over a particle residence time of about two seconds. The EFR
has been described more fully by Hutchings et al [1996] and a schematic diagram of
the EFR is shown in Appendix 3.

Coal runs on the EFR involved feeding approximately 100g of coal over a period of
40-60 minutes. Each coal was run through the EFR twice, once to collect a deposit
on an un-cooled mullite deposition probe and once to collect an ash sample from the
combustion atmosphere. The deposition probe and the ash collection probe were
inserted at the same position – at a residence time of about one second and at a gas
temperature of about 1250°C. Ash samples were also collected in a cyclone at the
bottom of the EFR, after a residence time of about two seconds.

5.3 Mineral transformations in the EFR


Both single minerals and mineral pairs were used in the EFR. Ten grams of mineral
sample were entrained in the primary air of the EFR over a period of about ten
minutes, and ash was collected in a cyclone at the bottom of the EFR after a
residence time of about two seconds. Ash samples were also collected in a water-
cooled probe at a furnace position with a residence time of about one second and at
a gas temperature of about 1250°C. Although it was intended to characterise the
probe ash samples, the quantity collected was too small for analysis for some of the
runs and the cyclone ashes were characterised for consistency.

Mineral transformations in the EFR used both single minerals and a comprehensive
set of mineral pairs. The mineral pairs were prepared by adding 10, 25 or 50wt% of
pyrite, siderite A, siderite B, calcite or dolomite to quartz, kaolinite or illite. Mineral
pairs involving the same minerals were run consecutively, with increasing levels of
addition (0, 10, 25, 50 and 100wt%), to minimise the possible effects of sample carry-
over in the reactor.

5.4 Ashes and deposits collected from the CTF


Coal samples were also combusted on the CTF and ash samples were collected at
ports 722 and 740, as shown in Appendix 3. Each deposit was assessed and given a
slagging propensity using the Jones index, see Table 5.1. The CTF ash samples
were also characterised at Imperial College using CSEM and other techniques.

5.5 Sample characterisation using Computer Controlled Scanning Electron


Microscopy [CCSEM]
Ash samples were mounted in epoxy resin and prepared as a polished cross-section
perpendicular to the settling direction to eliminate the effects of density segregation.
Digital back-scattered electron images were collected at 100x and 500x
14
magnifications at random points across the sample cross-section. Clusters of pixels
with intensities corresponding to the average atomic number range of ash were
located, and their size and shape measured. Larger ash particles, analysed at 100x
magnification, were over-sampled during this analytical procedure and corrected for
abundance during the data interpretation. About 1000 chemical analyses of the areas
of the cross-section corresponding to those regions of the image were obtained by
energy-dispersive X-ray spectrometry [EDS].

Coal ash Jones Index


at 1200°C
Harworth 5.6
Bailey 3.25
PRB 1 9.2
PRB 2 7.2
Bowen Basin 3.25
Hunter Valley 0.75
Prodeco 2.75
Goedehoop 2.0
Talcher 2.25
PRB 2/ PRB 1 blend 8.25
PRB 1/ Bailey blend 4.8

Table 5.2 Slagging propensity of coal ash deposits, using the Jones Index for
CTF deposits at 1200°C

The data for each ash sample was interpreted in terms of the size, shape and
chemical composition of individual ash particles. Particles were assigned to a
chemical type, defined on the basis of their chemical type, as shown in Table 5.3.

Circular sections were cut through the mullite coupons and attached deposits,
mounted in epoxy resin and prepared as polished cross-sections. Digital back-
scattered electron images were collected at 100x and 500x magnifications, usually at
twenty points across the deposit cross-section. Pixels in the images with intensities
corresponding to the average atomic number range of deposit were located. A
regular grid of one hundred points was superimposed on each image, and points
lying on the deposit were identified. Deposit porosity was calculated for each image
from the fraction of points identified as lying on the deposit. Up to 1200 point
chemical analyses of the deposit were obtained and assigned to a chemical type on
the basis of their chemical composition. Definitions of chemical types are shown in
Table 5.3.

5.5.1 Slagging index based on CCSEM characterisation


The data for each deposit sample was interpreted in terms of porosity, and the
average chemical composition and chemical distribution of individual deposit point
analyses. In addition, a ‘slagging index’ was calculated for each deposit. This index,

15
which is described more fully by Russell (2002) estimates the potential of the deposit
to consolidate with time and to become difficult to remove. These studies have

Chemical type Compositions


non-CFAS (CaO + Fe2O3 + Al2O3 + SiO2) < 80 wt%
Si-rich SiO2' > 90 wt%
Fe-rich Fe2O3' > 90 wt%
Ca-rich CaO' > 90 wt%
Si-alsil (Al2O3' + SiO2') > 80 wt% and SiO2' > 70 wt%
Alsil (Al2O3' + SiO2') > 80 wt%
Ca-alsil Fe2O3' < 10 wt%
Fe-alsil CaO' < 10 wt%
Ca+Fe not previously assigned

Table 5.3 ‘Chemical types’ as used for CCSEM characterisation. Analyses are
assigned sequentially to the above ‘chemical types’ using the following scheme,
where CaO', Fe2O3', Al2O3' and SiO2' analyses are normalised to (CaO' + Fe2O3' +
Al2O3' + SiO2') = 100%.

shown that the slagging index of EFR deposits correlate well with slagging behaviour
observed during larger scale combustion.

The deposit slagging index was obtained using the following procedure:
1. Point analyses lying in the CaO-Al2O3-SiO2 system (CaO+Al2O3+SiO2 > 50%)
were selected and normalised to CaO+Al2O3+SiO2 = 100%
2. The percentage of analyses with normalised CaO contents between lying
between 5% and 40% was calculated.
3. Point analyses lying in the Fe2O3-Al2O3-SiO2 system (Fe2O3+Al2O3+SiO2 >
50%) were selected and normalised to Fe2O3+Al2O3+SiO2 = 100%
4. The percentage of analyses with normalised Fe2O3 contents lying between
10% and 50% were calculated.
5. The percentages in (2) and (4) were added together.
The resulting slagging index was a number between 0 and two hundred. Point
analyses that lie between both limits were included twice in recognition of the
synergistic effects that CaO and Fe2O3 together make to a deposit consolidation
[Russell (2002)].

5.6 Sample characterisation using X-ray diffraction analysis


Coal, ash and deposit were analysed by X-ray diffraction (XRD) to identify the
crystalline phases present. X-ray spectra were acquired over the range 5-65°2θ
using CuKα radiation. Diffraction peaks were identified by comparison with a
computerised database of diffraction spectra. The abundance of crystalline phases
was identified as either major, minor or trace, while the amount of glassy phase was
estimated from the size of the glassy ‘halo’ in the spectrum.

16
5.7 Results from EFR mineral trials
Samples of common coal minerals were reacted in the EFR to investigate their
behaviour when subject to the time-temperature history of a pulverised coal fired
boiler. Mineral pairs, consisting of clay (kaolinite or illite) with a Fe-bearing mineral
(pyrite, siderite A or siderite B) or with a Ca-bearing mineral (calcite or dolomite),
have been reacted in the same way. Ash samples from the reactor runs were
collected and characterised. The majority of ash particles from each run were derived
directly from a single mineral, in proportions directly related to the composition of the
mineral pairs.

Most of the ash particles were derived directly from a single parent mineral, but some
ash particles with intermediate compositions were produced by interaction. The clay-
derived ash particles showed small increases in iron oxide content through
interaction with pyrite or siderite, possibly by vaporisation and condensation; similar
changes in lime content were observed with calcite and dolomite interactions.
Ash particles of intermediate chemical composition were observed in all runs – iron-
aluminosilicate for the Fe-bearing mineral pairs and calcium-aluminosilicate for the
Ca-bearing mineral pairs, see Fig 5.1 The abundance of these intermediate ash
particles generally increased with the proportion of Fe- or Ca-bearing mineral. The
presence of ash particles with intermediate compositions indicates that physical
interaction between clay and other minerals had occurred during their transformation
in the EFR.

Since reacting mineral pairs in the EFR experience comparable conditions to


excluded minerals in a utility boiler, the low level of interaction between the majority
of ash particles observed in this study may be used to reinforce the wide-spread
observation that chemical equilibrium between coal ash particles is not achieved.
Predictions of coal ash behaviour based on thermodynamic equilibrium over the
entire ash stream are unlikely to be directly applicable to transformations in
pulverised coal fired boilers.

5.8 Ashes from EFR and CTF coal trials


Most of the particles in the coal-derived ash samples were found to be derived from a
single mineral, and showed the same transformations as the ash particles derived
from the pure mineral runs on the EFR. Some ash particles with intermediate
chemical compositions were observed; these particles have formed from the
association of two minerals in the parent coal particles. The proportion of
intermediate ash particles closely reflects the degree of mineral-mineral association
in the parent coal. In general, there were slightly more ash particles of intermediate
composition than there are mineral occurrences of intermediate composition,
because of agglomeration of two or more transforming mineral occurrences on the
surface of a burning char particle.

The size distributions of particles in the coal-derived ash samples were similar
to those of the mineral occurrences in the parent coals. Ash particles at both
ends of the size range are less abundant than coal mineral occurrences of
equivalent size, probably because of fragmentation of some larger minerals
and agglomeration of fine included mineral matter during combustion. The
finest particles in all the CTF and EFR ash samples consistently showed lower
SiO2 contents and higher Fe2O3 and CaO contents than the average for each
ash sample. The extent of these differences, and the relative changes in Fe2O3
and CaO contents, varied between coals. The variation in composition for the
17
finest ash particles is probably a direct reflection of the variation of mineral
proportions for the smallest mineral occurrences in the parent coals.

5.8 Deposits from EFR coal trials


In general, good agreement was observed between the bulk chemical compositions
of the deposits, the ash samples and the parent coals, and between the CTF and
EFR deposits from the same coal. Deposits from the PRB coals showed CaO
depletion, possibly due to the presence of calcium species in the gas phase that later
condensed on cooling ash particles but not on the deposit. A small but consistent
difference in behaviour between Al2O3 and SiO2 on deposition was observed. No
significant differences in microstructure between CTF and EFR deposits were
observed, although the CTF deposits were considerably thicker than those obtained
in the EFR.

Deposits showed a wide range in the degree of sintering, from dusty through sintered
to partially fused and almost fully fused ash. Deposit microstructures reflected these
variations. The extent of sintering correlated well with the chemical composition. The
EFR index, based on the chemical distribution of the deposit and measured by
CCSEM, estimated the potential of the deposit to consolidate. EFR index predictions
of slagging potential differ from those of the base/acid ratio, because the base/acid
ratio does not fully reflect the differing behaviours of CaO and Fe2O3 [Gibb (1986)].
Deposits from PRB 1- PRB 2 and PRB 1-Bailey coal blends had intermediate
properties to deposits from the individual coals, with no evidence of non-linear
behaviour.

5.9 Deposits from CTF coal trials


A well-established index for predicting the slagging propensity of coals with a
bituminous ash is based on ash basicity, defined as the ratio of the basic oxides
(Fe2O3+CaO+MgO+ K2O+Na2O) to the acidic oxides (SiO2+Al2O3+TiO2). This
parameter has been included in the table of coal and ash characteristics in Appendix
2. A linear relationship was found between the Jones Index, (see Table 5.1) and the
basicity of the ash, with the exception of two of the deposits, namely those from
Goedehoop and Harworth coals. Goedehoop gave a deposit that was less fused and
less sintered than might have been expected from the high CaO content. The ash
has the characteristics of a lignitic ash, defined as ∑CaO+MgO < Fe2O3, and would
not necessarily be expected to fit an index derived for bituminous ashes. PRB 1 gave
the most sintered ashes, the deposits being fully fused and glassy at 1200°C. This
behaviour is undoubtedly due to the high Na content of the coal. Na has long been
recognised as giving rise to sintered ashes in boilers burning lignites and sub-
bituminous coals. Na species are highly volatile at the combustion temperatures and
can be transferred to the aluminosilicate ash particles before and after deposition. Na
has a pronounced effect at lowering the viscosity of a silicate material and therefore
plays a key role in ash slagging and fouling.

PRB 2 was another PRB coal, although the Na content was not so high as in the
PRB 1. The PRB 2 deposits were well sintered, but not so fused as those from the
PRB 1 coal. Hunter Valley gave the most benign deposits, being no more than dusty
coverings of no mechanical strength. The ash has a SiO2 content of over 80% and a
low Al2O3 content. The mineral matter was almost entirely quartz and kaolinite, both
refractory materials.

18
Al2O3

Kaolinite

SiO2
CaO

Al2O3

Kaolinite+ 10wt% calcite

SiO2
CaO
Al2O3

Kaolinite+ 25wt% calcite

SiO2
Al2O3 CaO

Kaolinite + 50wt% calcite


SiO2
CaO

Al2O3

Calcite

SiO2
CaO

Fig. 5.1 CCSEM analyses of ash particles from kaolinite and calcite mixtures. During
the transformations the transfer of Ca species to the aluminosilicate ash particles is
evident, providing particles with a range of Ca-aluminosilicate compositions.

19
Harworth coal gave a deposit that was more sintered and fused than the base: acid
ratio would have predicted. The coal has relatively high levels of CaO, MgO and
Fe2O3. While each of these oxides has the ability to break the covalent bonding in
aluminosilicates, lowering the viscosity, it has been observed [Russell(2002)] that in
combination a certain synergy is exhibited giving a more powerful fluxing action than
a simple chemical analysis might suggest.

6 Mineral Transformations and Ash Formation Processes

6.1 Transformation from coal minerals to ashes


The impact of coal minerals on most aspects of plant operation can be described in
terms of the decomposition of the key mineral groups, i.e. the clays, quartz, sulphides
and carbonates, as listed in Table 6.1.

Mineral Chemical composition


Clays Kaolinite Al2Si2O5(OH)4
Illite Based on muscovite [KAl3SiO10(OH)2], but the
proportions of K, Al and Si may vary and these
elements may be partially replaced by Na, Ca, Mg
and Fe
Montmorillonite Like illite, but with Ca and Na more abundant than K
Quartz SiO2
Pyrite FeS2
Carbonates Calcite CaCO3
Dolomite CaMg(CO3)2
Ankerite Ca(Mg,Fe,Mn)(CO3)2
Siderite FeCO3
Minor Apatite Ca5(PO4)3(OH)
Crandallite CaAl3(PO4)(SO4)(OH)6
Feldspar Compositions between NaAlSi3O8 and CaAl2Si2O8
Garnet (Mg,Fe)3Al2Si3O12
Rutile TiO2

Table 6.1 Key mineral groups for the minerals found in the coals

6.2 Clay mineral transformations


Pulverised coal particles that contain the clay minerals tend to consist mostly of
mineral matter. Two main types of clays are found in bituminous coals: kaolinite and
illite. Kaolinite has a chemical composition that varies little from the formula

20
Al2Si2O5(OH)4., while the illites have a wide range of chemical compositions based on
muscovite [KAl3SiO10(OH)2], but in which the proportions of K, Al and Si may vary
and these elements may be partially replaced by Na, Ca, Mg and Fe. In lower rank
coals, montmorillonite is often more commonly found than illite. A wide range of
chemical compositions is also possible in montmorillonite, with Ca and Na being
more abundant than K. The proportion of kaolinite to illite is generally higher in
Southern Hemisphere coals, where the clays may be more widely dispersed through
the organic coal matrix and in individual pulverised coal particles. Clays melt to give
an aluminosilicate liquid that usually contains small crystals of mullite [Al6Si2O13].
Pure kaolinite melts more slowly than the other clay minerals, because of the more
restricted chemistry and the absence of cations such as K, Na and Ca. However, in a
seam of mixed clay minerals the combustion behaviour will be controlled by the local
average chemical composition, rather than by the properties of individual clay grains.
On cooling, liquid ash particles derived from clay minerals supercool to form an
aluminosilicate glass with the crystallisation of some mullite.

6.3 Quartz transformations


Quartz has a composition that differs little from that of pure SiO2. Like the clays,
quartz enters the coal formation process as a river-borne product of rock weathering.
Larger grains of quartz tend to form excluded particles in pulverised coal, while
smaller grains may be intimately mixed with the clay minerals. Quartz grains melt
more slowly than other minerals and large quartz grains often survive their passage
through the pulverised coal flame with evidence of only minor surface melting.

6.4 Pyrite transformations


Pyrite usually forms in coal after deposition; as clusters of crystals called ‘framboids’
formed under anaerobic conditions during the coalification process, or as veins
infilling seams and cracks following coal formation. Framboids, typically 10-20µm in
diameter, consist of spherical clusters of 1µm pyrite crystals and are usually found
within organic pulverised coal particles, although some individual pyrite crystals may
be released by milling. Pyrite from veins is generally more massive and occurs as
excluded pyrite particles with angular morphology. Small grains of pyrite may also be
present in seams of coal mineral matter, intimately associated with clay minerals and
quartz. During combustion, pyrite readily loses sulphur to form pyrrohtite [FeS1-x]
which then oxidises, passing through a molten iron oxy-sulphide stage before forming
magnetite [Fe3O4] or hematite [Fe2O3], [Groves et al (1987)].

6.5 Carbonate mineral transformations


The main carbonate mineral in Northern Hemisphere bituminous coal is calcite
[CaCO3], but dolomite [CaMg(CO3)2] and ankerite [Ca(Mg,Fe,Mn)(CO3)2] are
frequently associated with calcite. Siderite [FeCO3] is more common in Southern
Hemisphere coals, although the other carbonate minerals may also be present
[Bailey (1998)]. Carbonate minerals show some solid solution in each other, resulting
in a variation in composition. Carbonates generally formed in coal following the
coalification process, are typically found in veins. After milling, carbonate minerals
tend to form excluded particles in the pulverised coal. All carbonate minerals lose
CO2 during combustion, forming agglomerates of small oxide particles with
compositions reflecting the parent mineral [ten Brink (1996)].

21
6.6 Minor minerals
Several other minerals are found in many coals at low levels of abundance; feldspar,
garnet, rutile, apatite and crandallite. These minerals form ash particles with
distinctive chemical compositions, but do not affect power station operation or ash
properties.

6.7 Mineral associations


Multiple mineral occurrences present in single pulverised coal particles may form
separate ash particles during combustion, or they may combine to form a composite
ash particle. Since clay minerals make up a large proportion of coal mineral matter,
these composite particles usually have an aluminosilicate component. Associations
with quartz or pyrite give rise to silica-rich or iron-rich aluminosilicate ash particles.
Associations between clays and carbonate minerals are not common, because
carbonates tend not to occur as included mineral matter.

6.8 Mineral transformations – an overview


Pulverised coal particles entering the power station boiler contain a range of minerals
with a variety of mineral associations. The coal minerals also undergo a wide range
of transformations as described and reviewed by Wibberley and Wall (1982), Raask
(1985), Cunningham et al, (1992), Couch (1994), Wigley and Williamson (1995),
Bryers (1996), and Scott (1999):
• Minerals such as carbonates or hydroxides decompose at high temperatures
to form the respective oxides
• Calcite may fragment as it heats up rapidly
• Many minerals melt at flame temperatures
• Sulphide minerals loses sulphur and oxidise in the combustion environment
• Alkalis in transforming clay minerals or combusting macerals may vaporise,
and may later condense onto cooler ash or deposits
• Minerals that are associated in a coal particle, and deposited ash particles, will
chemically interact
• After deposition, aluminosilicate ash particles will sinter at rates depending on
the local deposit temperature and chemical composition
• A deposit matrix will often crystallise – the phases formed will depend on the
local chemical composition

Most ash particles follow the cooling combustion gas stream and leave the boiler
without further transformation. Ash particles that do deposit enter a different
environment in which further interactions and transformations occur. The mineral
transformations after deposition determine the effect of deposits on heat flow in the
boiler, the rates at which deposits consolidate and the effectiveness of sootblowing
operations.

22
7 Ash particle interactions and transformations

Most ash particles do not deposit – they follow the cooling gas stream and leave the
boiler without interacting with other ash particles or with the boiler itself. If ash
particles do impact on the boiler wall they may simply rebound, especially in the
cooler convective section of the boiler. However, ash particles that do deposit
experience a dramatic change in environment. The deposit will be cooler than the
arriving ash particle, because of the heat extraction through the underlying boiler
tubes, but the time scale for interaction in a deposit is much greater than the transit
time of mineral matter through a boiler. Processes that are ongoing in the deposited
ash, such as melting and crystallisation, may continue in the deposit. The deposited
ash particle will also be in close contact with other ash material providing the
opportunity for an additional range of transformations.

7.1 Transformations prior to interaction with other ash particles


Fusion of larger quartz grains, homogenisation of particles derived from multiple
mineral precursors, and growth of existing crystalline phases such as mullite in
aluminosilicate melts, are processes that can continue in an ash particle after
deposition. The rate of these processes will usually decrease, because of the cooling
of ash particles on deposition, but the longer time available at approximately constant
temperature in the deposit may allow further fusion and crystallisation to occur.
Fouling deposits, in the convective section of the boiler, are too cool for ash particles
to show significant further transformation. In slagging deposits, the local temperature
of the deposit is critical in determining rates of transformation.

7.2 Quartz
On deposition, quartz starts to dissolve slowly in the aluminosilicate matrix of a
deposit. Although smaller quartz grains fuse and are assimilated in deposits more
readily, larger quartz grains may persist for longer than the interval between
sootblowing in power station boilers or sample collection in combustion test rigs. The
presence of solid, untransformed, quartz in the deposit may cause an increase in
viscosity, reducing the rate of sintering and enhancing deposit removability. In
dissolving, the quartz will cause the aluminosilicate liquid to become locally silica-
rich, with increased viscosity and reduced diffusion rates. The presence of
undissolved quartz will certainly mean that the chemical composition of the deposit
matrix differs from the bulk deposit chemistry.

7.3 Clays
Liquid clay-derived ash particles, mainly aluminosilicate in composition, will sinter to
form dense homogeneous materials in deposits at a sufficiently high temperature
(Nowok et al, 1990). Compaction and a reduction in porosity changes the properties
of the deposit. Denser deposits conduct heat more efficiently, reducing both the
fireside temperature of the deposit and the effect of the deposit on heat transfer out
of the boiler. However, the deposit is more strongly bonded to the boiler tubes and
removal by sootblowing becomes more difficult.

7.4 Pyrite
Solid iron oxide particles, derived from pyrite, are physically and chemically very
different from the glassy aluminosilicate matrix of a typical deposit. Even in slagging
deposits, rates of dissolution are often slow, with spheres of iron oxide persisting for
days or longer and behaving like residual quartz.

23
7.5 Particle interaction
Ash particles interact with each other by a process of diffusion and viscous flow. The
ash particle interactions are driven by a decrease in the free energy of the system
when gas-solid interfaces are replaced by solid-solid interfaces. The sintering
process leads towards a more uniform chemical composition that is related to coal
ash composition, though this need not be the same as the bulk coal ash composition,
since deposition processes are usually partially selective. The rate of sintering will be
strongly dependent on the local temperature of the deposit, and on the chemical
compositions of the ash particles involved. Aluminosilicate melts with high
concentrations of network modifiers such as Ca2+, Fe2+ and K+ will sinter more rapidly
than typical illite-derived ash material, while pure aluminosilicate material derived
from kaolinite will sinter slightly more slowly. Progress towards a uniform bulk
chemical composition may be hindered by the need to dissolve residual quartz grains
or crystalline phases.

The sintering and consolidation of deposits determines the effect of the deposit on
heat transfer through boiler walls and also the ease of deposit removal by
sootblowing (Wain, 1991). Post-deposition transformations are mainly driven by the
interaction between deposited ash particles; deposits are the only stage in the
passage of mineral matter through a power station at which such widespread
interaction between different mineral matter components is possible.

7.6 Crystallisation
Crystalline phases will usually grow from the aluminosilicate melt at the temperatures
at which slagging deposits form,. Mullite is frequently the stable phase in
aluminosilicate melts at deposit temperatures. Anorthite [CaAl2Si2O8] may also
crystallise if the concentration of CaO in the bulk coal ash is high, or if deposition
processes have preferentially selected lime-rich ash particles because of their lower
viscosity. Gehlenite [Ca2Al2SiO7] and anorthite, but not mullite, crystallise from even
more lime-rich deposits such as those from coals in which montmorillonite or
crandallite is abundant. The magnetite [Fe3O4] that forms as pyrite or siderite
transforms is not stable in the aluminosilicate melts that form boiler deposits, but
hercynite (a spinel [Fe(Fe,Al)2O4] in which Al3+ has replaced Fe3+ in the magnetite
structure) may be stable in iron-rich aluminosilicate deposits.

7.7 Homogenisation
The chemical homogenisation of a deposit by sintering, diffusion, dissolution and the
growth of locally stable crystalline phases may interact in a complex manner. When
magnetite dissolves, for example, the surrounding aluminosilicate melt becomes
enriched in iron, and hercynite may crystallise around the dissolving magnetite
particles. Hercynite will persist until diffusion of iron away from the original magnetite
lowers the local iron concentration sufficiently to allow the hercynite to begin to
dissolve. Given the various obstacles to deposit homogenisation, it is not surprising
that only the hottest boiler deposits approach phase equilibrium on a macro-scale.
Unstable crystalline phases and incomplete chemical homogenisation usually persist
in slags from regions higher in the radiant section of the boiler or from regions of
deposits close to the boiler tubes. Deposits from cooler regions of the boiler or from
immediately adjacent to boiler tubes may show little transformation beyond the first
stages of sintering (Wigley et al, 1990).

24
7.8 Condensation
As previously stated, the surfaces of slagging deposits are usually too hot for
condensation of vaporised elements to occur. However, significant condensation may
occur on fouling deposits in the convective zones of the boiler. Condensation of alkali
sulphates or chlorides may bind together deposit ash particles that do not themselves
interact, forming a dense deposit that obstructs gas flow and can be difficult to
remove. Under the right conditions, lime and unreacted calcite particles may
sulphate, sintering and bonding a fouling deposit.

7.9 Rates of transformation


Typical times for mineral transformation during pulverised coal combustion in utility
boilers are summarised in Table 7.1. Individual mineral occurrences show a range of
transformation rates, depending on mineral size and associations and on mineral
environment in the boiler. With the exception of the melting of excluded quartz,
mineral transformations during pulverised coal combustion typically occur with
timescales similar to or shorter than the time taken for mineral matter to pass through
the radiant zone of a utility boiler.

Mineral Transformation Comparative


time
Pyrite Sulphur loss to form FeS Much shorter
All clays Glass formation from melt Much shorter
Siderite and ankerite Decarbonation Shorter
Pyrite Oxidation of FeS Shorter
Calcite and dolomite Decarbonation and Slightly shorter
partial transfer of Ca to clays
Illite and Fusion and partial Slightly shorter
montmorillonite devolatilisation
Kaolinite Fusion Similar
Quartz Fusion Much longer

Table 7.1 Typical times for major mineral transformations during coal combustion,
in comparison with the transit time of mineral matter through the radiant zone of a
power station boiler. Mineral size, association and path through the boiler have
strong effects on transformation times.

When ash particles deposit within the boiler, the rate at which further transformations
occur strongly depends on the local deposit temperature. Typical times for mineral
transformation following deposition, based on a location within a pendant
superheater deposit, are summarised in Table 7.2, in comparison with a typical 8-
hour sootblowing cycle. In contrast to mineral transformations during combustion, the
transformations that occur following deposition are considerably slower, taking hours
or days and frequently never reaching completion.

25
Mineral origin Transformation Comparative time
Clays Sintering and fusion Slightly longer
All minerals Crystallisation from melt Longer
Pyrite Dissolution of iron oxide Much longer
Quartz Dissolution Much longer

Table 7.2 Typical times of major mineral transformations after deposition, based on
conditions at a pendant superheater deposit and in comparison with an 8-hour
sootblowing cycle. Local deposit temperature and chemical composition have strong
effects on transformation times.

7.10 Summary
In a power station boiler, coal minerals transform in response to the high
temperatures reached in pulverised coal combustion. Clay minerals and quartz fully
or partially fuse to form a viscous melt. Pyrite loses sulphur and oxidises through a
liquid Fe-S-O stage to solid iron oxide. Carbonate minerals lose CO2 to give mostly
solid oxide products. In general, the ash particles produced during mineral
transformations are derived from a single mineral and display limited chemical
compositions. Mineral occurrences that contain two intimately associated minerals, or
different mineral occurrences that coalesce during combustion, will invariably interact
to produce aluminosilicate melt particles with a wider range of chemical
compositions. Organically-bound alkalis may volatilise and condense on ash particles
in the cooler convective region of the boiler. Calcium, from calcite, may interact with
transforming clay minerals to produce calcium-rich aluminosilicate ash particles.
There is, therefore, a direct correlation between the physical and chemical properties
of the coal minerals and of the ash particles produced during combustion; the mineral
transformations affect the subsequent behaviour of the ash particles, including their
deposition propensity.

Ash particles that do deposit encounter other ash particles over a much longer time
period than the boiler transit time. In slagging deposits the deposited ash particles
may interact at the elevated temperatures, while in fouling deposits particles will
experience alkali sulphate and chloride condensation. Slags sinter by a process of
diffusion and viscous flow – a wide range of aluminosilicate chemical compositions is
possible, and crystallisation may occur.

Although transformations that are on-going in the deposited ash particles may
continue, most of the transformations in slagging deposits are a consequence of
particle interaction at elevated temperatures. Since the temperature range between
and within deposits is very wide, the range of possible interactions and their level of
completion is very variable. Post-deposition transformations determine the
microstructure of the deposit, and therefore the thermal properties and potential for
removal, all of which affect boiler operation.

26
8. Implications for Power Station Performance
8.1 Types of boiler ash
The most abundant non-combustible species in coal are those of the aluminosilicate
clay minerals, together with quartz. They account for between 60% and 90% of the
total mineral matter in coal. The most common species of clay minerals are
muscovite-illites (potassium aluminosilicates), kaolinites (aluminosilicates) and mixed
layer illite-montmorillonites of variable composition. One other non-silicate mineral,
found only in relatively small amounts in coal, can have a significant influence on
boiler plant from a depositional and plant maintenance standpoint and this is pyrite
(FeS2). This mineral is hard and very abrasive, it decomposes and fuses at a
relatively low temperature, and also emits sulphur-containing gases when the coal is
burned.
Ash is produced in two forms in coal fired pulverised fuel boilers, namely furnace
bottom ash (fba) and pulverised fuel ash (pfa). A 500 MWe coal fired unit will typically
produce 20 tonnes of ash per hour when operating at full output. Furnace bottom ash
(fba) comprises large pieces of fused material that have formed in the boiler furnace
in close proximity to the combustion process, plus deposits of slag that have formed
on the walls of the furnace, have become detached, fall to the bottom of the boiler
into the main ash hopper from where it is periodically removed. Fba typically makes
up between 5 and 15% of the total ash produced in the boiler. Pulverised fuel ash
(pfa) is by far the most predominant form of ash produced in coal fired pulverised
boilers and typically represents 85% to 95% of the total ash formed. Pfa is carried out
of the boiler furnace in suspension form by the flue gas and comprises finely divided
ash particles (and unburned carbon char) that are removed from the gas stream by
an electrostatic precipitator (ESP).

8.2 Washing and handling of coal


Coal, as mined, has a specific gravity ranging from 1.2 to 1.7 depending upon its
rank, the moisture content and the nature and amount of the mineral matter that it
contains. A ‘pure’ clean bituminous coal would typically have a density of 1.3. The
shale associated with the coal might have a specific gravity of about 2.5 and the
pyrite present would have a specific gravity of 4.9. Table 8.1 illustrates the variation
in specific gravity of common constituents of coal.
Typical limits for ash content of internationally traded steam coal are <14%, which
usually introduces the need for some degree of coal preparation to reduce the run of
mine ash to an economical level. Coal is generally screened into size fractions and
then washed, often using density separation processes. The performance of
conventional cleaning technology

Component Specific Gravity


Coal macerals 1.15 -1.5
Aluminosilicates (shales, 2.0 -2.6
clay and sandstones)
Gypsum 2.3
Calcite 2.7
Pyrite 4.8 - 5.0

Table 8.1. Specific gravity of common constituents in coal

27
depends on the mineral type and distribution. Generally, a large proportion of the
inorganic matter associated with the coal is in the form of particles such as quartz,
pyrite and calcite. The efficiency to which coal cleaning removes these particles,
depends to a large extent on the degree of crushing that has taken place in order to
liberate this material as separate particles. Further crushing is sometimes
undertaken after a first-stage density separation in order to further liberate the
mineral matter for subsequent removal.

8.3 Abrasive and erosive materials in coal


The sliding of hard particles over a surface or between two surfaces causes abrasive
wear. In full-scale plant the main areas of concern are in coal handling, crushing and
milling equipment. Erosive wear is associated with the impact of hard particles,
carried in a fluid at significant impact velocities, on metallic and other surfaces. The
key areas of concern at full scale are the internal surfaces of coalbunkers and
chutes, mills, exhauster fans, PF pipework/riffleboxes and the burner internals.
Erosive wear can have an effect on structural integrity, performance and
maintenance requirements. Erosive wear of internal boiler tubes and other surfaces
in the convective section of the furnace can also be of concern.

In the case of both mechanisms, the key uncertainty is the abrasiveness or


erosiveness of the coal or ash material. This is dependent on the ash content of the
coal and on the concentrations and size distributions of the harder constituents of the
ash. Ash-free coal would not cause significant abrasion or erosion in coal plant; the
wear damage is caused chiefly by the mineral impurities in coal.

Coals contain a wide range of minerals, but it has been shown by Wells et al
[2004(a,b)] that quartz and pyrite, minerals that are harder than steel, are the main
components of coal that are responsible for wear and abrasion. The other hard
minerals, orthoclase, kyanite, topaz and alumina are generally present in trace
quantities and have a minimal effect on overall wear and abrasion. Quartz is
particularly hard and shatters to produce sharp-edged fragments. Quartz also tends
to occur in the form of comparatively large particles of ‘free’ mineral matter, whereas
much of the pyrite is dispersed in the coal substance and clay sediment. Other
minerals such as the clays, carbonates, sulphates and phosphate minerals are
relatively soft and do not cause any significant erosion wear. Table 8.2 below gives
the concentration and hardness of different mineral species in a typical bituminous
coal.

The DTI project Cleaner Coal Project 218 has investigated coal mineral matter in
some detail, characterising the size, shape and degree of exclusion of the mineral
matter in a range of power station coals by Scanning Electron Microscopy (SEM).
The results showed that quartz is effectively twice as abrasive as pyrite on a wt%
addition basis of the free minerals to a UK coal, which is in agreement with
Raask[1985]. The grain size of angular quartz was found to be critical in determining
erosion and abrasiveness, up to a critical size of approximately 100µm. Above this
size, the abrasiveness of the coal/mineral addition mixture becomes independent of
grain size. The particle shape was shown to be a key factor in determining
abrasiveness with angular particles causing more abrasion than rounded quartz
particles. The effect of adding both angular and rounded particles was not
cumulative suggesting that the different abrasion mechanisms of angular and
rounded particles are not mutually exclusive.

28
Constituent Approximate Mohs Vickers Hardness
wt % in coal Hardness (kg/mm2)
number
coal substance 85 1.5-2.5 10-70
quartz 1.6 7 1200-1300
pyrite 1.5 6-7 1100-1300
silicate minerals
kaolin 3 2-2.5 30-40
illite 3 2-2.5 20-35
muscovite 3 2-2.5 40-80
orthoclase <0.1 6 700-800
kyanite <0.1 6-7 500-2150
topaz <0.1 8 1500-1700
carbonates and other minerals
calcite 0.5 3 130-170
magnesite 0.1 4 370-520
siderite 0.2 4 370-430
alumina rare 9 1200

Table 8.2 Concentration and hardness of mineral species in bituminous coals

8.4 Coal Milling


The grindability and density of minerals in coal determines their ultimate size in
pulverised coal after milling. Aluminosilicate clay minerals, which frequently occur in
coal as dispersions of sediment, appear to concentrate in the finest fractions, i.e.
below 20 microns. Silica, in the form of quartz, tends to appear in a size range of 20
to 75 microns, although some larger particles may be found in PF. Pyrite tends to
concentrate in a size range of 10 to 45 microns. This is despite pyrite being difficult to
grind. The reason is that the air classification systems used on coal mills are unable
to remove the dense pyrite particles from the grinding table until they become very
finely divided.

Strict controls on NOx limits and dust emissions from coal-fired power stations means
that optimising electrostatic precipitator performance and maintaining low levels of
carbon in ash whilst achieving low excess air levels are of critical importance to
utilities. It is widely recognised that the degree of carbon burnout achieved in a
particular furnace configuration is directly related to the size distribution of the PF
supplied to the burners.

It is to be expected that the particulate mineral matter in a typical PF would have a


distribution spectrum weighted towards the smaller sizes compared with that of the
coal particles for two main reasons. Firstly, coals contain a proportion of the total
mineral matter in the form of fine clay sediment particles of 0.1 to 10µm in size, which
may be liberated as individual particles. Secondly, the density of coal is significantly
29
less than that of mineral matter, resulting in them being retained longer within the mill
and ground finer than the coal before they can pass through the classifier. The extent
to which coal milling releases minerals as discrete particles depends on the origin of
the coal and the degree of washing the coal has undergone.

A fineness specification of 65±3% less than 75µm and <1% greater than 300µm was
specified for all the coals delivered from the grinders (Durrans) for testing on the
E.ON combustion test facility (CTF). However, there were some significant
differences in fineness and particle size distribution. Three of the coals, Talcher,
Bowen basin and PRB 1, were delivered as a considerably finer PF compared to the
remaining coals, which were very consistent in particle size.

8.5 Burnout and combustion efficiency


All combustion testing was undertaken on the E.ON 1MW CTF. It should be noted
that the CTF is a single burner facility in which the flow of fuel and air can be
accurately controlled. Tramp air is minimal. In full-scale utility boilers, there are often
variations in fuel and airflow to different burners and tramp air in-leakage resulting in
fuel-rich zones having a detrimental affect on combustion efficiency. As a guide, it is
assumed that LOI figures obtained at 1% excess O2 on the CTF would be more
representative of the values obtained at 3% O2 on full-scale plant. Further details of
the combustion trials have been given in Section 4.6 of this report. Burnout
predictions were made by Nottingham University by determining the proportions of
“unreactive” macerals and coal rank for the coals, the Reactivity Assessment
Programme (RAP), as described in Section 4.

LOI has been used to represent the proportion of carbonaceous material left unburnt
in the ash, but with ash contents of the parent coals varying from 8-38%, the
combustible loss (CL) gives a better assessment of the overall combustion efficiency.
LOI and CL for each coal were given in Table 4.1. Apparent discrepancies in LOI
and CL are due to the different ash contents of the coals.

The highest combustible loss values were associated with coals having the highest
% “Unreactive” content. These included coals such as Goedehoop, Hunter Valley
and Bowen Basin. Higher levels of unburnt carbon are a characteristic of South
African coals, such as Goedehoop, due to the relatively high levels of inertinite
(>50%) and lower levels of more reactive vitrinite. The low combustible loss values
were associated with the lower rank coals such as PRB 1, PRB 2 and Talcher.
Experience of burning Prodeco at full-scale is in agreement with these rig-scale
results – burnout tends to be poor despite a high vitrinite and low inertinite content,
one possible reason being that at least two different seams, each with a different
combustibility, are extracted from the same mine and then blended.

8.6 NOx Emissions


NOx emissions will be influenced by a number of parameters, including fuel
composition and plant operations. The primary fuel characteristics that affect NOx
emissions are the nitrogen, volatile matter and moisture content. The highest NO
emissions were obtained from the Bowen Basin, Goedehoop and to some extent the
Hunter Valley coals. This is a consequence of the relatively high DAF nitrogen
content of the Goedehoop and Hunter Valley (2.03% and 1.86% respectively) and
the relatively high Fuel Ratio for all three coals (Goedehoop 2.31, Bowen Basin 2.15
Hunter Valley 1.83). Previous experience at full and rig-scale has shown that there is

30
a reasonable correlation between NO emissions at a given oxygen level and the
product of Fuel Ratio and DAF N%.

It has been reported that at full scale, the Bowen Basin coal might not give such high
NOx emissions as might be expected. Although the Bowen Basin gave relatively
high NO emissions during the testing on the CTF, it is a relatively wet coal by
bituminous standards (18-20% moisture). As all the coals for the CTF tests were
ground by an external contractor (Durrans), they were all received in a relatively dry
condition as PF (typically <10%). It is possible that the additional moisture, in a high
moisture coal such as Bowen Basin, released during the milling process would
induce a degree of flame cooling and this might result in lower NO emissions than
those seen during the CTF testing. This effect could be similar for all coals, but
perhaps most relevant for Bowen Basin, which although bituminous, contains
moisture levels approaching those of sub-bituminous coals.

PRB 1 and PRB 2 coals gave the lowest NO emissions, as might be expected given
their analysis. High DAF VM, low DAF nitrogen and high moisture (even after
grinding) all contribute to lowering NO emissions. The blend of PRB 1 and Bailey
gave broadly additive behaviour in terms of NOx emissions, compared to the two
individual fuels.

8.7 Electrostatic precipitation of fly ash


Approximately 85-95% of the ash formed in the coal combustion process in a pf
power station boiler remains in suspension in the flue gases as they exit the
combustion chamber. The commonest method of dust collection is the use of
electrostatic precipitators (ESP). An ESP electrically charges the ash particles in the
flue gas to allow collection and removal. The process consists of four stages, (i)
charging, (ii) collection, (iii) rapping and (iv) ash removal.
The electrical conductivity of the fly ash and the dielectric strength of the bulk ash are
two properties important to the electrostatic collection process. For effective
operation, a small but definite electric current, in the form of charges carried by gas
ions and particles, flows between the high-tension discharge electrodes and the
collecting electrodes. This current must pass through the layers of collected ash,
which normally coat the plates. The ash, therefore, must be able to conduct the ionic
current to the grounded metal surfaces of the plates. The electrical conductivity
required is very small, about 1010ohm.cm-1. In practice, it is usually more convenient
to express this in terms of resistivity, which is the inverse of conductivity.

In general, the coal and ash components that produce high resisitivities are SiO2 and
Al2O3, both excellent insulators in their own right, whereas those that reduce the
resisitivity are sulphur and Na. Although during combustion most of the pyritic
sulphur in the coal ultimately forms sulphur dioxide, a small percentage is converted
to sulphur trioxide, can significantly affecting the resisitivity. In a boiler flue gas, there
is always moisture present that can react with the sulphur trioxide to form condensed
sulphuric acid. This then uses the surface of the ash particles for condensation,
thereby reducing their resistivity, as conduction is now able to proceed through this
absorbed or adsorbed acidic layer. The extent of surface conditioning is dependent
on the amount of sulphur trioxide present, the gas temperature and hence the vapour
pressure. For coals containing in excess of 1.5% sulphur, there is sufficient natural
surface conditioning at normal plant operating temperatures of around 130ºC to give
acceptable values of resistivity for effective electrostatic precipitation. The presence
31
of Na ions causes the ash particle to behave very differently. Instead of surface
conditioning, the Na ions act as charge carriers, so high resistivity effects are
negated to an extent. Potassium does not appear to behave in the same manner to
Na. Ca and Mg in coal are able to produce sulphates that are not effective
conductors, thus the addition of sulphur trioxide will not be effective. [Parker (1997)].

Although not an ash particle, the presence of increased amounts of unburnt carbon,
as a result of fitting low NOx burners, can adversely affect ESP performance. The
lower resistivity of carbonaceous particles means that they are readily re-entrained
into the flue gas stream and subsequently emitted, [Wu (2000)]. The use of additives
to combat the effects of increased unburnt carbon in PFA is an issue that seems to
require further investigation.

In addition to increased unburned carbon, low NOx burners produce coarser and
more irregularly shaped fly ash particles than with unstaged combustion. Increasing
particle size would be expected to improve ESP collection efficiency, however, more
irregular particles will experience greater drag forces that decrease the particle
migration velocity and lowers the ESP efficiency [Wu (2000)].
There is also some evidence to indicate that the fitting of low NOx burners increases
the split of fly ash to bottom ash from about 70:30 for unstaged combustion to 85:15
or higher with low NOx burners.

8.8 Ash resistivity


As coal burns the high temperature promotes decomposition of and chemical
interaction between mineral particles that in turn determines their resistivities. Parker
[1997] has studied the composition of ashes with their collectibility using an ESP and
provided the data shown in Table 8.3

Ash type Resistivity (ohm.cm)

Conductive ash 104 to 108


Normal resistivity ash 108 to 1010
Moderate resistivity ash 1010 to 1011
High resistivity ash 1011 to 1013

Table 8.3 Resistivity of pulverised fuel ashes

Experience has shown that coal ashes with resistivities greater than 5 x 1011 ohm.cm
are difficult to collect. Low resistivity ash (less than 108 ohm.cm) tends to suffer from
excessive re-entrainment, but such low resistivities are rarely encountered. A range
of ash resistivities between 109 and 5 x 1011 ohm.cm is likely to produce satisfactory
and predictable ESP operation.

Dalmon and Raask [1972] found that the hydrated silicate materials normally present
in coal, i.e. kaolin, illite and chlorite, had resistivities in the range 1010 to 1011 ohm.cm.
After heating to around 800OC the resistivity increased by one to three orders of
magnitude because of the loss of chemically bound water in the original crystal
structure. Unhydrated silicates, quartz, kyanite, muscovite and albite had high

32
resistivities of 1013 ohm.cm. and above. Quartz, kaolin, kyanite and muscovite
appeared to retain their resistivities up to the PF flame temperature. The resistivities
of silicate particles containing iron and sodium decreased by three orders of
magnitude after fusion of the particles.

8.9 Conclusions
Although there are some changes to the resistivity of the coal minerals as the coal
burns, the effects are relatively small and are generally well understood. This means
that electrostatic precipitators can be designed to collect most of the particulate
material in the flue gas stream based on a knowledge of the mineral matter in the
coal and the chemistry of the coal ash.

In most cases, deviations from ESP design ash characteristics result in the
production of a higher resistivity ash associated with the need to burn low sulphur
coals for environmental reasons. Many of the low sulphur coals originate from the
Southern Hemisphere and contain highly refractory ashes of high resistivity.
Conditioning agents such as sulphur trioxide and sodium compounds can alleviate
such problems. The effect of unburnt carbon in fly ash on ESP performance is well
known and may require the development of new conditioning agents to restore ESP
performance.

9. Ash sales and ash disposal


Currently, coal fired power stations in the UK generate in excess of 5 million tonnes
of ash per year of which approximately 50% arises from the use of indigenous coal
and the balance arises primarily from the combustion of Southern Hemisphere coals.
The mineral content of coal is transformed into two types of ash during the
combustion process, furnace bottom ash (fba) and pulverised fuel ash (pfa) each of
which has significantly different impacts on power station operation, ash sales and
disposal.

9.1 Furnace bottom ash


Furnace bottom ash (fba) comprises large pieces of fused slag that have formed in
the boiler furnace in close proximity to the combustion process plus deposits of slag
that have formed on the walls of the furnace. The fba falls to the bottom of the boiler
into the main ash hopper and is periodically removed. Fba typically makes up
between 5 and 15% of the total ash produced in the boiler. The removal frequency is
dependent on the rate of formation of fba but is typically once or twice a day.
Normally fba is water jetted out of the ash hopper to the crushers and onwards into
storage bunkers from which the water is drained and recycled and the fba removed
by a grab. There is a ready market for fba in the UK for a range of activities including
block making, granular fill and other lightweight granular aggregate uses. As fba is
effectively a by-product with the whole production being sold there are effectively no
disposal issues.
There are few criteria in the sales specification for fba (other than size and moisture
content both of which are post boiler processes). The key factor that can be affected
by the combustion process and the fuel type is the quantity produced. The amount of
fba can vary from 5% to 15% of the ash entering the boiler. The impact of the
variation can be seen to be far more marked in terms of fba production than pfa. A
three fold increase in fba production would have a major impact on the main plant
operations in terms of “ashing out” frequency and could even lead to reductions in
33
load output on the main boiler if the fba could not be removed quickly enough. The
key drivers for the formation of fba are the ash content in the coal and the slagging
propensity of the ash [Couch (1994)].

The relationship between ash fusion temperature and combustion temperature in the
furnace also acts as an indicator of slagging propensity and fba formation. There is
now good evidence of a number of boilers where the rate of fba formation has been
significantly reduced with the lower flame temperatures following the installation of
low NOx burners.

9.2 Pulverised fuel ash


Pulverised fuel ash (pfa) is by far the most predominant form of ash produced in coal
fired pulverised boilers and typically represents 85% to 95% of the total ash formed.
A 500 MWe boiler will typically produce 20 tonnes of pfa per hour and for many
years, disposal of pfa has been a key issue for the power industry. Many applications
for its use have been developed for high quality pfa, but overall there is still greater
pfa production than a market place for its use.

9.3 Uses of pulverised fuel ash


Fresh pfa removed from ESP has the most value as it has good pozzalanic and
physical properties that make the product attractive to the cement and allied
industries. In order to facilitate the handling of pfa it is conditioned with
approximately 10% water to reduce the dusty properties and allow transport etc to
the disposal or storage areas. The addition of moisture reduces the pozzalanic
properties of the ash and hence renders the material less attractive for the cement
replacement market. Ashes with a high CaO content tend to agglomerate when
conditioned. Any pfa that cannot be sold or dry stored is normally slurried and then
used for land reclamation in old sand and gravel workings etc or disposed of in
purpose made disposal lagoons.

There are a number of profitable applications for this pfa, however, which include
lime/ash mixtures for a range of civil engineering application as well as a bulk fill
material for road-works, land reclamation etc. The preferred downstream use for pfa
by the power industry is as an addition in concrete, primarily due to the revenue that
can be accrued from the relatively high price that pfa can achieve in this market.

9.4 Specifications for use of PFA as a cement replacement material


The physical and chemical properties of pfa for this market are closely controlled by
standards. In the UK, BS 3892 Parts 1 and 2 have set the specification for many
years and it is expected that this standard and its European counterparts will be
harmonised in the form of a new standard (EN 450) which it is hoped will be agreed
during 2004.

Both BS 3892 and the current EN 450 place limitations on chloride, sulphur and
calcium oxide contents, fineness, moisture and loss on ignition contents of the pfa.
Details of these standards are contained in the UKQAA Technical Data Sheet
1.1[2000], One of the key differences between the two standards is the basis by
which the calcium oxide content of the pfa is measured. In BS 3892, the limit is
based on 10% of total CaO in the ash whereas in EN 450 it is based on 2.5% of free
CaO in the ash. The BS was formulated when indigenous coals formed the main

34
supply to UK power station and there are virtually no deep mined coals that would
give ashes that would approach this CaO limit.

Mainland European coal fired power stations have traditionally used significant
quantities of international traded coals many of which come from Southern
Hemisphere sources. Many of these coals, particularly from the Southern
Hemisphere have CaO contents well in excess of the levels found in UK coals.
Increasing amounts of Southern Hemisphere coal are now being used in UK power
stations due to a number of reasons, including price, availability and lower sulphur
content and as a result there levels of total CaO in ash are rising. Thus, there is the
potential for a reduction in the usage of pfa in concrete additions if the total CaO
parameter remains in use. However, there is as yet no evidence that high CaO ash
coals are unsuited for use as a cement replacement in the UK.

A large number of ash and deposit samples have been collected from CTF and EFR
trials as part of this project and total CaO concentrations have been determined for
these samples. Free lime was not directly measured for the ashes and deposits.
However, as part of the CCSEM characterisation of the chemical variability of the
samples, the proportion of ‘Ca-rich’ ash particles or deposit point analyses has been
determined. The ‘Ca-rich’ chemical type is defined by the CaO content making up
more than 90wt% of the sum of (CaO + Fe2O3 + Al2O3 + SiO2) contents. Lime (CaO),
calcite (CaCO3), calcium sulphate and calcium phosphate all report in the Ca-rich
chemical type.

Although the CCSEM measurement of Ca-rich abundance is not the same as the BS
measurement of free lime, there is probably a strong correlation. Lime is associated
with aluminosilicate glass in all the other forms of Ca-bearing material in coal ashes
and deposits. Only the ‘Ca-rich’ material is sufficiently concentrated to be a possible
source of free lime, although not all of the ‘Ca-rich’ material need be ‘free’. The
abundance of Ca-rich ash particles has been plotted against total CaO concentration,
determined by CCSEM, for CTF and EFR ash samples in Figure 9.1 This shows that
the abundance of Ca-rich remains below 2.5wt%, and frequently below 1wt%, for all
the ash samples with total CaO contents below 15wt%. However, the two ash
samples from the Powder River Basin coals gave high total CaO contents (above
20wt%) and the ashes that also contained significant amounts of ‘Ca-rich’ material.

Similarly, a CCSEM characterisation of the CTF and EFR deposits showed only one
deposit to have a Ca-rich content above 2wt% (~ 2.5wt%), see Fig 9.2. Based on
these analyses, it appears that significant “free lime” is only likely to occur in boiler fly
ashes with very high total CaO contents, while boiler bottom ash is unlikely to contain
any significant “free lime”, whatever the CaO content.

The project has demonstrated that the relationship between total CaO and free CaO
in ash is non-linear and it would appear that the BS3892 parameters are suited to UK
indigenous coals whilst the EN 450 parameters are applicable to a much wider range
of coals. It is important to note that this wider range of coals is now more reflective of
the basket of coal used in UK power stations.

Other uses of pfa are mainly for conditioned rather than fresh material and include
lime/ash bound granular materials which have been used for road and airfield
pavements, sub-base and road base and trench mixes. In these applications there

35
12

10

8
Ca-rich (wt%)

0
0 5 10 15 20 25
Total CaO (wt%)

Fig 9.1 CCSEM analyses of CTF and EFR ashes for all 23 coals, showing only three
coals with more than 2.5% of Ca-rich particles.

12

10

8
Ca-rich (wt%)

0
0 5 10 15 20 25
Total CaO (wt%)

Fig 9.2 Abundance of 'Ca-rich’ point analyses plotted against bulk (total) CaO
concentration, as determined by CCSEM of both CTF and EFR deposits.

are none of the strict specification criteria of the concrete addition applications
applied to the pfa. These applications also require the presence of typically 3% of
free lime. As shown in this project, very few coal ashes actually contain this level of
free CaO, and it is necessary to provide this as an addition to the pfa for the specific
application.

The free CaO content again can affect ash disposal into lagoons. It has been noted
in a number of instances that the pH of the lagoon water and run off can rise
significantly when coals with high total CaO contents are used on a long-term basis.
It has, in a number of instances, been necessary to take remedial actions such as
coal blending or neutralisation of lagoon and run off water.

36
10. Conclusions
This project was undertaken to give the coal-fired power generators a greater insight
into the combustion behaviour of a wide range of coals that are now available as
world-traded commodities. Many of these coals have an ash chemistry that can be
quite different to that of UK coals, a consequence of the varied geographical origins
and a different mineral assemblage. Emphasis has been placed on obtaining an
improved understanding of the mineral matter transformations, mineral interactions,
the formation of ash particles and boiler deposits. The effects of mineral
transformations on power station performance have also been considered.

1. The combustion behaviour of nine coals and two blends has been successfully
studied using the E.ON 1MW Combustion Test Facility (CTF) and the Imperial
College Entrained Flow Reactor (EFR). The coals came from North and South
America, Australia, South Africa and India. A further fourteen coals, including coals
from Indonesia, China, Russia have been studied using the EFR. The mineral matter
in all the coals and the nature of the ash and deposits generated during the CTF and
the EFR trials has been characterised using the CCSEM and other facilities facilities
at Imperial College.

2. Combustible loss values for ash samples taken at 1% excess oxygen in the CTF
were approximately 1%, indicating relatively good combustion performance for most
coals. Goedehoop, a South African coal with a high inertinite content showed poor
burnout performance, while the two low rank coals from the US PRB gave extremely
low combustion loss values (<0.5%). The performance of the coal blends appeared
to follow an additive behaviour.

3. Petrographic analysis of the coals and chars was used to predict the burnout
properties of the coals and chars. In general a good correlation was obtained
between the predicted and measured combustion loss. The two coals that did not
give the predicted burnout were Goedehoop and Prodeco.

4. NO production was highest from those coals with a high Fuel Ratio and a high fuel
N content. NO emissions were lowest from the low rank, high volatile matter US
Powder River Basin coals. The blended coals gave NO emissions broadly in line with
an additive relationship.

5. The ash deposits collected on ceramic deposition probes ranged from dusty
deposits with no mechanical strength to highly fused, strongly bonded slags of low
porosity. The highest slagging propensity was shown by PRB 1, a high Na PRB coal,
while the coal to show the lowest slagging propensity was Hunter Valley, an
Australian coal with a high quartz content and a high Al:Si. In general, a good
relationship was obtained between the Jones Slagging Index and the basicity of the
coal ash. Two notable exceptions from this relationship were Goedehoop, a South
African coal and Harworth, a UK coal. A detailed microstructural and chemical
analysis of these deposits revealed an unusual crystallisation sequence for the
Goedehoop ash (a coal with a lignitic ash) that reduced the sintering behaviour, while
the enhanced slagging propensity of the Harworth ash was thought to be a
consequence of the synergistic effects of relatively high levels of CaO and high
Fe2O3 in the ash.

6. Mineral matter in the coals was identified by XRD and CCSEM analysis and found
to consist of a relatively small range of clays, quartz, sulphides and carbonates.
37
Many other somewhat exotic minerals were encountered, but in such small
proportions that the effects on combustion were considered to be negligible. The
proportions and distributions of minerals varied widely from one coal to another.
While pyrite (FeS2) was commonly found to be the principle Fe containing mineral in
Northern hemisphere coals, siderite (FeCO3) was much more common in coals from
the Southern hemisphere. Southern hemisphere coals tended to have higher
proportions of kaolinite: illite than found in the Northern hemisphere coals.

7. Mineral transformations were studied using the EFR that simulates the time
temperature conditions that coal and mineral particles experience in a pulverised coal
boiler. Combinations of minerals, eg. clays mixed with pyrite and calcite, showed that
a limited amount of chemical interaction took place between decomposed mineral
particles prior to deposition.

8. Transformations continued to take place after ash particles had deposited,


although at rates that were much slower than at the high temperatures encountered
during the combustion process. Deposits were generally very heterogeneous, with
residual quartz and iron oxide particles dissolving only slowly in an aluminosilicate
melt. Sintering of aluminosilicate ash particles occurred readily, with the rate of
sintering increasing with an increase in the bascity of the ash.

9. Slagging predictions for the CTF deposits based on a CCSEM characterisation of


the deposits agreed well with the Jones Slagging Index. CCSEM characterisation
was used to explain the anomalous slagging behaviour of the Goedehoop and
Harworth deposits.

10. Furnace bottom ash (FBA) and electrostatic precipitator ash (PFA) find a ready
market as cement replacement materials if the ashes fall within the required
specifications concerning carbon content and the amount of free CaO in the ash.
Only the Powder River Basin coals and Goedehoop gave ashes that might not meet
the current specifications.

11. The current coal characterisation methods that have been developed in recent
years to predict combustion performance and coal ash slagging behaviour were, with
just one or two exceptions, considered to be satisfactory in predicting the combustion
and slagging behaviour of the coals. In the case of the coals and ashes that did not fit
the established relationships, then a more detailed examination was necessary to
establish the reasons for anomalous behaviour.

11. Further work


There seems little doubt that as the regulations on gaseous and particulate
emissions from coal-fired power stations become ever more restrictive, then the
blending of coals with different characteristics will become a favoured option for
many generators. The combustion behaviour of blends is difficult to predict,
particularly if the coals that make up the blend are of a very different nature. This
project has shown that a limited amount of interaction can occur between
decomposed mineral particles prior to deposition. Thus, as previous studies have
shown [Manton et al (1996 &1999)], the slagging propensity of a coal blend may be a
non-linear function of the slagging propensity of the constituent coals. Further work
on blended fuels is required if these phenomena are to be understood.

38
World-wide attention is now being focused on the fate of some of the more
hazardous trace elements present in coals, in particular Hg and As. The trace
elements are known to reside in both the organic fraction of the coal and the mineral
matter. Their fate on combustion is controlled by a combination of the high
temperature physical chemistry and the mineral matter transformations. Continued
use of coal for power generation will require a greater understanding of these issues.
Only with an improvement in our knowledge of these phenomena can methods to
counter the less undesirable effects devised.

12. References
Bailey, CW, Bryant, GW, Matthews, EM and Wall, TF. Investigation of the high-
temperature behaviour of excluded siderite grains during pulverized coal combustion.
Energy and Fuels, 12, 464-469 1998.
Bryers, RW. Fireside slagging, fouling and high-temperature corrosion of heat-
transfer surface due to impurities in steam-raising fuels. Prog. Energy Combust. Sci.,
22, 29-120 1996.
Cloke, M., Lester, E. and Thompson, A.W., Fuel, 81 (2002), 727-735.
Couch G. Understanding slagging and fouling in pf combustion. IEACR/72, IEA Coal
Research, London, UK 1994.
Cunningham, ATS, Gibb, WH, Jones, AR, Wigley F and Williamson, J. The effect of
mineral doping of a coal on deposition behaviour in Inorganic transformations and
ash deposition during combustion (editor Benson SA). ASME, New York, 271-284
1992.
Gibb, WH. The role of calcium in the slagging and fouling characteristics of
bituminous coals. J Inst. Energy, 54, 206-212 1986.
Groves, SJ, Williamson, J and Sanyal, A. Decomposition of pyrite during pulverised
coal combustion. Fuel, 66, 461-466 1987.
Hutchings, IS, West, SS and Williamson, J in ‘Applications of advanced technology
to ash-related problems in boilers’, Eds LL Baxter and R DeSollar, Plenum, New
York, 1996, 201-222
Jackson, PJ. and Jones, AR, (1989) Controlled slagging and fouling trials in
500MWE boilers firing individual coals, in “Slagging and Fouling due to Impurities in
Combustion Gases” ed R.E. Barrett, United Engineering Trustees, New York.
Lester, E, Watts, D, and Thompson, A.W. A novel automated image analysis method
for maceral analysis, Fuel, 81, (2002) 2209-2217
Manton, NJ, Riley, GS and Williamson, J., A laboratory assessment of the slagging
propensity of blended coals, 212th American Chemical Society National Meeting,
Fuels Division, Vol. 41, No3, pp 1113-1117, 1996.
Manton, NJ, Riley, GS and Williamson, J., Changes in slagging behaviour with
composition for blended coals, in “The Impact of mineral Impurities in Solid Fuel
Combustion”, eds. R.P.Gupta, T.F. Wall and L. Baxter, Kluwer Academic/ Plenum
Publishers, New York 1999, pp 297-308
Nowok, JW, Benson, SA, Jones, ML and Kalmanovitch, DP. Sintering behaviour and
strength development in various coal ashes. Fuel, 69, 1020-1208 1990.

39
Raask, E. Mineral impurities in coal combustion. Springer-Verlag, New York, 1985.
Russell, NV, Wigley, F and Williamson, J The roles of lime and iron oxides on the
formation of ash and deposits in PF combustion, Fuel, 81, (5), 673-681, 2002
Scott, DH. Ash behaviour during combustion and gasification. IEACR/24, IEA Coal
Research, London, UK 1999.
ten Brink HM, Eenkhoorn, S and Weeda, M. The behaviour of coal mineral
carbonates in a simulated coal flame. Fuel Processing Technology, 47, 233-243
1996.
Parker, K.R., (ed.) “Applied Electrostatic Precipitation”, Blackie Academic &
Professional, 1997, ISBN 0 7514 0266 4
UKQAA Technical Data Sheet 1.1, “PFA and Fly ash as an addition in concrete
specified to BS EN206-Part 1: 2000”.
Wain SE, Livingston WR, Sanyal, A and Williamson, J. Thermal and mechanical
properties of boiler slags of relevance to sootblowing in ‘Inorganic transformations
and ash deposition during combustion’ Ed. Benson, SA, ASME, New York, 459-470
1991.
Wells, JJ, Wigley, F, Foster, D, Gibb, W.H. and Williamson, J. “The relationship
between excluded mineral matter and the Abrasion Index of a coal”, Fuel, 83, 359-
364, 2004(a).
Wells, JJ, Wigley, F, Foster, D, Gibb, W.H. and Williamson, J. “The nature of mineral
matter in a coal and the effects on erosive and abrasive behaviour”, Paper accepted
by Journal of Fuel Processing Technology, publication late 2004(b).
Wibberley LJ and Wall TF. Alkali-ash reactions and deposit formation in pulverized-
coal-fired boilers. Fuel, 61, 87-92, 1982.
Wigley, F and Williamson, J. Mineral matter in UK power station coals - distribution
and association in ‘Coal Science’ Pajares, JA and Tascón, JMD eds, Elsevier 195-
198 1995.
Wigley, F, Williamson, J and Jones, AR. Transformations of coal mineral matter in
the formation of rig deposits in ‘Mineral matter and ash deposition from coal’ Bryers,
WR and Vorres, KS. United Eng. Trustees, New York, 347-358 1990.
Wu, Z, “Prevention of particulate emissions”, IEA Coal Research Report CCC/40,
2000, ISBN 92-9029-352-7

Acknowledgments
Financial support for this work was provided by the Department of Trade and Industry
under the Cleaner Coal R&D Programme (Project 120) and by E.ON (Powergen),
TXU Europe and Rio Tinto Technology. In addition, E.ON, TXU Europe and Rio Tinto
Technology provided coals, combustion test facilities and much technical experience.

40
Appendix 1

Personnel involved and contact details of collaborating organisations.


Organisation Personnel Contact details
involved

Imperial College Prof J Williamson Department of Materials


London Mr F Wigley Imperial College London
South Kensington
London SW7 2AZ
Tel: 020 7594 6747
Fax: 020 7594 6748
Email: jim.williamson@imperial.ac.uk

E.ON (Powergen) Dr W Quick E.ON UK plc


Dr WH Gibb Power Technology Centre
Ratcliffe-on-Soar
Nottingham NG11 0EE
Tel: 011 5936 2440
Fax: 011 5936 2205
Email: will.quick@e.on.co.uk

Rio Tinto Technology Dr C Cross Rio Tinto Technology Development


Ltd
PO Box 50
Castlemead
Lower Castle Street
Bristol BS99 7YR
Tel: 011 7927 6407
Fax: 011 7927 3317
Email: chris.cross@riotinto.com

TXU Energy Mr PR Cooper Precol Associates Ltd


52 Kelvedon Drive
Rushmere St Andrew
Ipswich
Suffolk IP5LQ
Tel: 014 7372 7997
Fax: 014 7372 7997
Email:

University of Dr M Cloke SChEME


Nottingham Dr A Thompson University of Nottingham
University Park
Nottingham NG7 2RD
Tel: 011 5951 4169
Fax: 011 5951 4115
Email:
michael.cloke@nottingham.ac.uk

41
Appendix 2 Proximate, ultimate and ash analyses of coals
Coal Harworth Bailey PRB 1 PRB 2 Bowen Hunter Prodeco Goedehoop Talcher Jacobs Ranch/ PRB 1/ Bailey
basin Valley PRB 1 blend blend
Origin UK Eastern USA USA PRB USA PRB Australia Australia Colombia South Africa India USA PRB USA PRB
Moisture(%) 2.5 2.3 18.0 19.7 4.8 3.2 3.7 2.8 9.7 18.9 11.2
Ash, %ar 14.4 8.9 3.7 5.8 7.6 9.8 8.6 13.1 39.7 4.79 5.6
Vol Matter,% ar 31.4 34.5 34.5 35.0 27.8 30.7 35.5 25.4 24.0 43.7 34.8
GCV,kJ/kg, ar 28870 30970 24120 22450 28900 29910 29620 28230 15110 23404 27450
NCV, kJ/kg, ar 27836 29912 22942 21278 27950 28934 28508 27356 14342 22231 26313
Sulphur, %ar 2.25 1.26 0.33 0.33 0.39 0.46 0.66 0.70 0.37 0.46 0.8
Chlorine, %ar 0.20 0.21 0.01 0.01 0.01 0.02 0.5 0.01 0.01 0.02 0.17
Carbon, %DAF 82.40 83.67 76.28 77.42 81.76 82.99 81.39 83.45 74.52 75.89 80.61
Hydrogen, %DAF 5.52 5.32 4.44 4.37 4.48 4.87 5.49 4.52 4.94 4.39 4.89
Nitrogen, %DAF 1.78 1.66 0.98 1.12 1.77 1.86 1.68 2.03 2.07 1.1 1.41
Oxygen, %DAF 7.35 7.70 17.87 18.63 11.53 9.73 10.12 9.16 17.71 17.99 11.92
Vol Matter, 37.8 38.9 44.1 47.0 31.7 35.3 40.5 30.2 47.4 45.2 41.8
%DAF
Fuel Ratio 1.65 1.57 1.27 1.13 2.15 1.83 1.47 2.31 1.11 1.20 1.39
Calorific Value, 34,741 34,876 30,805 30,134 32,991 34,379 33,774 33,567 29,862 3.670 32.993
%DAF

Normalised ash compositions (wt%)


SiO2 50.8 56.4 36.3 39.2 61.5 81.6 63.4 43.1 67.2 38.6 48.5
Al2O3 26.1 25.37 19.7 20.9 31.0 13.2 20.0 33.3 24.3 19.1 24.9
Fe2O3 14.5 10.66 6.2 6.8 4.1 2.8 7.2 4.8 2.9 6.4 9.7
CaO 1.2 2.07 20.9 23.0 0.48 0.34 2.5 10.9 1.1 22.6 7.6
MgO 1.2 0.86 5.6 4.6 0.24 0.29 2.4 2.6 0.76 5.40 2.43
K2O 3.9 2.26 0.74 0.54 0.27 0.85 2.36 0.53 1.77 0.46 1.97
Na2O 0.8 0.47 7.57 1.45 0.10 0.14 0.80 0.29 0.10 3.99 2.99
TiO2 1.0 1.62 1.80 1.78 1.89 0.65 1.00 1.70 1.31 1.78 1.51
BaO 0/1 0.10 0.88 0.66 0.09 0.03 0.16 0.36 0.08 0.81 0.30
Mn3O4 0.05 0.03 0.06 0.03 0.01 0.03 0,07 0.08 0.03 0.05 0.02
P2O5 0.3 0.15 0.21 1.06 0.39 0.07 0.16 2.36 0.46 0.68 0.16
SiO2/Al2O3 1.94 2.22 1.84 1.87 1.99 0.20 3.18 1.29 2.76 2.02 1.95
Base:acid ratio 0.26 0.20 0.71 0.59 0.05 0,05 0.18 0.24 0.07 0.66 0.33

42
Appendix 3 Schematic diagrams of the combustion test facilities

E.ON Combustion gTest Facility,


g pshowing locationy of sampling
p ports.
g
Port 622, 626
(Flue gas temperatures)

Port 740 Port 726


(Flyash sample) (Deposition sample)
Port 722
(Flyash and deposition sample)
Ports 204, 205, 206
(OFA Ports)

Cyclone
(Flyash sample)
To Stack Port 313 (Opposite side)
(Char sample)

Burner
Bottom Ash Hopper

Imperial College London Entrained Flow Reactor

Feeder

1600°C

Probe deposit c ollection position


Uncooled mullite probe 1500°C
Gas temperature ~1250°C 1500°C
1450°C
Probe ash collection position
N2 gas quench probe
Gas temperature ~1250°C 1200°C
Same position as probe deposit

Cyclone ash collection 1200°C


Comparable to utility PFA

Cyclone

43

Anda mungkin juga menyukai