factors influencing properties of carbonelectrodes a calcined-anthracite or petroleum coke, the...

145

Upload: ngominh

Post on 15-Jul-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

Loughborough UniversityInstitutional Repository

Factors influencingproperties of carbon

electrodes

This item was submitted to Loughborough University's Institutional Repositoryby the/an author.

Additional Information:

• A Master's Thesis. Submitted in partial fulfilment of the requirements forthe award of Master of Philosophy at Loughborough University.

Metadata Record: https://dspace.lboro.ac.uk/2134/28434

Publisher: c© J.R. Farrington

Rights: This work is made available according to the conditions of the CreativeCommons Attribution-NonCommercial-NoDerivatives 2.5 Generic (CC BY-NC-ND 2.5) licence. Full details of this licence are available at: http://creativecommons.org/licenses/by-nc-nd/2.5/

Please cite the published version.

This item was submitted to Loughborough University as an MPhil thesis by the author and is made available in the Institutional Repository

(https://dspace.lboro.ac.uk/) under the following Creative Commons Licence conditions.

For the full text of this licence, please go to: http://creativecommons.org/licenses/by-nc-nd/2.5/

LOUGHBOROUGH UNIVERSITY OF TECHNOLOGY

LIBRARY

AUTHOR/FILING TITLE

___________ E~f.!.!': ~T~_~_.,.----~~-~--------- __ :

ACCESSION/COPY NO. .'

--- _______________ ~~:t.'2:~ L~ ~".. _______________ _ VOL. NO.

28 JUN 1996

27 Jtm 1S97

26 JUN 1998

CLASS MARK

009 '2~1 02 .

. ~IIIIIIIII~I~I~I~I~I~IIIIIII~IIIII~~II~

,~: . '\

::.

~ "'~- , .. ' .. -~;,~

Factors Influencing Properties

of Carbon Electrodes

by

John Robert Farrington

A Master's Thesis

submitted in partial fulfilment of the requirements

for the award of the degree of

Master of Philosophy

of the

Loughborough University of Technology

!" ~

•• J ~

' . .

© J. R. Farrington 1990

- ...

L.ougtIborough UnIIi~ 01 Technology t.Jbmry

DI!II& ...16 't:~

la-

I::: • .r.:

(;f q<.f'2.::!' I I) '\... "

Acknowledgements

I should like to thank my supervisor, Dr. I.W. Patrick, for. his advice and guidance

throughout this project.

I should also like to thank Dr. A. Walker for explaining many technical points to me,

and Mr. D. Hays for his assistance with the experimental work. The method used to

measure electrical resistivity of carbon cores was jointly developed with Mr. G. Ogden.

Abstract

Aluminium is obtained commercially from the Hall-Heroult Process, the electrolytic

reduction of bauxite (aluminium oxide). The electrodes used are made from different

types of carbon; the bulk of the electrode, known as the filler, is typically made from

either a calcined-anthracite or petroleum coke, the particles of which are bonded

together by the binder component of the electrode, which is usually coal-tar pitch.

Large quantities of electrodes are consumed each year; attrition is by a variety of

causes, and since there is no commercially realistic alternative to the Hall-Heroult

Process, any means of reducing loss of electrodes would improve the economics of the

extraction process.

The principal objective of this research project was to examine a range of electrode

materials used in the aluminium industry, investigating the relative contributions of their

pore structural characteristics and the quality of the bond between binder and filler to

their mechanical strength and electrical resistivity, with a view to identifying those raw

materials which contribute to inferior binder-filler bonding. The literature contains a

diversity of views on manufacturing methods for electrodes; this reflects the largely

empirical nature of the process.

Results were obtained for the tensile strength of carbon specimens, SEM fractography

of fractured carbon specimens, image analysis of polished carbon specimens and

electrical resistivity of carbon cores. Some tests were also made of the wettability (the

readiness with which pitch· wets a coke surface) and real density of fillers. Seven

electrode materials were examined, of which two used calcined anthracite as the filler

while the other five used petroleum coke. All used coal-tar pitch as the binder.

The results obtained supported previously-published work by indicating that the choice

of raw materials has a key effect in determining electrode properties, and that the

strength and porosity· characteristics of a carbon can be related using an

empirically-derived equation of the form S.N = k + a.W/p2, where S = tensile strength,

N = number of pores, W = mean wall size (thickness of carbon matrix) and P = mean

pore size. In fractography, a method was successfully developed for identifying and

categorising features observed on fracture surfaces of electrode specimens, and an

attempt was made to relate the data from this work to the image analysis results.

Contents

Page

1 Introduction 1

2 Literature Review 4

2.1 Use of Carbon Electrodes In the Aluminium Industry 4

2.2 Carbonisation 6

2.2.1 Introduction 6

2.2.2 Carbonisation of Coal 7

2.2.3 Carbonisation of Pitch 9

2.2.4 Mesophase Chemistry 11

2.2.5 Factors Affecting Mesophase Properties 14

2.2.6 Pore Development During Carbonisation 16

2.2.7 Effect of Carbonisation Conditions on Binder and Filler Properties 17

2.2.8 Graphitisation 18

2.3 Electrode Manufacture

2.3.1 Raw Materials

2.3.2 Specification of Binders and Fillers

2.3.3 Factors Affecting Electrode Quality

2.3.4 Industrial Scale Manufacturing Methods

2.4 Strength of Brittle Materials

2.4.1 Basic Theory

2.4.2 Fracture of Brittle Materials

2.5 Strength of Carbons and Graphltes

2.5.1 Brittle Fracture Theory and Carbons

2.5.2 Porous Brittle Materials

20

20 21

25

27

33

33

35

41

41

44

3

3.1

3.2

3.3

3.4

3.5

3.6

3.7

4

4.1

4.2

4.3

4.4

4.5

4.6

5

5.1

5.2

5.3

5.4

5.5

5.6

5.7

6

2.6 Electrical Resistivity 01 Carbons and Graphltes

2.6.1 Introduction

2.6.2 Structural Aspects

2.7 Conclusions to LIterature Review

Experimental Procedures

Electrode Materials Used

Tensile Strength Determination

Pore Structural Analysis

Fractography using Scanning Electron Microscope

Determination of Electrical Resistivity

Determination of Real Density

Wettability

Results

Tensile Strength

Pore Structural Analysis

Fractography

Electrical Resistivity

Real Density

Wettability

Discussion and Conclusions

Tensile Strength

Tensile Strength and Porosity Results

Tensile Strength and Fractography Results

Tensile Strength, Porosity and Fractography Results

Resistivity Results

Conclusions

Recommendations for Further Work

References

45

45

45

47

51

51

51

52

53

56

56

57

68

68

71

72

74

74

74

117

118

119

121

123

125

126

128

131

1 Introduction

Carbon and graphite electrodes are used in the manufacture of a variety of materials.

The principal industrial use is in the electrolytic extraction of aluminium from bauxite

by the Hall-Heroult process, the most important industrial process for the production of

this metal. Carbon or graphite electrodes are also used in the production of steel in

electric-arc furnaces, and in the production of sodium, magnesium, lithium and

chlorine. All the electrodes studied in this research project were carbon electrodes used

in the extraction of aluminium from bauxite.

A carbon electrode consists of two solid phases: the large particles which make up the

bulk of the electrode are known as the filler, while the component of the electrode

which is used to cement the filler particles together, and which also makes its own

contribution to the mechanical and electrical properties of the electrode, is known as the

binder. The filler in electrodes used in aluminium extraction may be petroleum coke (the

most common) or a calcined anthracite; the binder is usually coal-tar pitch.The

properties of the finished product, especially its mechanical strength and electrical

resistivity, depend to a large extent on the quality of the bond achieved between the

binder and the filler during the manufacture of the electrode; it is therefore as important

to specify the details of the manufacturing process, as well as the raw materials used, to

produce electrodes of consistent quality.

Whatever application they are used in, carbon or graphite electrodes must be

mechanically strong, in order to maintain structural integrity when they are moved.

Electrodes must also be able to carry a high electrical current without undergoing

excessive Joule (resistive) heating, and should therefore be oflow electrical resistivity.

They must show adequately high thermal conductivity, to be able to dissipate heat

effectively. They must be able to withstand the maximum temperature likely to be

reached in the operation of the plant; for example, in electrodes made with

electro-calcined anthracite, an irreversible reduction in strength can take place at high

temperatures, probably because of shrinkage of the binder component away from the

filler particles. Electrodes must also be resistant to thermal stresses, that is, large and

perhaps rapid temperature fluctuations caused by a need to shut down the plant at short

notice, or because of fluctuations in the current. They must be resistant to chemical

attack by, for example, electrolytes or evolved gases; they must be capable of being

fabricated reliably and on a repeatable basis; and they must be capable of being

produced and used on an economic basis. If the electrode is so badly made that it fails

on a gross scale, not only will a new electrode be needed, but there is likely to be a high

cost because of the lost production which will ensue. It is therefore essential to be

1

confident that the process used to manufacture the electrodes will result in electrodes of

consistently adequate quality. This is of particular importance in the aluminium

industry, because of the large numbers of electrodes used, and the fact that the

electrodes are gradually used up in the electrolytic reaction. Some of these requirements

are mutually exclusive, and a compromise is usually necessary.

The structure of carbon and graphite electrodes is closely related to their required

properties. A key feature is the two-phase structure; the properties of the binder and

fJ.!ler on their own are significant, and so is the quality of the interface between them. A

mechanically strong electrode of low electrical resistivity must be fabricated from

suitable binder and filler materials, and be well-bonded at the binder-filler interface.

The other major structural feature which determines the properties of the electrode is

porosity. Pores are present in all types of manufactured carbon and coke-based

material, and play a key role in determining the strength of the artifact; it is likely that

they correspond to the Griffith critical cracks which play a vital role in the fracture of

brittle materials.

The objectives of this research project were (1) to examine a range of

commercially-available electrode materials used in the aluminium industry,

investigating the relative contributions of their pore structural characteristics and

binder-filler bond quality to their strength and electrical resistivity, and (2) to identify

those raw materials which contribute to inferior binder-filler bonding.

The experimental techniques used were: measurement of the tensile strength of carbon

specimens, which also provided specimens for SEM fractography; polarised light

microscopy, which was used in conjunction with the scanning electron microscope

(SEM) to identify constituents of the carbons, such as electro-caIcined anthracite (ECA)

and petroleum coke; SEM fractography, which involved developing and using a

point-counting method to determine the mode or modes of failure of fracture specimens;

image analysis, which was used in conjunction with reflected light microscopy to

obtain pore structural data on the carbons; wettability, which involved studying the

readiness with which pitch wets a coke surface; and real density, which was used in

conjunction with apparent density to obtain values for the porosity of the carbons, as an

alternative method to the use of the image analysis system.

Some of the techniques used were developed for this research project, while others

were modified versions of existing techniques used for other materials. Seven different

electrode materials were examined.

2

This thesis is divided into a review of the relevant literature. a description of the

experimental techniques. a results section and a discussion.

3

2 Literature Review

2.1 Use of Carbon Electrodes in the Aluminium Industry

In tenns of the amount of carbon used, the principal industrial use of carbon electrodes

is the electrolytic extraction of aluminium from bauxite. The aluminium industry

worldwide requires approximately eight million tons a year of carbon for use in anodes

[1].

The principal ore is bauxite, hydrated aluminium oxide, which also contains impurities

in the fonn of silica, iron oxide and titanium oxide. These impurities must be removed

before the aluminium can be extracted. The purification process, invented by Bayer in

1890, depends on the fact that hydrated aluminium oxide dissolves in heated caustic

soda (sodium hydtOxide), but the impurities do not, enabling almost pure aluminium

oxide (alumina) to be separated. In the Bayer process, the bauxite is pulverised,

washed and calcined until it is a powder. The powder is then mixed into a hot solution

of sodium hydroxide and pumped into large tanks containing caustic soda, which

dissolves out the aluminium hydroxide, fonning sodium aluminate. Crystals of

aluminium hydroxide are fonned when the sodium aluminate solution is seeded with

crystals of aluminium hydtOxide. After cleansing, the aluminium hydroxide crytals are

calcined to drive off the water, and the resulting alumina, is ready to be used for the

extraction of aluminium.

Alumina can be electrolysed into aluminium metal and oxygen by a high-amperage

current, but it cannot easily be liquified as it has a melting point of greater than 2000°C.

However, if the alumina is dissolved in cryolite (sodium aluminium fluoride), in the

ratio of 5% alumina, the mixture will melt at about 900°C. without appreciable

decomposition of the cryolite. This step is an integral part of the Hall-Heroult process.

Calcium fluoride may be added to the mixture to reduce the melting point further.

Unlike steel, aluminium is not made in bulk in a large furnace, but is produced in a

fairly large number of comparatively small units known as pots. A modem reduction

furnace consists of a shallow rectangular steel container lined with refractory bricks,

with an inner lining of carbon serving as the cathode. Direct current enters the cell

through the anodes, which are bulky carbon blocks suspended from above and dipping

into the molten solution of alumina in cryolite. (See Figure 2.1). A modem lOO,OODA

furnace is about IOm long and 5m wide, and produces up to 750kg of aluminium every

24 hours. Such cells typically work in a series of 150-200 called a pot line, rated at

about 75MW and using about 15kWh of electricity for every kilogram of aluminium

4

produced. The anodes are gradually oxidised to carbon dioxide (as well as being

subject to other losses, described below), and must therefore be replaced as they wear

out.

Another type of furnace uses anodes baked as the reduction proceeds: these are known

as Soderberg anodes, in which carbon paste is fed into a steel tube above the

electrolyte, the paste becoming hardened as it descends towards the pot.

Since aluminium is more dense than the molten solution of alumina in cryolite, the

aluminium produced sinks to the bottom of the cell, whence it is periodically extracted

as a liquid either by siphoning or by being tapped from the bottom of the cell. It may be

cast into ingots of the required shapes and sizes, or it may be re-smelted for further

purification.

The oxygen liberated at the anodes reacts with the carbon to form carbon monoxide,

which burns at the top of the furnace to form carbon dioxide. This is the principal

means by which anode material is lost, other losses being by reaction with the cryolite

and sloughing off. Because of the scale of aluminium production, it is uneconomic to

use the more expensive graphite anodes, although these have lower resistance and

hence lead to smaller Joule heating losses.

The economic production of aluminium depends largely on the availability of low-cost

electricity. The stoichiometry of the reaction requires a minimum of 0.33kg of carbon to

be consumed for the production of lkg of aluminium. A larger amount is always

required in practice, the efficiency of Hall-Heroult cells typically being 75-80%.

Excessive carbon consumption occurs because of losses through re-formation of

aluminium oxide, air oxidation of the anode (air burn), reaction of the anode carbon

with carbon dioxide, and dusting (small-scale crumbling of the anode which results in

small particles of carbon falling into the electrolyte). Some of these effects are

inter-related; for example, in the areas where reaction with carbon dioxide takes place,

the surface of the anode is likely to become pitted and fissured, which will promote

dusting. Similarly, an area of the surface which becomes pitted will be more prone to

reaction with carbon dioxide [2]. Thermal shock on start-up also acts to reduce the life

of the electrode. Since anode costs form approximately 20% of the total cost of

aluminium production, ways of minimising excessive carbon consumption are

potentially of considerable economic benefit. About 12 kWh of electricity, 1 kg of

bauxite and 0.5 kg of carbon anode are required to produce 0.5 kg of approximately

95% pure aluminium [3].

5

2.2 Carbonisation

2.2.1 Introduction

Carbonisation may be defined as the destructive distillation of organic material out of

contact with air, accompanied by the formation of carbon, plus liquid and gaseous

products. Carbonisation of coal yields coke; carbonisation of wood, sugar and other

materials yields charcoal.

Chemically, carbonisation is essentially an aromatic growth and polymerisation

process, in which a small aromatic structure is polymerised to an aromatic polymer,

which may then become three-dimensionally ordered to form graphite.

6

2.2.2 Carbonisation of Coal

Coal became widely used as a metallurgical fuel in the seventeenth century. after

supplies of wood for charcoal-making became inadequate (mainly because of the use of

wood for ship-building). The hearth process of heating coal in rounded heaps. known

as beehive ovens. predominated for almost a century. until more sophisticated

nineteenth-century ovens enabled by-products to be recovered.

Identifying the physical and chemical principles behind carbonisation has been a slow

process; for several hundred years. carbonisation was carried out on an empirical basis.

Details of the mechanisms involved have only been worked out this century. using

modern analytical techniques.

As coal is heated. it undergoes chemical changes which result in the evolution of gases.

leaving a solid carbon residue. Not all coals can form coke; those which can (coking

coals). soften and become plastic in the temperature range 400-500°C. The coal

particles coalesce. then the resulting mass swells and subsequently resolidifies. forming

a porous solid. With a further increase in temperature. the solid undergoes non-uniform

contraction at a rate which partly depends on the amount of volatile matter in the coal.

Factors affecting the quality of the resulting coke include the type of coal. rate of

heating. coal particle size and pressure used.

The study of carbonisation of a coal· resulted in the discovery of the carbonaceous

mesopluise. which was discovered during work on one of the principal uses of coals.

namely coke production. Polarised light microscopy was used extensively for this

work. Working with samples of coal from a seam which had been modified by a

geological source of heat. Taylor observed a loss of optical anistropy as the vitrinite

component became plastic in samples from points successively nearer the source of

heat; subsequently. he noticed the formation and growth of small spherical bodies [4].

When these bodies reached a size when they started to interfere with each other's

growth. a mosaic-like structure began to form. Completion of this structure coincided

with the resolidification point of the coal.

Efforts to observe mesophase formation in coals in the laboratory have been almost

totally unsuccessful; nearly all subsequent work has not been carried out on coal. but

on pitch and related substances. The formation of a liquid-crystalline phase during

carbonisation of certain organic materials such as coal-tar pitch and petroleum pitch was

fust described by Brooks and Taylor [5]. The formation of mesophase usually takes

7

place between 350-450°C. As described above. small spherules emerge from an

optically isotropic matrix. The spherules subsequently grow (at the expense of the

matrix material). and coalesce into large anisotropic regions which show the properties.

especially long-range order. of nematic liquid crystals. With increasing temperature. the

mesophase begins to predominate and becomes viscous. and eventually freezes to form

coke. Any mesophase structures remaining are frozen in place. Various chemical and

mechanical factors affect the degree to which the microstructure features are present in

the resulting coke; these features play a part in determining the theImal and electrical

properties of the graphite.

There are two types of coking coal [6]:

Prime quality - these coals have a volatile matter content of between 19.6 and 32 wt

% (measured on a dry. mineral-free matter basis). Only a small amount of intra granular

swelling takes place. because the plastic phase is too viscous and the rate of

devolatilisisation too low. However. adjacent particles adhere to each other. and the

resulting coke has numerous small-diameter pores with thick walls. which enhances its

mechanical strength.

Lower quality - these coals have a volatile matter content of greater than 32 wt %.

Considerable intragranular swelling may take place. and the whole mass may foam.

The resulting coke haspores of various sizes. with thin walls. and is of relatively low

mechanical strength.

8

2.2.3 Carbonisation of Pitch

In general, pitches will, upon heating to the appropriate temperature, pass through an

optically isotropic plastic or liquid phase, and then form the optically anisotropic

mesophase. Selective area electron diffraction has revealed [5] that the mesophase

consists of aggregations of planar liquid crystal-like structures, as described below,

which form spheres as a compromise between the tendency of the liquid crystals to

form two-dimensional arrays, and the need to minimise surface energy. The mesophase

spheres have properties which resemble those of nematic liquid crystals [7].

Mesophase Development and Liquid Crystals

Liquid crystal systems were discovered in 1888. They were noticed as being systems

with unusual fluid properties, namely being more highly ordered than ordinary liquids

but without actually being crystalline. The term "liquid crystal" is actually a misnomer,

but one that has become generally accepted. Liquid crystal systems may be divided into

two categories: nematic (thread-like) and smectic or cholesteric (soap-like). The

mesophase spheres which are produced in pitch carbonisation resemble nematic liquid

crystals, with the important difference that the transformation is irreversible.

Mesophase formation is observed as the separation out of spherical droplets from the

isotropic liquid, and their subsequent coalescence to give a molecular arrangement

which may be likened to a shoal of fish [8].

Factors which assist molecules to form nematic liquid crystals include [7]:

- Rigidity along their principal molecular axis.

- The presence of benzene rings, producing flat portions in the molecule.

- The existence of strong dipoles and easily polarisable groups.

- A structure extended either linearly, two-dimensionally, or both.

In the carbonisation of aromatic hydrocarbons, petroleum pitches and coal-tar pitches,

pyrolysis of the isotropic liquid results in a complex series of reactions detailed in the

section on mesophase chemistry (2.2.4). In the absence of agitation, the liquid crystals

produced tend to stack extensively. This continues as the liquid crystals are modified by

increasing heat treatment temperatures (HIT) to form semi-coke and then coke. This

mechanism produces a stacked lattice structure, which can then undergo further

changes to form graphite.

9

Growth of Mesophase Spheres

As suitable precursors of the liquid crystals are formed, they encounter similar

molecules in the initially isotropic liquid, and form planar aggregations by means of

dipole effects and Van der Waals forces. To minimise surface energy, the aggregates of

liquid crystal-like molecules form spheres, in which the planar structures are observed

as lamellae, mostly stacked parallel to the equatorial plane of the sphere but with some

aligned at right angles to the interface between the sphere and the liquid phase [9].

Insoluble material is never found in the spheres; some of it may be present in the

starting material, but most or all is produced during carbonisation. Insoluble particles

introduced into the starting substance are found to behave in the same way as

naturally-occurring insoluble material [8].

The spheres grow with increasing temperature, until eventually they coalesce and the

liquid phase disappears. The spheres grow to between 10 and 100~m in diameter; they

usually reach a maximum size in the temperature range 400-500°C. Subsequent

coalescence of the liquid crystal structures to form semi-coke leads to the anisotropic

structures observed under the optical microscope under polarised light. Four categories

of sphere growth and coalescence have been described [7]:

(1) Spheres originate at approximately the same time thoughout the mixture (Le. at

the same HTT), and so coalescence occurs over large volumes at the same time.

This behaviour is seen during carbonisation of coal-tar pitch, which contains

relatively small amounts of QI.

(2) Spheres originate at different times and grow at different rates, resulting in a large

range of sizes. Coalescence is non-uniform. This behaviour is observed in the

carbonisation of petroleum pitch [10].

(3) Spheres originate at the same time throughout the mixture but do not grow

beyond a relatively small size. The anisotropic units produced on coalescence are

much smaller than those produced in (1) or (2) - approximately 2J.lm in diameter

as opposed to about 25~. This behaviour is observed with Gi1sonite pitch and

carbons from heterocyclic compounds [11].

(4) The sphere growth pattern is affected by the presence of solids within the liquid

phase. In such cases, sphere coalescence is severely restricted by small particles,

approximately 20nm in diameter, such as carbon black [12]. However, larger

particles may produce an increase in the size of the so-called domains, by

providing a surface on which further growth can take place [13].

10

2.2.4 Mesophase Chemistry

The degree to which a carbonaceous material undergoes transformation to mesophase,

and subsequently graphitisation, depends on its chemical makeup. The following

substances show a transition to pitch (if they are not pitches already), then to

mesophase, then to a semi-coke [14].

Substance Temperature at which

mesophase forms eC)

Coke-oven pitch 425

Vertical-reton pitch 425

Petroleum bitumen 425

Vitrinites from bituminous coals 460

Naphthalene, anthracene and other 400

polynuclear aromatic hydrocarbons

Under the conditions in which carbonisation takes place, it is likely that the more

reactive hydrocarbons in the pitch phase undergo dehydrogenation and condensation to

form larger, planar species, which separate out from the pitch as the mesophase, in

partially-ordered molecular assemblages. For example, dibenzanthrone, a 7 -ring

polynuclear quinone, was found to form a coke-like material at 520°C [5].

The chemistry of mesophase, and hence of carbonisation, is extremely complex,

involving many different types of chemical reaction including bond cleavage,

polymerisation, molecular rearrangement, and hydrogen transfer. The diagram below

shows a generalised scheme for the transformation of an aromatic hydrocarbon to

carbon and graphite.

11

AROMATIC HYDROCARBON

.... 300-500°C

COMPLEX, FUSIBLE MIXTURE OF HYDROCARBONS

SEMI-COKE (INFUSIBLE POLYMERIC HYDROCARBON MIXTURE)

.... 1000'C

2-D CARBON POLYMER

.... 3000"C

GRAPHITE (3-~ STRUCTURE)

Chemically, mesophase formation involves low molecular weight components being

driven off, while higher molecular weight components start to associate and precipitate

out as anisotropic spheres from the lower molecular weight isotropic phase. The entire

pitch is eventually transformed to mesophase and then to an anisotropic, infusible, solid

coke. Mesophase pitches exhibit typical properties of nematic liquid crystals, but also

behave as supercooled glasses, exhibiting nematic textures at room temperature in the

solid state. Solid pitches behave as eutectic glasses, exhibiting glass transitions and

melting to give a low-viscosity liquid over a wide temperature range. These properties

are significant with regard to the transformation to carbon.

Unless the precursor is already aromatic, the first stage in the carbonisation process is

the formation of aromatic molecules from aliphatic molecules or substituents [15]. The

subsequent reaction sequence, or the reaction sequence if the starting material is

already aromatic, is essentially a free-radical process, involving a series of

bond-cleavage reactions.

(1) C-H and C-C bonds are broken, producing free radicals.

(2) Molecular rearrangment takes place.

12

(3) Polymerisation occurs, in two stages, with free-radical intermediates being

involved in both:

(i) Polymerisation through a reactive radical produced by hydrogen

dissasociation, or by a disproportionation reaction.

(ii) Formation of stable radicals from naphthalene oligomers by the loss of a

single hydrogen atom in a second-stage condensation process.

(4) Large aromatic molecules undergo condensation.

(5) Side chains are eliminated.

Production of large aromatic molecules does not necessarily mean that a graphite-like

structure will be formed. Before this can happen, the aromatic molecules must have the

correct shape and reactivity to be able to form a two-dimensional structure which can

provide the basis for the formation of graphite.

13

2.2.5 Factors Affecting Mesophase Properties

As described previously, certain carbonaceous compounds show the development of

anisotropic structures upon heating to the appropriate temperature. This section will

consider which substances fall into this category, with specific reference to the factors

which influence anisotropy development.

The structure (or lack of structure) established upon the coalescence of the liquid

crystals is largely carried through to the semi-coke and graphite. The structure formed

. depends on several factors:

- The chemical composition of the pyrolysing system

- The presence and size of solids.

- The rate of heating.

- Presence of factors such as convection currents and gas evolution.

Since it is observed that solid surfaces provide a nucleation site for mesophase growth

[8], insoluble particles probably provide such sites but are subsequently excluded from

the mesophase spheres.

If a pitch which contains solids, and a pitch which does not contain solids are

carbonised under the same conditions, the pitch with the solids tends to contain more

but smaller mesophase spheres than the pitch without the solids. There is, however,

little difference in the proponion of mesophase present

If the carbonaceous compound is not agitated when it is being heated, the mesophase

spheres tend to be larger, as there are fewer nuclei available.

Mesophase formation and properties are influenced by the kind of pitch used, that is,

whether it was derived from coal (coal-tar or coal-liquids), from components of

petroleum such as naphtha-tar pitch or decant-oil pitch), or from organic materials [16].

Another factor which influences mesophase formation is the type (if any) of

pre-treatment used, such as solvent fractionation or hydrogenation. Finally, the

heat-treatment conditions are significant; these include temperature, residence time,

heating rate, gas-blowing rate and stirring rate.

The most important factors influencing the rate of growth of mesophase spheres are

temperature and time. Generally speaking, the slower the rate of carbonisation, the

fewer and larger will be the spheres. Below a limiting temperature (400°C in the case of

many coke-oven pitches), no sphere formation occurs, even when a large period of

14

time is allowed. At a temperature slightly above the limiting temperature, sphere

formation starts. The longer the time allowed, the greater the proportion of pitch will

be converted to mesophase; the increase with time is apparently asymptotic.

A t temperatures well above the the limiting temperature, prolonged heating leads to

complete conversion to mesophase. The longer the time allowed, the more complete the

conversion. Agitation, and presence of insoluble particles, seem to assist mesophase

formation; the insoluble particles presumably act as nucleation centres. Surfaces act in a

similar way. If the mesophase is not agitated by stirring or bubble formation, the

available nuclei for sphere growth are removed, and the result is fewer but larger

spheres compared to the case when agitation does take place [8].

Before the spheres begin to interfere with each other's growth, their profiles are

circular. When the mesophase spheres form the dominant phase, distortions are seen in

the shape of the spheres, usually on a local basis at first, before becoming more

widespread.

Mesophase properties play the main part in the establishment of the structure and

properties of carbons and graphites formed by the process of carbonisation.

Petroleum-derived precursors yield a wide range of mesophase morphologies ranging

from fIne-textured and isotropic cokes to the fibrous structure of needle-coke.

15

2.2.6 Pore Development During Carbonisation

It has been reponed [17] that the pore structure of coke is largely determined during the

plastic stage. Pore fOl1llation can be divided up into four stages:

(1) Pore nucleation within larger coke panicles.

(2) Intraparticulate pore growth, with consequent swelling of the particles leading to

fusion.

(3) Continued growth of pores, reaching a maximum size after fusion.

(4) A decrease in pore size, leading to compaction of the structure near the

resolidification temperature.

It has been demonstrated that pore formation within single particles of coal is dependent

on particle size. Although fusion can only take place when pore growth causes the

particles to swell and fill the inter-particular voids, softening is still the principal

requirement for fusion, since pore growth is dependent both on the amount of volatile

matter in the pores, and on the surface tension, which decreases with increasing

fluidity. Since pores in coke are highly interconnected, it is likely that compaction of the

coke is associated with their partial deflation.

16

2.2.7 The Effect of Carbonisation Conditions on Binder

and Filler Properties

The propenies of the pitch used as the binder in pre-baked carbon anodes in the

aluminium industry have a significant effect on the propenies of the anode. Coal-tar

pitches are usually preferred to petroleum pitches, as they give a more effective binder

coke in the finished (baked) electrode [18]. This occurs because growth and

coalescence of the mesophase spheres are inhibited by the presence of

quinolene-insoluble material (QI) in the early stages of carbonisation, with the result

that the flow texture of carbon does not develop very well [7, 8, 19]. In order to obtain

a pitch with the optimum amount ofQI, various methods of removing QI from coal-tar

pitch, or of negating its effects, have been proposed [20], including the use of

additives, and co-carbonisation with QI-free coal-tar pitch.

The size and shape of the structural units in the resulting coke is a direct consequence of

the formation, growth and coalescence of spherical mesophase units, which are

themselves formed from large planar molecules produced by pyrolysis of the pitch

constituents (see Section 2.2.4) The pre-graphite structures are developed during the

lifetime of the mesophase, and an extensive lamelliform morphology results if the

conditions are favourable for sphere coalescence. Poor sphere coalescence may occur if

the pitch has low aromaticity [21], a high fluid-phase viscosity, or ifforeign atoms or

inen particles are present in significant quantities [8]. In such cases, examination of the

resultant carbon under polarised light reveals small, isochromatic, randomly-orientated

areas.

Considering cokes used as fillers in electrodes, some work has been done on needle

coke, which is used in applications where a highly-ordered structure is required.

Studies on decant oil [22] indicated that (1) a significant proponion of the feedstock

evolved at an early stage of carbonisation, and hence had little effect on the resulting

coke structure, and that (2) the structure of the resulting coke was determined partly in a

similar way to the structure of binder coke, as described above, and partly by the

manner in which gases were evolved, the latter factor being the chief determinant of the

axial textural arrangement which gives rise to the large scale needle-type structures.

Feedstocks used to produce needle coke must be chosen carefully; they must have a

high transformation temperature, in order to form a very fluid mesophase which can

easily be deformed by bubble percolation to produce a fine, fibrous microstructure

[23]. Extensive mechanical deformation as the mesophase hardens is essential for the

formation of a good needle coke.

17

2.2.8 Graphitisation

Graphite occurs naturally, but can also be manufactured by heating cenain forms of

carbon to temperatures of over 2000°C. These carbons are known as graphitising

carbons, and are generally those carbons which pass through the liquid stage involving

mesophase. Examples of precursors of graphitising carbons are high-temperature

coal-tar pitches, petroleum bitumen, polymers such as PVC and aromatic hydrocarbons

such as naphthacene [8]. Such precursors are usually rich in hydrogen and deficient in

oxygen.

Graphitising carbons are usually coke-like in appearance and exhibit optical anisotropy

under the polarising microscope. They have a compact structure, with a nearly parallel

arrangement of crystallites, little cross-linking between crystallites and small pores.

Consequently, the crystallites retain a high degree of mobility; with increasing

temperature, adjacent crystallites become aligned in parallel, eventually producing a

large-scale graphite structure. By contrast, non-graphitising carbons do not form a

well-ordered graphite structure even on heating to 3000oC. They have

randomly-orientated crystallites, much cross-linking of crystallites and large pores.

Precursors of non-graphitising carbons include coals with a high percentage of volatile

matter (little hydrogen and large amounts of oxygen). Upon heating, they do not pass

through a mesophase stage, and a rigid cross-linked structure is fonned even at low

temperatures, so a graphite-like structure cannot subsequently form.

The process of graphitisation (th.e movement of atoms to produce a three-dimensional

graphite lattice) commences at'approximately 17000 C and is complete by approximately

3000°C. The rate of graphitisation increases greatly with temperature, and is more rapid

in carbons which already show some degree of graphite-like structure.

The usual method of heating is electrical, in which case the graphite formed is known

as electrographite. Graphitisable carbons may also be formed by deposition from the

vapour phase, by passing an organic gas over a surface at a temperature of

800-1200°C. Such carbons are known as pyrolytic carbons. Ifthe surface is at 2000°C

or a higher temperature, graphite of a very high degree of perfection is formed directly;

this is known as pyrographite. If the surface temperature is between 1200-2000°C,

graphite may form, but there will be a tendency for soot to be incorporated in the

deposit, which will hinder further graphitisation.

On a crystallographic basis, graphitisation may be considered to be the development of

an ordered crystallographic structure from an initial partially-organised graphitisable

carbon. Modern concepts of the process indicate an increase in crystallographic

18

ordering as a result of elimination of side chains. between 400-70QoC. followed by

elimination of non-carbon atoms such as hydrogen and sulphur. and annealing of

inter-layer defects between 700-27OQ°C.

Since industrial graphitisation does not produce single-crystal graphite. the properties

of the binder and filler developed before graphitisation remain after the process is

complete. and therefore largely determine the properties of the fmished product

19

2.3 Electrode Manufacture

2.3.1 Raw Materials

Pre-baked anodes for use in the extraction of aluminium are produced by the slow

carbonisation of a pitch-coated, calcined, compacted aggregate of filler particles.

Several materials may be used as the fIller, the most common being petroleum coke

produced by the delayed coking process. Calcined anthracites are used in some cases

for aluminium electrodes, while needle coke may be used in specialised applications.

The binder is usually a coal-tar pitch obtained from the COking industry. Petroleum

pitches can also be used as binders, but coal-tar pitches are generally preferred, as the

highly disordered microstructure of the binder coke formed during baking results in a

high proportion of edge carbon atoms available for chemical bonding and consequently

a stronger anode. By contrast, highly aromatic petroleum pitches have extensively

aligned carbon basal layers, and so exhibit poor bonding characteristics at right angles

to these layers [18].

20

2.3.2 Specification of Binders and Fillers

Binders Used In the Aluminium Industry

The most widely-used binder for aluminium electrodes is coal-tar pitch from the high

temperature destructive distillation of coking (bituminous) coal. Pitch is the residue

obtained from the distillation of coal tar after all the light oils, intermediate fractions,

and heavy oils such as creosotes and anthracene have been distilled off. Pitches are

classified according to the carbonising temperature of the ovens or retons used in

preparing the tar, such as coke-oven tar, horizontal-reton tar and low-temperature tar.

Tars made by high-temperature carbonisation are generally more suitable for use as

electrode binders, because a high carbonising temperature corresponds to a high coking

value and carbon/hydrogen ratio, which is associated with high electrode strength.

Binder Properties

The functions of the binder are:

(1) To plasticise the coke powder, allowing the correct shape of electrode to be

formed.

(2) To provide, by carbonisation, a dense body of baked carbon in the electrode.

(3) To bind the filler particles together to provide a mechanically strong artifact.

The binder must lose its fluidity and produce an irregular macromolecular structure

which penetrates into the irregularities on the filler panicle surface. As the pitch

solidifies, it must transform successfully into a binder coke with the required propenies

of high mechanical strength and low electrical resistivity.

Clearly, it is imponant to identify the factors which affect the quality of the bonding

achieved between binder and filler. Because a number of inter-related factors affect the

propenies of the pitch, there are diverse views in the literature regarding the optimum

raw materials and conditions used in pitch production and electrode manufacture. This

reflects the empirical nature of the process.

Binder Specifications

Binders for carbon and graphite electrodes must meet the following requirements:

21

(l) High carbon yield, usually 50-60 wt % of pitch.

(2) Good wetting and adhesion properties to bind filler particles together.

(3) Suitable softening behaviour at the mixing temperature.

(4) Minimal amounts of impurities.

(5) Low cost, wide availability, consistent from batch to batch.

Attempts have been made, by adding carbon black or mesophase to the pitch, to

increase the QI levels and coke yield [24], but such modified pitches have not proved to

be satisfactory binders because the additives do not wet the filler particles and hence

contribute nothing to the binding ability of the pitch. The carbon yield can be increased

by the addition of sulphur or nitro-aromatic compounds, but binder graphitisability is

decreased.

Although coal-tar is still the main source of binder pitch (from carbonisation of coal in a

coke-oven), other sources, chiefly petroleum pitches, have been used. Synthetic resins

such as furfuryl alcohol may be used as binders for special-grade graphites, but are

unsuitable for use in electrode production because they increase resistivity [25].

Coal-tar pitches with a high quinolene insoluble (QI) content produce highly-ordered

lamellar coke whether carbonised alone or in an anode mix [26]. There may be

considerably less effective bonding between binder and filler, and hence a reduction in

anode strength, if thin layers of mesophase are allowed to form around the ftller

particles [27].

Commonly employed tests used to characterise pitches used as binders in anodes for

aluminium production by the Hall process include softening point, ash content, coking

value and QI and toluene insoluble (TI) materials content [28]. Other tests are for

carbon yield, molecular weight, spectra of various kinds, elemental analysis, viscosity

and purity [27]. Some of these tests are concerned with the practical side of the

extraction process; for example, softening point determines the minimum mixing

temperature of the coke/pitch blend (usually 140-16O°C), while ash content affects the

purity of the aluminium. The aromaticity of pitch is an important consideration for

anode production; as described in the section on carbonisation, aromatic compounds

form coke more easily than non-aromatic compounds, so coking value can be used as a

measure of aromaticity.

22

There have been several attempts to correlate the properties of the binders and fillers

used in the manufacture of carbon and graphite electrodes with the properties of the

actual electrodes. with a view to achieving the ability to tailor electrode properties

precisely. The amount of primary and secondary QI. and B-resins (TI - QI) has a

significant effect [29]; an increase in primary QI at the expense of gamma-resins (the

toluene-soluble fraction) gives higher bulk density and strength. and lower resistivity.

than an increase in B-resins. while a higher content of B-resins results in a higher pitch

softening point. but lowers the binder pitch requiremenl However. an increase in the

amount of secondary QI increases the binder pitch requirement. with only a small

negative effect on other pitch properties.

Effect of QI Content on Pitch Properties

Since the presence of QI is beneficial in reducing exudation. but detrimental to the

formation of a good interface between binder and filler. a suitable pitch should contain

the optimum amount of QI. which has been reported as being about 10 wt % [30].

When co-carbonisation takes place within the restricted space of the interstices between

the filler particles in an anode mix. the QI particles tend to form areas of optically

isotropic material surrounded by lamellar carbon [21].

In summary. a binder pitch should have as high a coke yield and atomic C/H ratio as

possible. to give the artefact the highest bulk density and mechanical strength. whereas

a high B-resin content helps to give a lower electrical resistivity in graphite-based

artifacts. A pitch with a high primary QI content results in a hard. isotropic binder pitch

coke. with high thermal conductivity. high electrical resistivity and low thermal

conductivity.

Filler Properties

A considerable amount of work has been done on pitch properties. but less appears to

have been done on the influence of the filler particles. However. it has been shown that

the nature of the filler surface may play an important part in influencing the quality of

binder-filler bonding [31].

Filler properties which affect the quality of the resulting electrode include [32]:

(1) Roughness of the surface of the coke filler particles. This is important in

determining the degree to which the binder will wet the filler; if the binder cannot

penetrate the small irregularities of the filler surface. inferior bonding between

binder and filler may resull

23

(2) Chemical reactivity of the coke, which will be influenced by the presence of

surface groups such as 0, H, S and N.

(3) The degree of perfection of stacking of the graphite-like components of the coke.

The more highly graphitised the filler, greater the requirement for a high

proportion of non-polar pitch components.

(4) The rate of carbonisation of the fIller.

(5) Resistance to oxidation. The reactivity of the filler coke is important, especially at

the high temperatures of the Hall-Heroult cell.

(6) Porosity; whereas metallurgical cokes are almost impermeable to gases, carbon

anodes have a predominantly open porosity, which increases the area available

for oxidation.

24

2.3.3 Factors Affecting Electrode Quality

Importance of the Blnder·Flller Interface

The quality of the manufactured electrode cannot be specified simply by specifying the

qualities of the binder and filler; a key feature which influences the electrode propenies

is the interface between the binder and the filler. Desirable electrode propenies -

mechanically strong, compact and dense, low electrical resistivity and low chemical

reactivity - are more likely to be obtained if there is good contact between the binder and

the ftIler.

Binder-Filler Interface Quality

Several factors influence the quality of the binder-filler interface. If pitch containing

mesophase spheres is used as the binder, this may wet the coke more effectively than

just pitch [27], but the mesophase may go on to form shells round the filler particles

and prevent good contact with the binder [33]. It is therefore desirable to ensure that the

optimum amount of mesophase is present in the binder, as discussed in Section 2.3.2.

The viscosity of the pitch also affects the binder-ftller interface; if it is too viscous (a

rheological consideration), it will not be able to penetrate the surface of the filler'

particles, whilst if it is unable to wet the surface (a thermodynamic consideration) it will

not make good contact even if the viscosity is satisfactory [34, 35, 36,37].

When a pitch is carbonised alone, anisotropic liquid crystal-like spherical mesophase

units are formed, as described in Section 2.2.3. The manner of mesophase growth and

coalescence has been shown to play a key role in determining the size and shape of the

basic structural units in the resulting carbon; the growth and coalescence processes are

themselves dependent on the characteristic of the pitch and the conditions of

carbonisation [31].

If a petroleum pitch is both highly aromatic and carbonised with minimum agitation, the

result is a well-ordered, highly anisotropic carbon, with extensively aligned basal layers

arranged into lamellar structural units. However, the use of such pitches in electrode

manufacture is likely to lead to cracking, because of poor bonding between the binder

andftller.

The carbon in cokes from high-rank vitrinites is composed of lamellar components, the

lamellae being aligned circumferentially around devolatilisation pore surfaces. Cokes

25

from low-rank vitrinites are granular, and may be considered to be composed of

compressed mesophase units which have undergone complete carbonisation [38, 39].

Examination of fracture surfaces of coal-derived filler cokes using ion-etching and SEM

fractography indicate that parts of the surface may be regarded as being composed of

nodular structures 0.2-2Ilm in diameter, which may correspond to non-coalesced

mesophase units [40]. By contrast, the pore surfaces are observed to be quite smooth;

it may be that they are coated with a thin film of carbon deposited during coal pyrolysis,

the carbon forming lamellae aligned parallel to the pore surface.

Little or no bonding occurs between the lamellae of lamellar binder coke and the pore

surfaces of filler coke particles, or between granular binder coke and the pore surfaces.

However, strong bonding occurs between fractured filler coke surfaces and both

granular and suitably aligned lamellar binder cokes.

26

2.3.4 Industrial Scale Manufacturing· Methods

Synthetic graphites which are manufactured for use as aluminium electrodes are usually

made using a petroleum coke filler and coal tar pitch binder which are mixed and then

moulded or extruded into shape before being graphitised. The chemistry involved is

described elsewhere in this literature review; the whole process, from preparation of the

raw materials to despatch of the finished graphite, typically takes up to four months

[41).

The principal steps in the manufacturing process are [42]:

.J

(1) Preparation of raw materials.

(2) Mixing of materials

(3) Forming into green shapes

(4) Baking

(5) Impregnation

(6) Optional re-bake

(7) Graphitisation

(8) Finishing (cutting to correct shape)·

The Delayed Coking Process

This is the principal method of manufacture of the petroleum coke used as filler in

synthetic graphites. World production by this method was recorded as sixteen million

tons per annum in 1981 [43]. A wide variety offeedstocks is used, principally thermal

tar, decant oil, catalytic cracker slurry, coal tar pitch and ethylene cracker tar. This

ability to use a range of feedstocks is one of the reasons why the delayed coking

process is so popular.

The delayed coker feed is fed hot to the heater of the coker and heated to about 500°C.

The feedstock then flows into one of two coke drums where it thermally cracks to gas,

petrol, gas oil and coke. Over a period of hours, the coke accumulates in the drum,

while the light products pass back to the combination tower. When the first drum is

full, the feed is switched to the second drum and quenching of the lust drum

commences, l?"st by steam and then by water. Mter the water has been removed, the

now-cool coke is drilled from the drum using high-pressure water jets. (See Figure

2.2).

In the drum, coke formation takes place by polymerisation and homogeneous

nucleation of the feedstock molecules, to produce lamellar nematic liquid crystals. At

27

first, the system is highly fluid, and spheres may coalesce to form larger spheres, but

eventually, a fully-coalesced structure known as bulk mesophase is formed, beacause

increasing temperature or time leads to further polymerisation of the constituent

molecules. The bulk mesophase still shows plastic properties, and will undergo further

polymerisation before becoming a rigid semi-coke (with a heat-treatment temperature

of less than 500°C) or a green coke (with a heat-treatment temperature of greater than

500°C ).

Raw Materials

The chemical composition of the feedstock used in the delayed coking process largely

determines the quality of the coke. Feedstock constituents are usually divided into

asphaltenes, resins and aromatics. Asphaltenes are colloidal dispersions of hydrocarbon

molecules made up of C, H, 0, N, S, V and Ni; typical molecular masses are

3000-5000 amu [44]. Resins are similar, but with molecular masses 200-300 amu less

than the corresponding asphaltenes. Aromatics are composed of unsaturated but very

stable polycyclic six-carbon rings. Although different chemically, paraffinic and

naphthenic molecules are often included in this category.

Feedstocks with high asphaltene and resin contents are generally unsuitable for coke

maufacture because they produce a highly cross-linked coke with a large concentration

of impurities. Feedstocks with about 50 wt % aromatic constituents are suitable for the

production of coke for electrodes used in aluminium extraction. Feedstocks with greater

than 70 wt % of aromatic constituen.ts are suitable for the production of the

high-quality, easily graphitised coke known as needle coke, which is used in electrodes

for steel production and in certain electrolyic processes.

Cokes to be used in the manufacture of aluminium electrodes must have a low ash

content, because V, Ni, Fe and Si reduce the quality of the aluminium A relati,vely high

sulphur content is less important. Conversely, graphite electrodes must be made with

low-sulphur needle coke, because a filler with higher than 0.8 wt % sulphur is likely to

swell as the sulphur is eliminated as a gas. This swelling, known as puffing, at least

degrades artifact properties (decrease in bulk density, mechanical strength, electrical

conductivity and thermal conductivity) and may even result in gross failure of the

artifact.

So far, properties of filler coke have been considered, but the properties of the binder

are also important. As described in the section on binder and filler properties, the binder

must plasticise the filler to allow the artifact to be formed into the correct shape, and

subsequently carbonise to form a coke linking the filler particles, to provide both a

28

strong and dense artifact with the required thennal and elecoical conductivities.

Calcination

In general, green coke (that is, coke which has not been heat-treated and which contains

a high proportion of volatile matter) is not suitable for use as a filler coke because it

tends to shrink on heat treatment to a degree dependent on its volatile matter content

[45]. Because of this, green coke is generally heat-treated before being used as filler;

the heat treatment process is called calcination.

During calcination to 1400°C, the volatile matter content of the coke is reduced from

about 5-15 wt % to about 0.5 wt %, the volatiles mainly being released as gases,

including CH4, C2H6, H2, H2S and CH3SH. The C/H ratio of the material is

increased from about 20 to greater than 1000 [46]. The volatile matter content decreases

with increasing calcination temperature. Calcination is necessary because:

(1) Green materials cannot easily be bound to produce an electrode of the correct

density.

(2) They are difficult to mould or extrude.

(3) They emit volatiles during baking, resulting in a very porous electrode.

(4) They have a much higher elecoical resistivity than calcined materials.

Three types of calciners are in general use for cokes: rotary kiln, rotary hearth and

vertical shaft, the first being the most widely used.

Calcination usually results in coke filler particles 5-8cm in diameter. Particles about

1.3cm in size are required for aluminium anodes. Three opposing principles act when

the size of filler particles is decided. It is desirable to reduce porosity (and increase

strength) by packing small particles between large particles, but it is also desirable to

provide adequate voidage to enable the volatile pyrolysis products from the binder

phase to escaping during the subsequent baking operation, to avoid structural damage.

Also, the sizes of pores may need to be controlled.

Mixing

Manufacturing commences with mixing of the pitch and filler, which improves the

29

likelihood of obtaining a strong product. Mixing must be done at a sufficiently high

temperature to keep the binder fluid, and gently enough to avoid breakage of the filler

particles. The amount of binder must also be chosen carefully; there must be enough to

exclude air from the voids between the filler particles, taking into account that some

binder will penetrate some of the pores in filler particles. Following the mixing

operation, the hot mix is cooled to slightly above the softening point of the pitch, to

ensure its rheological properties are suitable for the forming operation. Total time for

mixing and cooling is typically 1.25 hours.

FormIng

The next stage is the forming process, which increases the density of the mix, resulting

in close contact between the binder-coated filler particles, a low porosity and a size and

shape of the so-called "green" artifact as close as possible to the finished product. The

two principal methods of forming are extrusion and moulding. Extrusion is the newer

and more widely used of the two methods. There are several types of extrusion

equipment, but all involve the use of a hollow cylinder known as a mud chamber, into

which the mix is placed; pressure is then used to force the mix through a die of the

correct shape. The extruded material is then cut to the correct length and cooled slowly

to avoid cracking.

The extrusion process produces a bulk anisotropy in the product; the filler particles

become aligned with their long axis parallel to the direction of extrusion. The degree to

which this takes place depends on the type of fIller and the details of the extrusion

process, and is more pronounced for needle-coke fillers. A consequence is that

extruded products have higher strength, Young's modulus and thermal conductivity in

the direction of the grain, while the electrical resistivity is smaller.

BakIng

The green artifact is subsequently baked, by being heated from 800°C to lOOO°C. This

stage has two functions: to convert the binder into solid coke and to remove shrinkage

in the artifacllfbaking is carried out too rapidly, the fmished product is likely to be of

poor quality; since the green artifact is almost impermeable, the gradual development of

a venting porosity early in the bake is essential to avoid a highly porous or cracked

structure. Also, too high a rate of heating will result in differential shrinkage, leading to

cracking of the artifact. A particular problem in this stage is the complete loss of

mechanical strength between 200-400oC because the binder is liquid. To avoid

slumping and distortion of the artifact, it must be packed in coke or sand which must

not only provide mechanical support but must be sufficiently permeable to allow

30

volatiles from the pitch to escape. Aluminium electrodes generally require two weeks to

fIre and fIve days to cool. while speciality graphite may require six weeks to fIre and

four weeks to cool [42).

Cleaning and Impregnation

A cleaning stage may constitute the conclusion of the manufacturing process for

aluminium electrodes and blast furnace refractories. For artifacts that are to be

graphitised. the next stage is impregnation. which reduces the porosity and increases

the density of the finished graphite. If the impregnant has large concentrations (greater

than about 3 wt %) of QI materials. it may not reach the centre of the product and a

so-called dry core will result. which may lead to cracking or gross failure of the

artifact. Commonly used impregnants are low melting point coal-tar or petroleum

pitches with less than 3 wt % QI [42).

Graphltlsatlon

The artifact is now ready for the actual graphitisation process. an electrical heat

treatment to about 3000°C. (See Section 2.2.6 for details of the chemistry and physics

of graphitisation.) The Acheson furnace. invented in 1895. is capable of heating several

tons of charge to temperatures approaching 3000oC. The furnace bed consists of

refractory tiles. The furnace ends are made of concrete; through them project several

water-cooled graphite electrodes which are connected to the secondary of a transformer.

The product is placed on a layer of metallurgical coke; its long axis may be placed

parallel to. or at 90° to the direction of current flow. depending on the manufacturer. A

coarse-sized metallurgical coke. called resistor pack. is used to fill the gaps between the

pieces to be graphitised. The furnace is then covered with a blend of fmer metallurgical

coke. sand and silicon carbide to provide thermal and electrical insulation.

A furnace of this type usually requires a day to pack. uses a heating rate of 4O-60°C per

hour. a total firing time.of about three days and a fInal graphitising temperature of

2700°C [47). Cooling and unloading takes eight to ten days; the resistor pack must not

be removed while the product is at a high temperature. to prevent oxidation taking

place.

Recent Developments

Demands for graphite with isotropic bulk properties. and for more rapid production.

have led to the development of more modem techniques such as isostatic hot pressing

31

(IHP), artifact densification and high pressure impregnation [48].

Isostatic hot pressing is a high-temperature, high-pressure compaction process, in

which the binder/filler mixture is put in a hermetically sealed metal container, which is

then placed in an autoclave. At high temperatures, the container becomes plastic and

and the pressure in the autoclave is transmitted to the mix.

Anifact densification is similar to IHP, except that the charge is a pre-fonned solid.

Very high temperatures (greater than 2200°C) are required to produce plastic

deformation of the solid carbon.

High pressure impregnation/baking is again similar to IHP, and incorporates the

impregnation and baking stages into the same operation. The impregnation stage takes

place at atmospheric pressure and the baking stage at approximately lOOMPa.

So-called "binderless" graphites may be used where very high strength and high

density are required; these are manufactured at very high temperatures and pressures,

typically 1700-2700°C and lOO-120MPa. The highest strength graphites are obtained

using the higher temperatures.

32

2.4 Strength of Brittle Materials

2.4.1 Basic Theory

Deformation of most solids occurs by the atoms or molecules of which the substance is

composed moving further apart or closer together, against the resistance of their

chemical bonds, which acts to keep the atoms at the most energetically favourable

distance apart. This distance is determined by the relative strengths of attractive and

repulsive forces between atoms.

Since the bonds between atoms are very strong, it requires considerable force to

compress (or extend) a block of steel in this way; even then, the movement of the block

is on a microscopic scale. There are special cases, such as rubber, where the molecules

are arranged in large-scale helical structures, which are deformed when the material is

compressed or stretched. Elastic materials (those which recover their original shape

when the applied load is removed) include not only rubber, but steel, concrete and

many other substances to which the word elastic is not generally applied.

Some substances, however, are not elastic; they do not recover fully (if at all) after

being deformed. These substances are termed plastic, and include examples such as

putty and plasticine.

Surface Energy

Surface energy in a solid is analogous to the surface tension in a liquid; energy is

required to create new surfaces, because chemical bonds are broken during the process.

Surface energy may be considered to be a measure of the resistance of the material to

crack propagation, or the toughness of the material. For materials with high surface

energies, the terms fracture energy or work of fracture are sometimes used. For most

materials, theoretical surface energies are approximately 1 Jm-2. It is found in practice

that surface energies from fracture stuclies vary widely:

Material Surface Energy, J m· l

Glass 1-10

Brick 3-40

Nylon 1000

Wood 10,000

Steel 100,000

33

The large values result from the consumption of energy in processes other than those

directly involved in failure; the low values are observed for brittle materials where

deformation of the StruCTure is only encountered over a few atomic layers.

34

2.4.2 Fracture of Brittle Materials

An application of the principles of brittle fracture theory is necessary in both the

traditional type of engineering structure, and in new materials of various types. The

latter category includes carbon used in load-bearing structures or in applications such as

aluminium electrodes where the size of the structure means it must bear a load as a

consequence of its size.

When a force is applied to a structure, it will deform to some extent, depending on

certain properties of the material of which the structure is composed. The terms which

describe the amount by which a substance defonns are stress and strain.

Stress is the load on a structure, typically measured in meganewtons per square metre

(MNm-2). The stress in a solid is analogous to the pressure of a gas, that is, it is a

measure of the external forces on the atoms of the material. When the material is not

subjected to a stress, the atoms are at their neutral or strain-free position [49].

Strain is the amount by which a structure will defonn under a load, per unit length of

the structure, that is, the fractional extension:

e = amQunt of stretch

original length

Strain can therefore be considered as a measure of how far the atoms in a solid have

been pulled away from their most thermodynamically stable positions.

If the material is stretched, the amount of stress increases. Eventually the bonds will

be unable to withstand any more stress, and will break; this is seen on a large scale as

the structure failing. Some of the energy released in the process of breaking bonds is

released as heat, hence the fractured ends of an object are often felt to be warm to the

touch. Some energy may also be released as sound, hence the snapping noise which is

heard when a brittle material breaks.

In 1807, Young realised that a consideration of the material of which a so:ucture was

composed, rather than the structure itself, led to the concept of stiffness.

Young arrived at the relationship:

at ress

strain

constant (stiffness of the material)

35

This constant for a material is known as Young's modulus for that material, and has

the units of stress (since strain is a dimensionless quantity). The initial region of the line

obtained by plotting stress against strain is straight; in this region, materials behave

elastically, and the gradient of the line gives the value of Young's modulus for that

material, which may be considered as the stress needed to produce 100% strain, or as a

measure of the stiffness of the material. The area under the line gives the strain energy

stored in the piece of material being tested. For a brittle material, the stress/strain curve

is virtually linear up to the point of fracture; this is not the case for other classes of

material.

Much of the pioneering work in this field was done by Inglis, whose first major

contribution to this field was to realise that although a material might appear to be

strong enough, the structure of which it was constructed could fail if there were locally

high concentrations of stress.

In a classic paper [50] Inglis showed that if, for example, a third of a structure was

removed, the stress at the edge of the gap such as a hatch cover in a ship was not

simply 4/3 the average but could be many times as high. The situation was worsened if

the gap formed a sharp re-entrant (that is, a square opening would be worse than a

circular one), and if the material was brittle.

Stress concentration at the tip of a crack approximates to:

s = -h/r

where 1 = crack length

r ::. tip radius

Inglis used the example of an elliptical hole in a flat plate in his calculations, and

showed that small, even SUbmicroscopic, flaws might be sources of weakness, since

the limiting case of an infinitesimally narrow ellipse is a crack. Inglis also realised that

the amount by which the stress is concentrated depends on the shape of the hole rather

than its size.

Inglis had shown that tough materials (those which can withstand relatively high tensile

stresses without fracturing) should be used when the structure of which they were

made was to be in tension. Brittle materials could be used .in situations where they

would be in compression; in tension, the defects present could spread, with failure of

the structure.

36

This work had provided a much clearer picture of what happened when a fracture

occurred, but not how it happened. Several important questions were left unanswered,

including why large cracks were found to propagate more easily than small ones, and

the effect of the radius of curvature of the crack tip.

The classic work of Griffith extended these ideas and described them more formally, by

introducing the energy-balance concept of fracture, which applies to fracture in general,

not just to brittle fracture.

Until Griffith's work, it was generally thought that all materials fractured at a critical

stress level which was characteristic of the material in question. The inadequacy of this

theory was shown by the observations that the fracture strength of a given material

often varied with the temperature, rate of loading or other factors.

Griffith's view was that an applied load causes strain energy to be stored within a

material; whether the structure breaks depends on whether or not it is energetically

favourable for the strain energy to create new crack surfaces.

From a consideration of the case of an isolated crack in a solid which was subjected to

an applied stress, Griffith suggested that brittle fracture occurs by the rapid growth of

such a crack through the material. However, this did not explain why the crack

extended, rather than the material deforming elastically or plastically.

Griffith formulated a critical crack criterion for the conditions, in terms of classical

mechanics and thermodynamics, under which the crack would extend. He suggested

that a body with an internal crack simply sought the lowest free energy configuration,

which would leave the crack in equilibrium and hence on the point of extension. A

small applied load would then supply the necessary energy for rapid crack propagation

to occur [51]. The equilibrium would be between the mechanical energy of the system

(work done by the applied load + strain potential energy of the medium + the energy

required to create new fracture surfaces) and the surface energy. The mechanical energy

must decrease as the crack extends; if this was not the case, the crack would stop.

Conversely, the surface energy must increase as the crack extends, as more bonds are

being broken.

Griffith's criterion is that the surface energy in the total areas formed by crack

propagation must be equal to the elastic strain energy which was stored in the region

before the crack spread. A surface crack therefore acts as a stress concentrator, so the

stress in the material near the tip of the crack is much greater than the applied stress.

37

The ~eeper the crack, the lower the stress needed for for the crack to propagate. Hence

the very high strength of carbon fibres and glass fibres is explained in terms of their

small surface area, which precludes the existence of very deep cracks.

According to Griffith, the stress concentration effect described by Inglis is simply a

mechanism for converting strain energy into fracture energy; fracture only takes place if

there is enough strain energy available. The basis of the Griffith critical crack criterion

is that while the energy debt of a crack increases linearly with the crack length, its

energy credit increases with the square of the crack length, since (to a reasonable

approximation) two triangular areas on either side of the crack will give up strain

energy. Up to the critical point, there is a net absorption of energy. Beyond that point,

energy is released at an increasing rate.

A crack or other defect may therefore exhibit a local stress concentration which is much

higher than the nominal tensile strength of the material, but failure will not occur

provided that no crack is longer than the critical length. In practice, however, all cracks

must start as being short, and subsequently extend.

Griffith formulated the critical condition for fracture as:

6f = (2"(E / ltc) 1/2

where 6f = failure stress

y = surface energy (Jrn-2 )

E Young's modulus (units of stress~ Nm-2 )

2c = critical crack size

This assumes a constant load rate and stress in one plane only.

Hence the critical crack length is both a necessary and sufficient criterion for fracture.

Griffith realised that the maximum stress at the tip of a crack at equilibrium must

correspond to the theoretical strength of the solid, that is, the strength of the chemical

bonds between the atoms or molecules of the material. It had been noted by previous

workers that solids always fell short or"their theoretical strengths by up to two orders of

magnitUde, and that when fracture occurred, it did not do so by an explosive separation

of all the constituent atoms, but along one plane. From these observations, Griffith

deduced that microscopic or submicroscopic flaws were present in all materials, and

38

that a thin specimen would be stronger than a thick one made of the same material

because it would have less room for defects.

It is found in practice that:

c = l/X . Work of fractUre per unit area of crack surface

where

Strain energy stored per unit volume of material

c = zm;

1t52

2c length

W work of

E Young's

5 = Average

of Griffith critical crack (m)

fracture (Jm-2 )

modulus (Nm-2 )

tensile stress in material near crack (Nm- 2 )

On the basis of the Griffith theory. the main safeguard against the occurrence of

brittle fracture is a high work of fracture. The work of fracture of a material is the

quantity of energy required to break a piece of material (of a set shape), by breaking

all the bonds in a plane and thus creating two new surfaces. Hence a material which is

tough has a high work of fracture, whilst a material which is brittle has a low work of

fracture; it is not brittle because it does not take much force to break it, but because it

does not take much energy. Materials may have similar tensile strengths, but may

differ greatly in values for work of fracture.

Summary of Grifflths' Treatment of Brittle Fracture

Griffiths recognised the imponance of the stress-raising capacity of existing flaws in

a material, and deduced that crack propagation occurred when the strain energy

available exceeded the surface energy of the two freshly-created surfaces. Since, for a

crack in the centre of a flat plate, the strain energy is proportional to the square of the

crack length while the surface energy is directly proportional to it, it follows that

above a critical crack size more strain energy is available than is required for the

creation of fresh surface. Therefore, the propagation of a critically-sized flaw leads

directly and inevitably to failure. In real materials, a considerably greater amount of

mechanical energy may be needed because the material exhibits some non-brittle

behaviour.

39

Stress Intensity Factors

An alternative approach to the theory of fracture of brittle materials involves

consideration of stress intensity factors, where fracture is characterised by a stress

intensity factor attaining a critical value. This approach leads to the conclusion that a

test piece with large cracks will fail at low stress levels, while a test piece with high

stress levels will fail at higher stress levels. If the applied stress does not exceed a

threshold value, crack propagation will not occur. If the critical value for the stress

. intensity factor is reached, failure ensues directly. For stress intensity factors between

the threshold and critical values, crack growth occurs until the critical value is

exceeded and fracture takes place.

Crack Stopping

A propagating crack may be stopped if it encounters conditions where the stress

concentration is reduced. In materials such as a carbon electrode, or composite, this is

most likely to occur at an interface. Because the interface adhesive strength will be

much less (perhaps 1/5) of the cohesive strength of the solid, an interface lying ahead

of a propagating crack will be broken before the crack reaches it. When the crack

does reach it, the crack tip radius will be greatly increased, and propagation is likely

to stop. In anode/cathode blocks, weak binder-filler interfaces may act as crack

stoppers and also act to lower the surface energy.

40

2.5 Strength of Carbons and Graphites

2.5.1 Brittle Fracture Theory and Carbons

Although it is convenient to consider the tensile failure of carbons in terms of Griffith's

theory, such materials do not strictly conform to classical brittle fracture theory. It has

been found that for metallurgical cokes, microcracks are developed at pores at low

stress levels [52]. Such cracks become stabilised, usually by being blunted on entering

another pore (see earlier). Eventually, however, a critically-sized flaw is formed and

failure ensues. A similar mechanism has been suggested for the failure of graphites

[53]. The detection of acoustic emissions from graphites at pre-failure stress levels [54]

has been associated with the generation of stable microcracks. Consequently, the

fracture energy is greater than the energy required to create the fracture surface.

Numerous factors influence the strength of cokes. One is the development of pores

during the carbonisation process; it has been stated that the key to understanding and

controlling the strength of metallurgical cokes lies in characterisation of the coke

porosity, since the pores are several orders of magnitude larger than any of the features

associated with carbon matrix, and are thus likely to correspond to the Griffith critical

cracks which may cause fracture of the material [55]. On this basis, any characterisation

of the porous structure from a strength point of view must take into account the size of

the larger pores and an indication of their shape, to enable the effects of stress

concentration to be determined. From the work of Griffith, the strength of the coke

should be related to maximum pore size by an inverse square root relationship. Patrick

and Stacey considered that to increase the strength of the coke, it is necessary to lower

the maximum diameter of the larger pores, and the volume porosity. Also, pores should

be as nearly spherical as possible.

For metallurgical and other cokes, the use of an equation of the form S.N = k +a.Wtp2

has been suggested [56], where S = tensile strength of the material in MPa, N = number of pores per mm2, W = mean wall size (thickness of carbon matrix) in ~m, and

P = mean pore size in ~m. The values of the constant k and coefficient a depend on

the raw materials used.

Recent work suggests that the same principles may be applied to electrodes made using

dectro-calcined anthracite as a filler [57]. Also, volume porosity and pore size

increased markedly with increasing heat-treatment temperature from lOOO°C to 2500°C.

The results suggested that pore structural data can be used to predict the tensile strength

of the material, and that HIT above lOOO°C led to a change in the mode of breakage.

41

Further work done in this area [58] indicated that the porosity of ECA or GCA

(gas-calcined anthracite) increased with an increase in HIT to 2500°C, and that

ECA-based materials differ somewhat in their characteristics from graphite and

GCA-based materials. Fractography using the SEM provided further indications that

this was the case; ECA-based materials showed a high incidence of interfacial failure,

while the other materials appeared to be well bonded, failure occurring predominantly

by a transgranular mode, implying that different parameters control the different modes

of failure.

Subcrltical Crack Growth In Carbons and Graphites

Subcritical crack growth is frequently found to precede mechanical failure in graphites

used in structural, that is, load-bearing, applications [59]. The easiest direction for slow

crack growth to occur is along the plane of the lamellae, while the most difficult is

perpendicular to the plane of the lamellae. Primary crack extension has been modelled

to occur by crack propagation along lamellae or fissures in filler particles. Secondary

cracks may form in advance of the main crack, perhaps initiating at pores.

Extensive SEM fractography [59] indicates some general micromechanical features

common to a variety of graphite materials:

(1) Although the fracture path intersects pores of all sizes, it does not go out of its

way to intersect pores.

(2) Microcrack density increases, due to the crack tip stress field, appear to be

minimal.

(3) The cracks and fissures observed in some petroleum coke fIller particles are

structural features which probably result from thermaL stresses during the

heating and subsequent cooling processes.

In the longitudinal direction of petroleum coke based artefacts, the fracture crack

frequently appears to be diverted around ususually-orientated filler particles, and

propagates through the binder phase. In the transverse direction, however, fractures of

filler particles are observed to be fairly numerous. Also, the tortuous crack path

indicates that the crack makes frequent detours around filler particles. Whether the crack

goes through or round a particle appears to depend on both energy considerations and

particle orientation.

42

--------

Toughness of Particles

The least tough fonn of solid carbon is the single-phase. glassy type. The most tough is

the single-phase. highly-crystalline pyrolytic graphite. in the direction perpendicular to

the basal planes. Two-phase graphites show intennediate toughness. depending on the

relative proportions of the two phases. the degree of anisotropy seen. and the porosity.

A further significant factor is the relative crystallinity of the two phases. the filler being

crystalline and the binder less so.

43

2.5.2 Porous Brittle Materials

The porosity of brittle materials such as carbon must be taken into account when

considering their strength. Early work in this area was done on ceramics, as stated by

the Ryshkewitch-Duckworth equation:

5 = 50 exp(-bp)

where 5 strength of a porous body

50 = strength of a similar non-porous body

p = fractional porosity

b constant

Knudsen [60) undertook a theoretical consideration of a material composed of sintered

spheres, assuming that the variation of strength with porosity reflected changes in the

load-bearing area within the specimen, the critical area being that traversed by an

irregular cross-section passing through the areas of contact between the spheres.

Knudsen continued his work by combining the Ryshkewitch-Duckworth equation with

an equation derived by Orowan [61):

5 = kG -1/2

where G = grain size

to derive an equation which took into account both grain size and porosity:

5 = 50G-1/2 exp(-bp)

where b is a factor which decreases as pores become more rounded.

It is possible to combine this equation with Griffith's equation to derive a more general

form of equation for the strength and porosity of porous brittle materials:

5 (2EY Iltc) 1/2. exp (-bp)

Simplified versions of this equation have been successfully used, which indicates that

some factors in it remain constant between different materials.

44

2.6 Electrical Resistivity of Carbons and Graphites

2.6.1 Introduction

It has been demonstrated that carbon powders are ohmic resistors [62], and that a

strong dependency exists in the range resistivity and pressure, such that only at fixed

and constant pressures is a comparison in the range resistivity values meaningful. A

series of straight lines were obtained by plotting a graph of m V against mA for different

pressures. Resistivities were found to range over three orders of magnitude, in the

order:

Electrographite < Petroleum coke < Acetylene black < Carbon black

Although carbon films are not intrinsic semiconductors, it has been observed that films

50nm thick can show the propenies of a semiconductor in the temperature range

277-397 cC, where electrons can be excited across the forbidden band and thus produce

non-localised electronic conduction [63]. At low temperatures, thermally-assisted

tunnelling through potential barriers may take place.

2.6.2 Structural Aspects

With heat-treatment temperatures in the range 400-500°C, it has been observed that

there is a decrease of resistivity and Young's modulus [64]. The reasons for the former

are unclear, but may reflect changes in the binder such as coalescence of the spherical

liquid crystals into bulk mesophase. The corresponding changes in Young's modulus

are more difficuitto account for.

A sharp decrease observed in the electrical resistivities of coals and a charcoal at

temperatures in the range 500-800°C may be accounted for in terms of the volume

fraction of the graphitic panicles distributed among microvoids [65]. Between

200-300°C, there appeared to be a small increase in resistivity, possibly due to the

formation of voids as volatile compounds are removed. With HITs in the range

500-800°C, resistivity decreased by twelve orders of magnitude. A coal with a higher

inorganic content showed a more gradual decrease in resistivity.

It has been noted [66] that for pitches at temperatures in the range 60-l60°C, electrical

conductivity and temperature susceptibility both changed with a change in temperature.

45

fonnation at 450°C is accompanied by a rapid increase in pitch resistivity. which

subsequently levels off and remains nearly constant until coalescence takes place. After

colaescence takes place. there are indications that current flow takes place at least partly

by solid state semi conduction.

46

..

2.7 Literature Review - Conclusions

This literature review has attempted to show that a number of factors play a part in

determining electrode properties, and that several of these factors are inter-related

Although the carbonisation process is well understood in general terms for

laboratory-scale experiments using simple precursors, there are still many aspects

which are poorly understood when considering the process on a large scale, using

complex precursors. An alternative, and complementary, method of studying

carbonisation is to note that a particular substance, or conditions used, gives an

electrode with desirable properties, and use this data both as a basis for electrode

manufacture and to advance the theory involved. The relative lack of detailed

knowledge of carbonisation on a large scale is reflected in the methods used in

industrial-scale electrode production, which work well most of the time, but may not

always give consistent results from batch to batch. It would be preferable to be able to

be confident that electrodes with the correct properties could be manufactured

repeatably, on the basis of a better theoretical understanding of the process. Since the

porous structure of electrodes plays an important part in determining their strength,

further understanding of pore development is a particular aspect of carbonisation which

would be likely to be useful in designing methods for producing mechanically strong

electrodes on a repeatable basis.

A reasonable amount of work has been done on the pore structural characteristics of

polished electrode surfaces, but there is scope for further work in this area, possibly

using more sophisticated image analysis techniques.

SEM study of fractured electrode surfaces has yielded a considerable amount of data on

modes of fracture of electrodes, but there is scope for further work in improving the

techniques used.

Although much work has been done on the electronic properties of carbon and graphite,

very little work has been done on the electrical properties of carbons used as electrodes,

in relation to their microscopic structure. It is likely that such investigation would be

rewarding both from a theoretical point of view and from the point of view of helping

to determine what factors affected the resistivity of the electrode.

The most significant deficiency in studying the diverse factors which influence the

mechanical and electrical properties of carbon electrodes has been the lack of relevant

data to enable the inter-relation of the factors to be understood. For example, as

47

mentioned above, a considerable amount of work has been done both on the pore

structural characteristics of electrode materials and their modes of fracture, but little

work has been done on relating these two groups of data for the same electrode

material. This research project was intended to obtain data for tensile strength, pore

structural parameters, modes of fracture and values of electrical resistivity for several

carbon electrode materials, and relate them to the overall properties of the electrodes.

Also, a method of detennining the wetting temperature of a particular combination of

binder and filler was to be designed, to provide an indication of the readiness with

which the binder will wet the filler during the initial mixing stage in electrode

manufacture.

48

." IQ c: ... CD I

!'l .....

en ()

:::T T CD

3 L III

o Siphon adle ...

()

~ C '" III

IQ ... III 3

.-~

-

~ r--

0 «"" ',". ". -::I: III

I

::I: CD ... 0 c: ... 0 CD

- - - -

f-- f-- r--

'. . --:~ "

. ..... '"

Carbon Cathode Lining

Insulation

-a

B

Anode Conductor

aked Carbon ode An

M

,i~1 El

olten uminium

olten ectro Iyte

,/~

a

V~

Cathode Conductor

teel asing

"'11 ca c ... (1)

N N

en o =r (1)

3 III -o

C III ca ... III 3 o -c (1)

III '< (1)

Co

o o 7:" (1) ... o

"C (1) ... III -o ::::I

Closed

0> >

:;::: o I'll c

700K

Vapour valves

Coke drums

0> > .... o «

Open Combination

tower

Vapour line

Coking heater

Top reflux

645K

Residual feed

Wet gas

310K

Compressor

Wild gasoline

475K

Light coker gas oil

575K

Heavy coker gas oil

3 Experimental Procedures

3.1 Electrode Materials Used

The seven carbons examined all used coal-tar pitch as a binder. Five used petroleum

coke as a filler, one used gas-calcined anthracite and graphite and one used

electro-calcined anthracite.

Carbon A4 was a carbon used for the refractory lining of blast furnaces. The fIller was

a mixture of gas-calcined anthracite (GCA) and graphite, and the binder was coal-tar

pitch.

Carbon C was another carbon refractory; the filler was electro-calcined anthracite

(ECA), and the binder was again coal-tar pitch.

Carbon P was a low-ash chemical-resistant carbon. The filler was petroleum coke and

the binder was coal-tar pitch.

Carbon AL6 and Carbon ALS were anode carbons from the aluminium industry. The

fIller in both was petroleum coke and the binder was coal-tar pitch. These carbons were

selected from a batch of electrode carbons which had been manufactured in the same

way but which differed in mean tensile strength.

Carbon Y was an ordinary aluminium anode carbon.

Carbon Z was an aluminium anode carbon which was baked at a higher temperature

than usual.

3.2 Tensile Strength Determination

The tensile strength of the electrode carbons was determined using the

diametral-compression test, in which the compressive load required to break a

cylindrical specimen across a diameter was measured [68]. Cores approximately J.5cm

in diameter were drilled from blocks of the carbons to be examined. The cores were

then sliced into specimens approximately lcm in length. These specimens were

washed, cleaned ultrasonically and dried. They were then inspected visually and any

damaged ones were rejected. 50 specimens for each carbon (except where there was

insufficient material) were weighed and used in the determination of tensile strengths,

which was carried out using an Instron Universal Testing Machine. The load required

51

to break each specimen at a cross-head speed of 5mm min-J was recorded, and the

mean tensile strength, standard deviation and standard error for each carbon were

calculated. The apparent density of each carbon was also calculated from core weight

and dimensions.

3.3 Pore Structural Analysis

Pore structural analysis was carried out using a Cambridge Instruments Quantimet

Q800 system, in which a television image obtained from a polished, resin-impregnated

carbon surface is viewed under reflected light. The individual dots of which the image

is composed are allocated to one of 64 grey levels, according to their brightness. The

image is then converted into a black and white image in which all pixels are allocated to

be black or white respectively, depending on whether their grey level is above or

below a threshold brightness. This detection level is set manually, such that the dark

pores in the detected image correspond to those seeri in the TV image.

Six specimens were selected from the tensile strength fracture specimens in such a way

as to use specimens covering as wide a range of tensile strengths as possible, to try to

find evidence of pore structural differences between specimens of differing tensile

strengths.

The selected specimens were embedded in araldite and polished using successively

finer grades of pad and abrasive; a1umina was used as the abrasive for the initial stages,

and diamond paste was used for the fine polishing. The objective was to attain a

scratch-free and relief-free flat-polished surface. They were checked for scratches under

an optical microscope between each stage of the polishing process, and polished further

as necessary.

The field parameters of interest are the volume porosity, number of pores, pore size and

wall size. These are known to be significant factors in determining the properties of the

carbon artifact. The volume porosity is'obtained by expressing as a percentage the ratio

of the black area to the total area of the measuring frame. The number of pores is the

count of features whose lowest point lies within the frame. Wall and pore sizes are

obtained using the number of intersections of the scan lines with the trailing edges of

pores. The pore size is then obtained from the ratio of the black area to the product of

the number of intersections and the pixel size, while the wall size is obtained in a

similar manner, but using the white area. The process is repeated for vertical scan lines,

and the mean of the values obtained in two directions used.

52

lava rl/N

where Iav = average intercept

r I a total intercept

N = number of features

I represents the number of intersections of scan lines or columns of the picture matrix

with image detail. Pore size is obtained from the total lengths of horizontal and vertical

intersections of scan lines with image detail.

Chord length (pore size) petected area (Equation 3.3.1)

r I

Closest approach Total area - Detected area (Equation 3.3.2)

r I

To determine the number of fields to be used for each specimen, measurements were

taken from each of the two halves. It was found that the accumulated mean value for

each parameter levelled out at about 40 fields; consequently, it was decided to use fifty

fields for each specimen, divided into groups of ten for convenience.

Three of the groups of ten were taken from one half-specimen, and two from the other.

To set the detection level, the detection level was determined for five fields for a

half-specimen, and the mean value calculated. This value was used as the detection

level in the accumulation of data for ten fields, then the procedure was repeated for each

subsequent group of ten fields. Although more accurate, it would have been too

time-consuming to set the detection level for each field.

At the start of each session, the instrument was calibrated using a microscope slide on

which were a group of circles of known size. If their size (measured along horizontal

and vertical diameters) corresponded well to the known size, the carbon specimens

were then exanlined.

3.4 Fractography Using Scanning Electron Microscope

A method of examining fracture surfaces from tensile strength experiments was

developed to enable the predominant mode or modes of failure of the carbons to be

determined. It involved use of a scanning electron microscope (SEM) to identify and

subsequently categorise features observed on fracture surfaces, calculating how many

53

sampling points would be needed to give an accurate view of a panicular carbon, and

carrying out the counting. Details of the method are described below.

The morphology of single filler particles was studied before fracture surfaces were

examined. After the different types of individual panicles had been identified, extensive

preliminary examination of fracture surfaces was carried out in order to be able to

categorise features reliably. A selection of electron micrographs is shown at the end of

this chapter to illustrate the types of particles and surfaces seen, in Figures 3.2 - 3.7.

Classification offeatures identified in fracture surfaces

Name of Feature

Filler fracture

Binder fracture

Filler side of interface

failure

Binder side of interface

failure

Highly porous ftller

Highly porous binder

Initials

FF

BF

IF

IB

HPF

HPB

54

Description

Observed where the fracture crack has

passed through a filler particle. Fracture

of petroleum coke particles reveals their

creased lamellar structure, while

fracture of anthracite grains reveals flat

surfaces bearing river patterns.

Observed where the fracture crack has

passed through the binder phase.

Formed where the fracture crack has

passed along the interface between

binder and filler. The filler side is

characterised by filler particle surfaces

with adhering pieces of binder.

The binder side of an interfacial failure

consists of flattened areas of binder

which may bear imprints of lamellae.

Observed where the fracture crack has

traversed a highly porous, thin-cell

walled filler particle. Pores at periphery

are usually filled with pitch coke.

Observed where a small piece of pitch

has become highly porous during

Pore surface

Interparticular fissure

Devolatilisation pore

Interlamellar fissure

Gross fissure

PS

IPF

DP

ILF

GF

binder carbonisation. Similar in

appearance to HPF, but no pores are

filled with coke.

Observed where the fracture crack has

passed through a pore. The pore

surface is covered by a thin film of

pitch coke in which primary QI

particles are evident.

A narrow void between two filler

particles.

A small, spherical pore in a filler

panicle, formed during original

carbonisation or calcination.

A fissure between lamellae of a

petroleum coke filler particle.

A large fissure traversing areas of filler

and binder.

Figure 3.1 illustrates the progression of a fracture crack and the origin of the observed

features. Some of the features arise from the passage of the fracture crack through

different components of the solid carbon matrix, such as fIller particles, while others

arise from the passage of the fracture crack through various types of voids in the

carbon such as pores.

It was decided to use five specimens for each carbon, to enable a reliable determination

of the characteristics of the particular material to be made. The five specimens were

chosen to be as close as possible to the mean tensile strength of the material, and to

±O.5 and ±1 standard deviations from the mean. The rationale behind this was to select

specimens which were representative of the material, rather than looking at them as

individual samples (as was the case with the image analysis work).!t was calculated that

a sampling target of 500 points would be adequate to give reproducibility at the 95%

probability level.

55

Having decided to use five samples and 500 sampling points per carbon, it was clearly

sensible to use lOO points per specimen and 50 points per half specimen. The points

were separated by 0.5mm in the X and Y directions, this being adequate to give

coverage of most of the fracture surface available, and also being easily accomplished

by turning the X and Y movement controls on the SEM through half a turn.

Prior to examination in the SEM, the specimens were gold-coated in a sputter-coater, to

prevent charge building up on their surfaces by the action of the electron beam. The

specimens were then mounted on stubs, and silver paint was applied to electrically

ground them. They were then examined in a Cambridge Instruments S604 scanning

electron microscope.

3.5 Determination of Electrical Resistivity

Five cores, measuring approximately 7cm in length by 1.5cm in diameter, were drilled

from a block of the carbon to be tested, then cleaned ultrasonically and dried. The

apparatus used to measure the resistance of a core consisted of a wooden base with two

plastic clips to support the core. A screw clip was slid over each end of the core, and

tightened to a constant torque. The distance between the inner edges of the clips was

measured using a pair of vernier calipers. For each core, at least eight measurements of

resistance were taken, for a range of lengths between the clips. The measurements were

used to plot graphs of length against resistance. The resistivity of the material was then

calculated.

3.6 Determination of Real Density

Principle

The method used to determine the real density of the carbons depended on comparing

the weight of the carbon and the weight of the same volume of water. Knowing the

density of water (1000 kgm-3), the density of the carbon could then be calculated using

equation 3.6.1.

Method

A density bottle was dried and weighed (weight = P), filled with distilled water and

reweighed (weight = W I), then dried again. A sample of the carbon to be examined

was ground to pass through a 150jlffi sieve. Approximately 1.5cm3 was transferred to

the density bottle, and the weight of the bottle plus carbon was measured (weight = W).

56

The density bottle was filled to between 1/4-1/2 of its capacity with distilled water,

which was boiled for ten minutes under reduced pressure. It was allowed to cool for

five minutes, then a drop of dispersive agent was added to reduce the surface tension

and allow all the fme carbon particles to sink. It was considered that the different

density of the dispersive agent would be negligible for the purposes of this test. The

density bottle was then filled with distilled water, and placed in a beaker of water at

room temperature for at least one hour, to allow the temperature of the contents to reach

equilibrium with the extemal temperature. More distilled water was added as necessary,

then the stopper inserted and the outside of the bottle was dried. The bottle was then

weighed again (weight = W2).

Calculation

Real density was calculated as follows.

Real density = (W P) (Equation 3.6.1)

(W1 - P) - (W2 - W)

Tests were carried out in duplicate, and the mean result taken.

3.7 Wettability

Introduction

One of the factors influencing the quality of the interface between the binder and filler in

a manufactured carbon artefact is the readiness with which the binder wets the filler

during the initial mixing stage. The flow of a liquid into a porous medium depends

mainly on two physicaVchemical properties: the viscosity of the binder (a rheological

property) and the contact angle the binder makes with the filler (a thermodynamic

propeny). Both of these properties are important: a pitch may show good wetting

properties but be so viscous that its use is impractical, whilst if the pitch does not wet

the coke adequately, it will be unable to penetrate it, however low the pitch viscosity.

The type of study described here is therefore important in providing further data to

characterise a pitch, in addition to the usual properties such as density, softening point,

QI content and TI content.

57

Non-wetting Wetting Spreading

Apparatus

The apparatus used consists of the following components:

A Variac power controller, connected to a hot-plate. on which was placed a

specially-designed circular steel block with four cylindrical cavities in the top, each

approximately 3cm deep by 1.5cm wide. A glass cylinder was placed over the top of

the central part of the steel block, and a glass connector over the top of the cylinder. A

rubber bung was placed in the neck of the connector, the bung had a 2mm diameter hole

drilled through its centre, and a longitudinal incision along a radius. The wire of a

hypodermic thermocouple probe was fed through the hole in the centre of the bung.

(See Figure 3.8).

The method is based upon that described by Heintz [36] but includes several

improvements: using a thermocouple to measure the temperature of the coke is both

more sensitive and more accurate than a thermometer, and the use of a steel block

assists in smoothing out fluctuations in the rate of temperatUre rise.

The experimental materials used were Alcan petroleum coke and coal-tar pitch. Coke

particles O.18-0.425mm in size were used; this size range was chosen because too large

a particle size would allow the binder particles to drop into inter-particular gaps, whilst

too small a particle size would bear little relation to the coarse-grained formulations

typically used in electric-furnace carbon electrodes.

Pitch particles 1.4-1. 7mm in size were used. Too small a size of pitch particle would be

difficult to observe, whilst pitch particles that were too large would tend to sag under

their own weight upon softening, making wetting temperature difficult to observe.

Upon softening, particles with sizes in this range formed spherical droplets on the coke

surface (until wetting took place).

The experiment was carried out as follows. The four cavities in the metal block were

filled with coke particles of the size range indicated, and the coke surfaces were

58

levelled. A pitch particle was placed close to the centre of three of the circular coke

surfaces, and the tip of the hypodermic thennocouple probe was insened into the coke

in the fourth cavity.

Preliminary experiments to determine the optimum operating conditions for the Variac

and hot-plate were carried out, to obtain a temperature rise rate of 5°C per minute over

the range BO-1BO°C. It was found that the best way of achieving this was to set the

Variac to 190, heat the block from ambient temperature to approximately BO°C, then let

the temperature stabilise at this value. The Variac and hot-plate were then switched on

again, at the same Variac setting, and the temperature reading noted when the

temperature started to rise. The temperature was checked every minute, to ensure the

rise rate was acceptable. The mean rise rate at the end of the experiment was was also

calculated.

The pitch particles were observed at frequent intervals: the temperature at which the

pitch fonned spheres was noted, and then the temperature at which wetting took place

was recorded. The wetting temperature was taken as the temperature at which there was

a clear flattening of the pitch spheres, as they started to wet the coke surface.

The experiment was repeated, a total of six specimens being used. This was to give

consistency both within a run and between runs.

The chief advantage of this apparatus is its simplicity; its chief disadvantage is the

subjective nature of the measurements. This could be overcome by using an optical

device to observe the moment at which wetting took place. The apparatus could be

used with solid pieces of coke, providing they were small enough to fit into the cavities

in the metal block; the hypodermic probe would be pressed against, or into, the solid

surface of the coke. The pieces of coke would have to fit into the cavities closely, to

ensure even heating took place.

59

FRACTURE CRACK THROUGH

Filler Particle

Binder

Interface Between Binder and Filler

Highly Porous Filler

Highly Porous Binder

Pore

Interparticular Fissure

Devolatilisation Pore

Interlamellar Fissure

FEATURE IDENTIFIER

FF

SF

IF .. 18 .. HPF

HPB

PS

IPF

DPF

ILF

Figure 3.1 - Diagrammatic Representation of Path of Fracture Crack Throug h Electrode

60

Figure 3.2 - ECA Particle Showing Conchoidal Fracture

Surface and River Patterns

61

Figure 3.3 - Petroleum Coke Particle Showing Lamellae

62

· lOlL

Figure 3.4 - Petroleum Coke Particle Showing

Pore Surface

63

Figure 3.5 - Filler Fracture in Petroleum Coke

64

Figure 3.6 - Binder Fracture in Petroleum Coke

65

Figure 3.7 - Filler Side of Interfacial Failure in Anode

Carbon made with Petroleum Coke

66

Thermocouple

Hypodermic Probe

Steel Block

Hot -.. Plate

Thermocouple Wire

Bung

• Glass Cylinder

Pitch Sphere (not to scale)

~---Coke

Sample

Figure 3.8 - Wettability Apparatus

67

4 Results

Results were obtained for seven carbon electrode materials. The results obtained using

each experimental technique will be described under the appropriate headings, and the

findings discussed in the next section.

4.1 Tensile Strength Measurements

Calcined-Anthracite Based Carbons

Summ;uy of Results for Carbon A4

11ean tensile strength = 4.89 11Pa

Standard deviation = 0.74

Standard error = 0.1

11ean apparent density = 1470 kgm·3

Standard deviation = 13

Standard error = 1

Summ;uy of Results for Carbon C

11ean tensile strength = 3.3811Pa

Standard deviation = 0.52

Standard error = 0.07

11ean apparent density = 1505 kgm-3

Standard deviation = 20

Standard error = 2

68

Petroleum-Coke Based Carbons

Summary of Results for Carbon P

Mean tensile strength = 6.32 MPa

Standard deviation = 0.61

Standard error = 0.09

Mean apparent density -. 1596 kgm·3

Standard deviation = 19

Standard error = 2

Summary of Results for Carbon AL6

Mean tensile strength = 6.36 MPa

Standard deviation = 0.65

Standard error = 0.14

Mean apparent density = 1564 kgm·3

Standard deviation = 17

Standard error = 3

Summary of Results for Carbon AL8

Mean tensile strength = 5.71MPa

Standard deviation = 0.75

Standard error = 0.16

Mean apparent density = 1570 kgm·3

Standard deviation = 27

Standard error = 5

69

Summary of Results for Carbon Y

Mean tensile strength = 6.62 MPa

Standard deviation = 1.01

Standard error = 0.14

Mean apparent density = 1550 kgm·3

Standard deviation = 13

Standard error = 1

Summary of Results for Carbon Z

Mean tensile strength = 5.82 MPa

Standard deviation = 0.88

Standard error = 0.14

Mean apparent density = 1537 kgm-3

Standard deviation = 17

Standard error = 2

Results for the tensile strengths of all the samples for a particular carbon showed an

approximatelY normal distribution, within the limits of the sample size. The two

calcined-anthracite based carbons had mean tensile strengths which were significantly

lower than those of the petroleum-coke based carbons. The latter group showed a wide

range of mean tensile strengths, although it is possible that this would also have been

observed if a larger number of calcined-anthracite carbons had been studied.

The real densities of the two calcined-anthracite carbons were similar. The real densities

of the petroleum-coke based carbons showed a much smaller range than the range of

their mean tensile strengths, and were slightly higher than those for the

calcined-anthracite group.

70

4.2 Pore Structural Analysis

Results from computerised image analysis are shown below; the data are the

cumulative data for all the specimens for a particular carbon. More detailed results are

shown in the tables at the end of this chapter.

Carbons are shown in increasing order of mean tensile strength.

Carbon Mean Tensile Porosity No. Pores Pore size Wall size

Strength (MPa) (%) per mm2 (Ilm) (Ilm)

C 3.38 26 2455 25.18 77.26

A4 4.89 32 1681 27.08 58.10

AL8 5.71 24 6344 14.99 49.52

Z 5.82 23 5247 20.80 68.40

P 6.32 23 4972 16.31 55.40

AL6 6.36 23 6461 15.70 54.16

Y 6.62 22 5716 19.45 70.77

These results indicate a clear difference between the two groups of carbons. The

porosity in both calcined anthracite-based carbons is distributed as a relatively small

number of large pores. whereas the petroleum coke-based carbons have a much larger

number of smaller pores. The calcined-anthracite carbons are more porous and have

lower tensile strengths than the petroleum coke based carbons.

71

4.3 SEM Fractography

Summary of Fractographic Data (Number of Observations per

Category)

Carbons are shown in increasing order of mean tensile strength.

Carbon Tensile Number of Observations per Category

Strength

FF BF IFF IFB HPF HPB PS IPF GF

C 3.38 83 157 31 25 0 1 200 3 0

A4 4.89 161 154 29 23 0 0 124 9 0

AL8 5.71 93 112 78 47 3 1 150 15 1

Z 5.82 171 110 120 20 0 0 59 18 0

P 6.32 149 187 3 21 3 3 79 25 0

AL6 6.36 178 142 18 13 1 5 123 11 5

Y 6.62 165 107 68 48 1 2 99 10 0

72

Summary of Fractographlc Data (Percentage of Observations per

Category)

Carbons are shown in increasing order of mean tensile strength.

Carbon Tensile Percentage of Observations per Category

Strength

FF BF IFF IFB HPF HPH PS IPF GF

C 3.38 16.6 31.4 6.2 5 o 0.2 40 0.6 0

A4 4.89 32.2 30.8 5.8 4.6 0 o 24.8 1.8 0

AL8 5.71 18.6 22.4 15.6 9.4 0.6 0.2 30 3 0.2

z 5.82 34.2 22 24 4 0 o 11.8 3.6 0

P 6.32 29.8 37.4 6.4 4.2 0.6 0.6 15.8 5 0

AL6 6.36 35.6 28.4 3.6 2.6 0.2 I 24.6 2.2 1

y 6.62 33 21.4 13.6 9.6 0.2 0.4 19.8 2 o

These results indicate some differences both between the two groups of carbons (calcined

anthracite-based and petroleum coke-based) and within the groups. The weaker of the two

calcined anthracite-based carbons (C) shows a much smaller percentage of filler fracture

than the stronger of the two (A4). The three strongest petroleum coke-based carbons, P,

AL6 and Y, all show comparable percentages of filler fracture. Although Z is only slightly

weaker than these, it too shows a high percentage of filler fracture, but also shows a very

high percentage of interfacial failure (filler side). AL8, the weakest of this group, shows a

much smaller percentage of filler filler fracture.

73

4.4 Electrical Resistivity

Results for the different cores for each carbon are shown in Figures 4.1-4.5.

Insufficient raw material was available to enable resistivity tests to be carried out on

carbons AL6 and AL8.

The calcined-anthracite based carbons showed lower electrical resisitivity than the

petroleum-coke based carbons; the range of values for the first group was 4.15 to 6.78

mQ cm-I, while for the second group it was 5.63 to 6.78 mf.l cm-I.

4.5 Real Density

The results show good agreement between two real density runs for each carbon, and

between the real density method of measuring porosity and the image analysis method,

except in the case of carbon A4. Because of the discrepancy, two funher real density

runs were carried out for this carbon, which confirmed the results for the first two

runs.

4.6 Wettability

Results were obtained for two successive runs; the average temperature rise rate was

4.7°C per minute. It was found that the size of pitch particles used was adequate to

allow sphere formation, wetting' and spreading to be observed. The temperatures at

which all these events took place were found to be consistent between the two runs.

The mean wetting temperature was found to be 157.5°C.

74

Table 4.1 Tensile Strength Measurements for Carbon A4

Specimen Weight Diameter Length Load Tensile Apparent No. (g) (mm) (mm) (divs.) Strength Density

(MPa) (kgm·3)

1 2.7556 14.89 10.60 78.7 6.22 1492 2 2.6030 14.90 10.42 66.9 5.37 1432 3 2.8640 14.89 11.06 61.9 4.69 1486 4 2.6422 14.88 10.35 38.8 3.14 1467 5 2.6596 14.88 10.50 49.5 3.95 1456 6 2.6792 14.89 10.40 55.5 4.47 1479 7 2.7600 14.89 10.75 25.8 2.01 1474

8 2.7036 14.90 10.75 67.2 5.23 1444 9 2.6878 14.87 10.57 58.5 4.64 1463 10 2.7636 14.90 10.76 63.3 4.92 1472 11 2.7733 14.89 . 11.00 67.5 5.14 1447 12 2.7643 14.90 10.70 55.2 4.32 1481 13 2.7798 14.86 10.95 60.5 4.64 1463 14 2.7092 14.90 10.52 49.0 3.90 1476 15 2.6718 14.88 10.45 64.0 5.13 1469 16 2.7700 14.89 10.68 56.3 4.42 1489 17 2.6390 14.88 10.50 66.2 5.28 1445 18 2.6447 1489 10.26 62.4 5.09 1480 19 2.7532 14.89 10.65 66.2 5.21 1484 20 2.6381 14.91 10.35 60.4 4.88 1459 21 2.6966 14.88 10.60 67.9 5.37 1462 22 2.7206 14.88 10.70 61.0 4.78 1461 23 2.7105 14.91 10.60 64.8 5.11 1464 24 2.6852 14.88 10.65 62.1 4.89 1449 25 2.7425 14.91 10.76 66.6 5.18 1459 26 2.6819 14.90 10.62 67.7 5.34 1448

27 2.7013 14.90 10.60 67.4 5~32 1461 28 2.8272 14.88 10.95 60.8 4.65 1484 29 2.7413 14.90 10.60 53.9 4.26 1482

30 2.7442 14.89 10.70 70.4 5.51 1472 31 2.6767 14.88 10.55 67.1 5.33 1458 32 2.7410 14.91 10.65 65.8 5.17 1473

33 2.7868 14.90 10.92 61.9 4.74 1463

34 2.8327 i4.88 10.95 76.9 5.89 1487

75

[Table 4.1 continued]

35 2.7154 14.87 10.68 70.2 5.51 1463 36 2.6650 14.89 10.37 53.1 4.29 1475 37 2.7033 17.89 10.57 66.1 5.24 1468 38 2.7824 14.89 10.81 69.6 5.39 1477 39 2.7690 14.88 10.72 51.6 4.03 1485 40 2.8332 14.90 11.10 53.5 4.03 1463 41 2.6709 14.89 10.41 74.0 5.96 1473 42 2.6726 14.89 10.41 54.2 4.36 1474 43 2.7654 14.91 10.71 60.8 4.75 1478 44 2.7683 14.89 10.72 68.8 5.38 1482 45 2.8099 14.90 10.95 64.1 4.90 1471 46 2.7122 14.92 10.60 73.3 5.78 1463 47 2.8132 14.87 10.91 67.9 5.22 1484 48 2.7581 14.89 10.63 78.2 6.16 1489 49 2.7755 14.92 10.78 57.8 4.48 1472 50 2.7727 14.92 10.70 59.9 4.68 1481

76

Table 4.2 Tensile Strength Measurements for Carbon C

Specimen

No.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

Weight

(g)

2.7702

2.8522

2.7316

2.8858

2.8353

2.8186

2.8502

2.8582

2.7904

2.8308

2.8503

2.8730

2.8892

2.7724

2.8476

2.7960

2.8867

2.8844

2.7702

2.8632

2.8437

2.8457

2.8887

2.9353

2.7746

2.8534

2.9191

2.8811

2.8101

2.7738

Diameter

(mm)

14.87

14.91

14.90

14.92

14.92

14.90

14.92

14.91

14.88

14.92

14.94

14.92

14.93

14.93

14.87

14.93

14.93

14.92

14.92

14.91

14.92

14.89

14.90

14.91

14.85

14.93

14.92

14.92·

14.92

14.90

Length (mm)

10.62

10.70

10.58

10.82

10.74

10.75

10.75

10.91

10.56

10.96

10.74

10.82

10.91

10.61

10.75

10.68

10.94

10.95

10.63

10.87

10.72

10.80

10.90

11.17

10.60

10.88

10.86

10.84

10.67

10.64

77

Load

(divs.)

42.3

43.3

24.5

39.0

48.1

46.7

48.1

49.5

36.3

42.8

48.2

38.5

42.4

44.6

37.7

42.3

49.7

40.6

34.0

58.8

48.6

40.6

43.8

42.0

44.0

52.0

53.0

40.0

51.5

46.1

Tensile

Strength

(MPa)

3.34

3.38

1.94

3.01

3.74

3.64

3.74

3.79

2.88

3.26

3.75

2.97

3.24

3.51

2.94

3.31

3.79

3.10

2.67

4.52

3.79

3.15

3.36

3.14

3.48

3.99

4.08

3.08

4.03

3.63

Apparent

Density (kgm-3)

1501

1526

1480

1525

1509

1503

1516

1500

1519

1477

1513

1518

1512

1492

1525

1495

1506

1506

1490

1508

1516

1512

1519

1504

1511

1497

1537

1519

1506

1494

[Table 4.2 continued)

31 2.8580 14.90 10.82 53.3 4.12 1514

32 2.7847 14.90 10.70 36.5 2.85 1492

33 2.8586 14.921 10.86 48.3 3.72 1505

34 2.6488 14.91 10.83 29.7 2.29 1400

35 2.8950 14.92 10.97 34.3 2.61 1509

36 2.8238 14.90 10.76 36.4 2.83 1504

37 2.8742 14.85 11.31 37.4 2.78 1466

38 2.9648 14.90 11.28 41.3 3.06 1507

39 2.8736 14.92 10.94 50.3 3.84 1502

40 2.8584 14.92 10.80 44.0 3.40 1513 41 2.8298 14.92 10.85 33.8 2.60 1491

42 2.7658 14.80 10.63 41.6 3.30 1512

43 2.7682 14.90 10.48 46.4 3.70 1524

44 2.8488 14.85 10.85 46.5 3.60 1515

45 2.9112 14.92 10.95 56.5 4.31 1520

46 2.7774 14.88 10.70 40.4 3.16 1492

47 2.8488 14.92 10.84 42.0 3.24 1502

48 2.8558 14.92 10.83 50.8 3.92 1507

49 2.8057 14.88 10.67 48.9 3.84 1511

50 2.8133 14.91 10.68 43.8 3.43 1508

78

Table 4.3 Tensile Strength Measurements for Carbon P

Specimen Weight Diameter Length Load Tensile Apparent

No. (g) (mm) (mm) (divs.) Strength Density

(MPa) (kgm·3)

1 3.0176 14.88 10.88 82.8 6.38 1594

2 3.0546 14.89 10.92 93.0 7.14 1606

3 3.0752 l4.90 11.02 81.0 6.15 1600

4 3.1049 14.90 11.14 72.9 5.48 1598

5 3.1260 14.89 11.17 93.0 6.98 1606

6 3.0389 14.90 10.95 83.7 6.40 1591

7 3.1008 14.89 11.28 95.8 7.12 1578

8 3.0465 14.89 10.92 88.2 6.77 1601

9 3.0761 14.90 11.05 83.0 6.29 1596

10 3.0402 14.88 11.00 83.3 6.35 1589

11 3.0096 14.89 10.83 84.0 6.50 1595

12 3.0767 14.90 11.07 92.8 7.02 1593

13 3.0207 14.90 11.02 81.2 6.17 1571

14 3.0157 14.89 10.82 87.5 6.78 1600

15 3.0278 14.88 10.90 79.8 6.14 1597

16 3.0085 14.90 10.80 85.8 6.65 1594 17 3.0577 14.90 10.98 73.0 5.57 1596

18 3.0232 14.89 11.02 80.0 6.08 1575

19 3.0726 14.89 11.04 94.5 7.17 1597

20 3.0365 14.90 10.85 71.8 5.54 1604

21 2.9958 14.89 10.74 83.8 6.54 1601

22 3.1071 14.89 11.21 78.8 5.89 1591

23 3.0090 14.90 10.88 76.0 5.85 1585

24 3.0476 14.90 10.92 91.8 7.04 1600

25 3.0911 14.88 11.16 83.8 6.30 1592

26 3.0300 14.89 10.90 83.8 6.44 1596

27 3.0682 14.89 10.95 92.0 7.04 1608

28 3.0305 14.89 11.01 81.3 6.19 1580

29 3.1068 14.90 10.86 88.0 6.78 1640

30 2.9850 14.90 11.14 72.5 5.45 1536

31 3.0040 14.90 10.85 84.5 6.52 1587

32 3.0081 14.91 10.80 59.0 4.57 1594

33 3.1288 14.89 10.80 82.0 6.36 1663

34 3.0460 14.89 11.28 86.3 6.41 1550

79

[Table 4.3 continued]

35 3.0701 14.91 11.00 75.5 5.74 1598

36 3.1516 14.89 11.05 91.3 6.92 1637

37 3.0950 14.89 11.32 96.3 7.13 1569

38 3.0842 14.88 11.04 58.0 4.40 1606

39 3.0493 14.90 11.10 76.3 5.75 1575

40 3.0928 14.90 11.04 87.8 6.66 1606

41 3.0742 14.89 11.08 82.5 6.24 1593

42 2.9611 14.87 10.70 75.0 5.88 1593

43 2.9357 14.89 10.61 89.2 7.05 1588

44 3.0308 14.89 10.84 78.0 6.03 1605

45 2.9290 14.87 10.49 79.5 6.36 1607

46 2.9846 14.89 10.70 77.0 6.03 i601

47 3.0380 14.90 10.84 83.8 6.47 1606

48 3.0200 14.90 10.80 78.3 6.07 1603

49 3.0580 14.89 . 10.97 87.5 6.68 1600

80

Table 4.4 Tensile Strength Measurements for Carbon AL6

Specimen Weight Diameter Length Load Tensile Apparent No. (g) (mm). (mm) (divs.) Strength Density

(MPa) (kgm·3)

1 2.5376 14.85 9.36 69.0 6.19 1565

2 2.6003 14.88 9.72 77.9 6.72 1538

3 2.4992 14.87 9.19 65.8 6.01 1565

4 2.5798 14.90 9.45 74.4 6.59 1565

5 2.5043 14.87 9.35 71.0 6.37 1541

6 2.5527 14.87 9.28 77.5 7.01 1583

7 2.5499 14.87 9.28 70.4 6.36 1581

8 2.5496 14.89 9.28 85.4 7.71 1577

9 2.5558 14.88 9.42 83.7 7.45 1559

10 2.5669 14.86 9.42 63.8 5.69 1570

11 3.6438 14.90 9.95 59.9 5.04 1523

12 2.4884 14.85 9.18 68.5 6.27 1564

13 2.6103 14.88 9.52 75.5 6.65 1576 14 2.6103 14.98 9.69 65.1 5.59 1528 15 2.5422 14.88 9.31 66.5 5.99 1569

16 2.5678 14.86 9.42 73.5 6.55 1571

17 2.5675 14.85 9.33 80.0 7.20 1588 18 2.5892 14.85 9.52 61.0 5.38 1570

19 2.5171 14.83 9.32 68.7 6.20 1563

20 2.6405 14.86 9.63 77.8 6.78 1580

21 2.5007 14.87 9.32 62.5 5.63 1544

22 2.5569 14.87 9.29 73.2 6.61 1584

23 2.5956 14.86 9.54 70.7 6.22 1568

81

Table 4.5 Tensile Strength Measurements for Carbon AL8

Specimen Weight Diameter Length Load' Tensile Apparent No. (g) (mm) (mm) (divs.) Strength Density

(MPa) (kgm·3)

1 2.4819 14.81 9,36 65.6 5,90 1538 2 2,6119 14.91 9,51 76.7 6.75 1572 3 2.5742 14.76 9,59 65.1 5.74 1568 4 2.5520 14.88 9,33 70.6 6.34 1572 5 2.4998 14.74 9,51 56.3 5.01 1540

6 2.5296 14,87 9.45 63.8 5.66 1541

7 2.5589 14,85 9.45 47.5 4.22 1563

8 2.5640 14,80 9,51 56.2 4.98 1566

9 2.5923 14,86 9,52 79.9 7.05 1569

10 2.5243 14,89 9.37 53.2 4.76 1546 11 2.7426 14.87 9.52 50.7 4.47 1658 12 2.5809 14.33 9.32 64.2 5.79 1602

13 2.6260 14.90 9.51 64.9 5.71 1583 14 2.6670 14.89 9.58 63.0 5.51 1598 15 2.6078 14.88 9.52 73.3 6.46 1574 16 2.6610 14.88 10.04 73.1 6.10 1523 17 2.4882 14.65 9.49 56.7 5.09 1555 18 2.6006 14.88 9.48 62.0 5.48 1577 19 2.5843 14.85 9.58 70.8 6.21 1557 20 2.6135 14.89 9.46 75.8 6.71 1586 21 2.6312 14.88 9.59 71.0 6.21 1577 22 2.5973 14.86 9.52. 61.4 5.41 1572

82

Table 4.6 Tensile Strength Measurements for Carbon Y

Specimen Weight Diameter Length Load Tensile Apparent

No. (g) (mm) (mm) (divs.) Strength Density (MPa) (kgm·3)

1 2.7264 14.85 10.19 95.6 7.88 1544

2 2.7825 14.90 10.33 91.9 7.45 1544

3 2.7934 14.89 . 10.29 81.3 6.62 1558

4 2.7727 14.90 10.23 99.8 8.17 1554

5 2.8058 14.90 10.24 92.7 7.58 1571

6 2.7833 14.93 10.19 83.7 6.86 1559

7 2.7635 14.92 10.20 80.8 6.62 1549

8 2.7729 14.92 10.30 62.3 5.06 1539

9 2.7750 14.88 10.20 74.7 6.14 1564

10 2.7758 14.90 10.25 78.9 6.44 1552

11 2.7651 14.90 10.25 99.0 8.09 1546

12 2.7781 14.90 10.26 82.7 6.75 1552

13 2.8142 14.93 10.31 82.1 6.65 1558

14 2.7657 14.88 10.24 81.2 6.65 1552

15 2.7532 14.89 10.22 78.6 6.44 1546

16 2.7805 14.90 10.25 80.7 6.59 1555

17 2.7652 14.93 10.19 100.3 8.23 1549

18 2.7878 14.90 10.22 68.7 5.63 1564

19 2.7337 14.91 10.18 77.7 6.39 1537

20 2.8981 14.87 10.80 86.3 6.70 1544

21 2.7289 14.89 10.09 75.0 6.23 1552

22 2.7488 14.89 10.22 81.3 6.67 1544

23 2.8078 14.86 10.36 64.9 5.26 1562

24 2.7824 14.92 10.24 89.0 7.27 1553

25 2.8231 14.90 10.30 84.5 6.87 1571

26 2.7773 14.91 10.25 62.6 5.11 1551

27 2.7052 14.90 10.28 67.0 5.46 1508

28 2.7812 14.39 10.27 84.5 6.89 1554

29 2.8157 14.89 10.25 89.3 7.30 1577

30 2.7752 14.91 10.24 98.7 8.07 1551

31 2.7736 14.88 10.17 95.6 7.88 1567

32 2.7575 14.91 10.27 47.7 3.88 1537

33 2.7556 14.86 . 10.60 81.5 6.45 1498

83

[Table 4.6 continued]

34 2.7670 14.89 10.18 87.4 7.19 1560

35 2.7819 14.91 10.25 91.3 7.45 1554

36 2.7619 14.90· 10.19 49.0 4.02 1554

37 2.7858 14.92 10.20 89.0 7.30 1561 38 2.7845 14.89 10.25 91.0 7.44 1559 39 2.8268 14.90 10.35 74.9 6.06 1566 40 2.7464 14.89 10.22 83.0 6.80 1542 41 2.7552 14.90 10.21 91.7 7.52 1547 42 2.7986 14.92 10.23 91.1 7.45 1564 43 2.7333 14.90 10.22 86.9 7.12 1533 44 2.7782 14.90 10.26 62.3 5.08 1552 45 2.8200 14.88 10.47 86.2 6.90 1548 46 2.7843 14.90 10.30 85.0 6.91 1550 47 2.7735 14.90 10.28 77.4 6.30 1547 48 2.7256 14.86 10.24 80.3 6.58 1534 49 2.7624 14.90 10.28 60.7 4.94 1540

50 2.7708 14.92 10.25 63.9 5.21 1545

84

Table 4.7 Tensile Strength Measurements for Carbon Z

Specimen

No.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

Weight

(g)

2.7753

2.6889

2.700

2.6886

2.7465

2.7336

2.7342

2.6993

2.7655

2.7706

2.7273

2.7281

2.8056

2.7716

2.7704

2.7723

2.7272

2.7691

2.7474

2.7899

2.7336

2.7661

2.7342

2.7984

2.7741

2.7346

2.6941

2.7775

2.7392

2.7115

2.7199

2.7412

Diameter

(mm)

14.90

14.92

14.91

14.91

14.90

14.90

14.90

14.90

14.91

14.90

14.90

14.92

14.91

14.90

14.91

14.90

14.90

14.89

14.90

14.91

14.99

14.90.

14.92

14.90

14.91

14.92

14.90

14.92

14.91

14.90

14.89

14.91

Length

(mm)

10.26

10.23

10.22

10.26

10.22

10.24

10.27

10.23

10.20

10.32

10.20

10.21

10.23

10.27

10.32

10.22

10.24

10.24

10.25

10.25

10.22

10.17

10.27

10.26

10.27

10.18

10.30

10.25

10.22

10.22

10.18

10.23

85

Load

(divs.)

71.8

60.0

61.5

44.3

74.2

65.4

53.0

61.4

69.0

76.9

55.6

66.6

77.2

78.1

82.0

79.5

85.7

76.5

85.0

95.1

62.3

87.1

72.9

74.7

71.0

75.0

66.8

76.0

78.2

69.5

62.9

51.8

Tensile

Strength

(MPa)

5.86

4.90

5.03

3.61

6.08

5.35

4.32

5.02

5.66

6.24

4.56

5.45

6.31

6.37

6.65

6.51

7.01

6.26

6.94

7.76

5.07

7.17

5.93

6.10

5.78

6.16

5.43

6.20

6.40

5.69

5.18

4.23

Apparent

Density

tkgm·3)

1551

1503

1512

1500

1540

1530

1526

1512

1552

1539

1533

1528

1570

1547

1537

1555

1527

1552

1536

1558

1515

1559

1522

1563

1546

1536

1499

1549

1534

1521

1534

1534

[Table 4.7 continued)

33 2.7761 14.90 10.24 86.3 7.06 1554 34 2.7471 14.92 10.18 65.2 5.35 1543

35 2.7750 14.89 10.22 76.5 6.27 1559

36 2.7568 14.91 10.20 61.0 5.00 1547

37 2.7157 14.90 10.27 63.7 5.19 1516

38 2.7851 14.91 10.26 79.4 6.48 1554

39 2.7623 14.89 10.29 74.8 6.09 1541

86

Table 4.8 Detailed Pore Structural Results for Carbon A4

Specimen Tensile Detected Count Chord Length Closest

No. Strength Area [No. of [Pore size] Approach

(MPa) [Porosity] pores] (Ilm) [Wall size] (%) (Ilm)

7 2.01 26.9 2932 23.4 63.8

30.5 1911 26.5 60.6

32.5 2772 23.9 50.0

24.3 4362 19.4 60.4

29.8 3233 22.9 54.0

4 3.14 28.7 1861 29.5 73.5

30.5 3570 23.5 53.7

35.2 2096 26.1 48.0

31.5 3225 22.2 48.4

30.9 2392 25.9 58.0

29 4.26 37.5 1924 27.5 45.9

34.9 2777 24.8 46.2

37.8 794 35.0 57.7

28.6 1872 26.1 71.3

36.7 1020 31.4 54.0

21 5.37 43.7 1205 36.5 47.2

33.3 2006 23.9 48.0

34.3 1478 26.8 51.4

34.8 1567 26.0 48.7

37.0 1208 29.8 50.8

1 6.22 33.3 1549 30.7 61.3

23.3 1279 33.3 109.5

28.0 2835 23.1 59.4

26.7 2865 25.9 68.7

34.9 888 33.1 61.9

87

Overall means 32.2 2145 27.1 58.1

Standard deviation 4.7 921 4.3 13.3

Standard error 0.9 184 0.9 2.7

Compared to the other electrode materials examined. carbon A4 shows a very high

porosity. and a fairly small number of pores. The mean wall size is twice the mean pore

size.

88

Table 4.9 Detailed Pore Structural Results for Carbon C

Specimen Tensile Detected Count Chord Length Closest

No. Strength Area [No. of [Pore size] Approach

(MPa) [Porosity] pores] (Il m ) [Wall size]

(%) (Il m )

3 1.94 34.2 1876 35.4 68.1

30.5 1108 34.5 78.6

21.8 3368 22.3 80.1

30.8 3004 25.9 58.3

23.0 3158 22.3 74.7

37 2.78 37.1 2992 28.1 47.7

22.9 3094 23.0 77.4

21.1 3694 21.6 81.0

31.4 2724 25.8 56.3

26.9 3526 23.6 64.1

46 3.16 16.8 3929 18.4 91.3

22.6 5331 18.0 61.5

21.4 3870 20.2 74.3

19.8 3467 22.9 92.6

21.6 2988 25.5 92.6

14 3.51 26.1 2443 35.9 101.6

28.1 2045 42.0 107.6

24.4 3949 22.7 .. 70.1

23.9 3051 26.4 83.8

26.5 3417 24.6 68.2

27 4.08 20.1 1830 25.0 99.3 26.4 2292 25.5 71.1

20.8 2574 26.6 101.6

26.0 4230 21.1 60.3

17.2 2488 25.1 120.6

89

[Table 4.9 continued)

20 4.52 25.0 3705 22.1 66.1

20.2 3567 22.1 87.1

24.2 4660 19.7 61.5

28.9 3821 21.6 53.0

29.7 1795 28.5 67.4

Overall means 26.2 3133 25.2 77.3

Standard deviation 4.8 918 5.5 17.6

Standard error 0.9 168 1.0 3.2

Carbon C shows fairly high porosity, with relatively thick walls (carbon matrix).

90

Table 4.10 Detailed Pore Structural Results for Carbon P

Specimen Tensile N.Area Count Chord Length Closest

No. Strength [Porosity] [No. of [Pore size] Approach

(MPa) (%) pores] (~m) [Wall size]

(~m)

38 4.40 21.0 7361 14.6 55.1

15.4 2540 14.3 78.5

19.9 7238 14.7 58.9

22.6 6398 16.8 57.4

19.9 6481 16.0 64.4

20 5.54 22.8 7897 14.7 50.0 25.0 9595 15.7 47.0

29.2 8219 17.5 42.3

21.5 6951 16.0 58.4

16.8 5282 15.5 76.8

46 6.03 24.5 6081 15.9 48.9

32.8 4572 20.0 40.9

30.5 5868 18.3 41.8

31.8 6330 17.5 37.6 25.7 6790 16.6 47.9

41 6.24 23.4 6023 16.1 52.7

26.7 6851 15.9 43.8

14.6 3484 16.7 97.9 25.0 5356 17.4 52.3

21.4 5416 15.6 57.5

16 6.65 28.2 6892 16.2 41.4

23.6 5820 17.9 57.8

30.2 7618 16.7 38.7 26.0 6497 15.9 43.6 24.0 6053 16.8 53.2

91

[Table 4. IO continued]

37 7.13 19.9 5070 16.6 66.7

18.5 5329 16.4 72.7

22.3 6834 15.8 55.1

18.1 5920 15.3 69.0

23.2 6250 16.2 53.8

Overall means 23.5 6344 16.3 55.4

Standard deviation 4.7 1356 1.2 13.6

Standard error 0.9 248 0.2 2.5

Carbon P is not very porous. The pores are numerous, but relatively small.

92

Table 4.11 Detailed Pore Structural Results for Carbon AL6

Specimen Tensile N.Area Count Chord Length Closest No. Strength [Porosity] [No. of [Pore size] Approach

(MPa) (%) pores] (jJ.m) [Wall size]

(jJ.m)

11 5.04 20.0 7235 16.9 67.4 20.3 7861 15.7 61.7 19.6 7369 14.8 60.7 22.2 7519 16.8 58.7 30.2 8769 16.4 37.8

10 5.69 16.4 7068 14.8 75.6

20.9 8654 15.0 56.7 21.7 8354 16.6 58.5 22.4 9532 15.4 53.2 17.7 6007 17.2 79.8

12 6.27 26.0 9879 15.1 42.9

22.0 9646 13.4 47.6

26.6 9004 14.7 40.6'

22.9 13683 13.4 45.2

30.2 8769 16.4 37.8

20 6.78 20.5 7335 15.6 92.9 25.6 7458 16.1 46.8

23.1 8342 15.1 50.3 28.5 7489 16.4 41.2

20.5 7866 14.9 57.6

17 7.20 23.0 7682 16.2 54.4

22.7 8650 14.1 48.0

16.9 7044 14.0 68.7 23.5 7286 15.5 50.6 22.9 7679 15.6 52.6

93

[Table 4.11 continued]

8 7.71 28.9 7342 16.0 39.3

26.7 8218 14.9 40.7

19.4 6613 23.3 62.7

25.4 8298 14.9 43.8

26.2 6940 16.6 46.7

Overall means 22.9 8244 15.7 54.2

Standard deviation 3.7 1391 1.7 13.2

Standard error 0.7 254 0.3 2.4

Carbon AL6 is not very porous. It has a large number of particularly small pores.

94

Table 4.12 Detailed Pore Structural Results for Carbon AL8

Specimen Tensile N.Area Count Chord Length Closest

No. Strength [Porosity] [No. of [Pore size] Approach

(MPa) (%) pores] (I1m ) [Wall size]

(I1m )

7 4.22 20.0 10093 12.2 48.5

18.7 7493 13.9 60.1

21.4 8656 14.3 52.3

20.5 8228 14.2 55.1

20.4 8014 13.6 53.0

10 4.76 27.3 8658 16.0 46.6

31.2 7712 16.8 37.1

29.9 7484 15.8 37.0

33.4 7299 17.2 34.4

33.5 9662 14.4 28.7

22 5.41 25.3 9903 13.8 40.7

21.2 9108 12.8 47.5

24.3 9036 14.4 44.7

24.8 9874 13.5 41.1

24.5 9535 13.4 41.4

12 5.79 22.8 8138 15.4 52.2

28.8 7306 15.9 39.5

21.0 7786 14.9 55.8

22.4 8638 13.8 47.8

23.8 9466 13.9 44.5

21 6.21 18.7 6992 15.2 65.8

24.2 6760 16.0 50.0

19.3 7740 13.7 57.3

25.3 7736 16.2 47.9

19.8 7077 15.5 62.9

95

[Table 4.12 continued]

2 6.75 22.4 6346 17.7 61.2

20.4 7061 15.5 60.5

19.1 6463 15.6 66.3 24.0 7944 15.8 49.9

23.3 8145 18.5 66.3

Overall means 23.7 8145 15.0 49.9

Standard deviation 4.2 1066 1.5 10.0

Standard error 0.8 195 0.3 1.8

Carbon AL8 is not particularly porous. It has a large number of small pores.

96

Table 4.13 Detailed Pore Structural Results for Carbon Y

Specimen Tensile N.Area Count Chord Length Closest

No. Strength [Porosity] [No. of [Pore size] Approach

(MPa) (%) pores] (Jlm) [Wall size]

(Jlm)

32 3.88 24.3 5679 26.0 80.9

21.0 7300 18.1 68.1

18.4 6131 20.4 90.5

18.4 5118 23.2 99.9

17.5 3572 24.0 112.9

49 4.94 22.2 7016 18.6 65.2

24.2 7257 19.1 59.9

24.0 8643 17.9 56.7

23.1 8434 18.3 61.0

23.7 11889 16.4 52.6

27 5.46 19.3 5956 20.2 84.5

19.9 7022 19.1 76.9

19.8 6311 18.3 74.3

18.8 6226 18.8 81.2

17.7 5763 20.1 95.5

47 6.30 22.4 6513 20.3 70.1

23.0 5889 19.6 65.9

26.5 7464 21.2 58.7

21.5 7281 17.8 65.0

22.1 8279 17.4 61.4

34 7.19 19.7 9507 16.6 67.8 22.1 8830 17.7 62.3

25.0 7663 19.2 57.7

21.1 6766 19.4 72.5

22.4 6612 20.1 69.3

97

[Table 4.13 continued]

1 7.88 26.6 8990 19.8 54.5

25.2 10772 17.8 52.9

19.6 7613 17.4 71.7

26.2 6187 22.7 64.2

20.8 8135 18.1 69.3

Overall means 21.9 7294 19.5 70.8

Standard deviation 2.6 1681 2.2 14.4

Standard error 0.5 307 0.4 2.6

Carbon Y has quite a large number of fairly small pores.

98

Table 4.14 Detailed Pore Structural Results for Carbon Z

Specimen Tensile N.Area Count Chord Length Closest

No. Strength [Porosity] [No. of [Pore size] Approach

(MPa) (%) pores] (Il-m) [Wall size]

(Il-m)

4 3.61 30.3 7902 21.3 49.1

25.0 8775 18.6 55.8 21.5 7614 19.8 72.3 21.0 7884 17.5 66.0 23.8 8055 19.2 61.6

7 4.32 22.5 7889 21.0 72.3

22.6 6913 20.6 70.5 24.9 5873 21.3 64.2 22.1 7088 18.6 65.8 25.2 6967 20.8 62.0

37 5.19 18.7 4229 20.8 90.6 22.5 7406 20.1 69.3

24.6 6513 23.3 71.1 26.2 6821 23.0 64.6

24.2 6680 20.4 63.9

5 6.08 22.6 6657 20.3 69.6 24.0 7028 21.4 67.9 23.1 6809 20.8 69.5

21.3 5772 20.8 77.0 21.7 6028 20.7 74.9

19 6.94 26.0 5903 24.1 69.6 26.1 5393 27.0 76.4

18.6 4443 21.4 94.0

22.4 6522 20.3 70.3

19.2 6009 18.9 79.4

99

[Table 4.14 continued)

20 7.76 21.5 7805 18.4 67.2

23.6 6948 20.2 65.4

20.6 6725 18.1 69.5

24.7 5721 22.7 69.2

26.0 6474 22.3 63.3

Overall means 23.0 6695 20.8 68.4

Standard deviation 2.5 1026 1.9 8.7

Standard error 0.5 187 0.4 1.6

Carbon Z has quite a large number of fairly small pores.

100

Table 4.15 - Detailed Fractographlc Data for Carbon A4

(Number of Observations per Category)

Specimen! No. of Observations in Each Category

half

FF BF IFF IFB HPF HPB PS IPF

8 14 12 0 1 0 0 23 0

22 21 3 0 0 0 3 I

10 15 19 3 1 0 0 11 1

13 15 4 2 0 0 15 I

35 11 16 2 6 0 0 14 1

13 13 4 2 0 0 15 3

16 21 17 3 0 0 0 9 0

15 19 1 2 0 0 13 0

36 22 7 7 6 0 0 6 2

15 15 2 3 0 0 15 0

TOTALS 161 154 29 23 0 0 124 9

GF

0

0

0

0

0

0

0

0

0

0

0

Carbon A4 shows a large number of binder and ftller fractures, and a relatively small

number of interfacial failures, indicating that it is well-bonded.

101

Table 4.16 - Detailed Fractographic Data for Carbon C

(Number of Observations per Category)

Specimenl No. of Observations in Each Category

half

FF BF IFF IFB HPF HPB PS IPF

12 8 4 5 5 0 0 28 0

14 12 7 2 0 0 15 0

18 6 19 4 0 0 0 19 2

9 20 3 3 0 0 15 0

25 6 13 3 1 0 0 27 0

6 17 1 2 0 0 24 0

43 4 20 3 2 0 1 20 0

10 21 3 6 0 0 10 0

48 13 11 2 1 0 0 22 1

7 20 0 3 0 0 20 0

TOTALS 83 157 31 25 0 1 200 3·

GF

0

0

0

0

0

0

0

0

0

0

0

Carbon C shows a large number of pore surfaces. and less filler particle fractures.

102

Table 4.17 • Detailed Fractographlc Data for Carbon P

(Number of Observations per Category)

Specimen! No. of Observations in Each Category

half

FF BF IFF IFB HPF HPB PS IPF

5 13 25 4 0 0 0 5 3 21 21 3 1 0 0 2 2

22 19 15 4 2 0 0 10 0 9 24 5 2 0 0 8 2

25 15 17 2 2 0 2 7 4 20 8 2 1 0 1 16 2

35 12 18 4 4 2 0 7 3 11 22 4 4 0 0 3 6

49 14 22 1 2 0 0 9 2

15 15 3 3 1 0 12 1

TOTALS 149 187 32 21 3 3 79 25

GF

0

0

0 0

0

0

0 0

0 0

0

Carbon P shows a very large number of binder and filler fractures, indicating that it is

very well-bonded and may therefore be expected to be of relatively high tensile

strength.

103

Table 4.18 - Detailed Fractographlc Data for Carbon AL6

(Number of Observations per Category)

Specimenl No. of Observations in Each Category

half

FF BF IFF IFB HPF HPB PS IPF

3 21 16 0 0 0 0 11 2

17 14 1 5 0 I 12 0

6 21 10 3 1 3 0 11 1

20 15 1 0 I 0 18 3

7 19 20 1 0 0 0 IQ 0

13 11 3 2 0 0 19 0

10 21 15 1 1 0 0 9 0

17 14 3 0 0 0 14 0

13 16 14 2 1 1 0 14 2

13 13 3 3 0 0 15 3

TOTALS 178 142 18 13 1 5 123 11

GF

0

0

0

2

0

2

0

1

0

0

5

Carbon AL6 shows a large number of binder and filler fractures. It would therefore be

expected to be of high tensile strength.

104

Table 4.19 - Detailed Fractographlc Data for Carbon AL8

(Number of Observations per Category)

Specimen! No. of Observations in Each Category

half

FF BF IFF IFB UPF UPB PS IPF

3 16 13 1 1 0 1 17 1

10 11 1 0 0 0 27 1

4 9 19 2 1 0 0 18 1

8 14 6 2 2 0 16 1

17 15 6 3 4 0 0 21 1

12 10 5 10 0 0 11 2

18 5 14 18 10 0 0 2 1

6 7 16 5 I 0 13 2

20 7 7 12 6 0 0 14 4

5 11 14 8 0 0 11 1

TOTALS 93 112 78 47 3 1 150 15

GF

0

0

0

1

0

0

0

0

0

0

1

Carbon AL8 shows a smaller number of binder and a much smaller number of filler

fractures than Carbon AL6. It would therefore be expected to be of lower tensile

strength.

105

Table 4.20 - Detailed Fractographlc Data for Carbon V

(Number of Observations per Category)

Specimenl No. of Observations in Each Category

half

FF BF IFF IFB HPF HPB PS IPF

3 20 11 4 0 0 0 13 1 18 12 4 2 0 0 12 0

5 10 15 10 5 0 0 10 0 13 12 8 8 0 0 9 0

9 16 5 10 3 0 0 16 0 16 9 2 17 0 0 5 1

18 20 12 2 4 0 0 11 1 16 13 7 1 0 2 10 1

43 16 10 12 5 0 0 6 2 20 8 7 3 1 0 7 4

TOTALS 165 107 66 48 1 2 99 10

GF

0 0

0

0

0

0

0

0

0

0

0

Carbon Y has a larger number of interface failures than carbons P, AI6 and ALS, and

would therefore be expected to be of lower tensile strength.

106

Table 4.21 • Detailed Fractographlc Data for Carbon Z

(Number of. Observations per Category)

Specimen! No. of Observations in Each Category

half

FF BF IFF IFB HPF HPB PS IPF

1 24 11 9 3 0 0 1 2 •

22 12 11 0 0 0 5 0

2 12 10 15 3 0 0 8 2

18 9 14 2 0 0 4 3

6 19 10 9 2 0 0 8 2

22 14 8 4 0 0 10 2

15 18 9 12 1 0 0 8 2

10 14 16 3 0 0 6 1

18 12 12 15 1 0 0 5 5 24 9 11 1 0 0 4 1

TOTALS 181 110 120 20 0 0 58 20

GF

0

0

0

0

0

0

0

0

0

0

0

Carbon Z shows a very large percentage of interface failures, especially filler side. It

would therefore be expected to be of relatively low tensile strength.

107

Table 4.22 - Results for Electrical Resistivity

Carbon

A4

C

P

AL6

AL8

Y

Z

Mean Resistivity (mn cm· l )

5.08

4.15

6.78

5.63

6.12

108

Table 4.23 - Results for Real Density/Porosity

Carbon

A4

C

P

AL6

AL8

Y

Z

Apparent Real

Density

(kgm·3)

1470

1505

1596

1564

1570

1546

1537

Density

(kgm·3)

1848

1853

1950

1889

1974

2025

1949

2058

2150

2171

2062

2038

2011

2098

Calculated Mean

Porosity

(%)

20

21

23

20

19

21

20

24

27

28

24

25

24

27

Calculated

Porosity (%)

20.5

21.5

20

22

27.5

24.5

25.5

Q800

Porosity

(%)

32

26

23

23

24

22

23

There is good agreement between the values for porosity obtained from the image

analysis data and calculated from the real density results, except in the case of carbon

A4 and possibly carbon C. Since these were the two carbons based on calcined

anthracite, it is possible either that the image analysis results for these materials are less

accurate than those for the petroleum coke· based carbons, or that the method for

measuring real density is unsuitable for use with these materials. It is considered that

the former is the more likely, since the accuracy of the method for real density should

not depend on the type of material used.

109

Table 4.24 - Results for Wettability Tests

Run 1

Time (min.) Coke Temperature (OC) Comments

0 82 1 87 2 92 3 97 4 103 5 109 6 115 7 121 Spheres start to fonn

8 127 9 132 Spheres fully fonned

10 138 11 143 12 148 13 153 14 158 Wetting temperature

15 163 Spreading temperature

16 167 17 171 Pitch almost disappeared

18 175 Penetration complete

Average temperature rise rate = S.2°C per minute.

110

----- --- -----

Run 2

Time (min.) Coke Temperature (OC) Comments

0 85

1 89

2 93

3 97

4 102

5 107

6 112 Pitch softening

7 117 Spheres StaIt to form

8 122

9 127 Spheres fully fonned

10 132

11 137

12 142

13 147

14 152

15 157 Wetting temperature

16 162

17 167

18 171 Spreading temperature

19 174 Penetration complete

Average temperature rise rate = 4.7°C per nlinute.

111

Figure 4.1 - Resistivity Results for Carbon A4

28

0 27- 6

26 0 0

6" • 25

0

24 0

C --S 23 0 <D

~ () s::: 0 0 - 22 (I)

·iii <D ~ 0 ~

21 0

6 .. 20 0

ca Legend

• • Core 1 19 6 0 Core 2

• Core 3 18- 60 0

• o Core 4

17 • 6 Core 5

1 1.5 2 2.5 3 3.5 4 4.5 5 5.5

Length (cm)

112

Figure 4.2 - Resistivity Results for Carbon C

28T-----------------------------------~

26 •

113

Figure 4.3 - Resistivity Results for Carbon P

27 0

26- • ~

25 0

• • 24-

0 .L1 L1 0

C 23-

g 0 ., • 0 22- • c:

00 -E • VI .iij • .,

21- 0 a::

• 20 L1 0

~o Legend 19 •• 0 • Core 1 . ~

0 Core 2 0 18 • Core 3

0 0 Core 4

L1 Core 5 17 , , , , ,

1.5 2 2.5 :3 3.5 4

Length (cm)

114

Figure 4.4 - Resistivity Results for Carbon Y

27

~ 26 D'V

25 +

24

~'V 23 ill

C +iI -.S 22

Q) 0 c 0 ..... 21-VI

"iii Q)

>+ c::: 20 M> Legend

0 • Core 1

19 0 Core 2

• Core 3

18 0 Core 4

.t /:;. Core 5

X Core 6 17 -

'V Core 7 0 + Core 8

16 , , , , 1 1.5 2 2.5 3 3.5 4 4.5 5

Length (cm)

115

Figure 4.5 - Resistivity Results for Carbon Z

26,-----------------------------------~

o 24

116

5 Discussion and Conclusions

The principal objective of this research project was to examine a range of different

electrode materials used in the aluminium industry, investigating the relative

contributions of their pore structural characteristics and binder-ftller bond quality to

their mechanical strength and electrical resistivity. As described in the literature review,

although metallurgical cokes do not strictly conform to classical brittle fracture theory,

the Griffiths theory may be applied successfully; recent work by Patrick, Walker,

S~rlie and others [57, 58] suggests that this may also be the case with baked electrode

carbons. On the basis of the Griffiths theory of fracture, the strength of a structure is

determined by the number and size of flaws (cracks) present; if a crack reaches its

critical length, it will extend rapidly and failure of the structure will ensue. Since the

largest flaws in a coke are associated wi th its porous structure, it is not unreasonable to

suggest that the pores may correspond to the Griffith critical cracks. This hypothesis

has been applied with some success to metallurgical cokes, but does not take into

account the two-phase structure of electrodes used in the aluminium industry. For such

materials it may be necessary to consider the effect of the quality of the interface

between the two phases (binder and filler) of the electrode; for example, a

poorly-bonded electrode may fail at the binder-filler interface before failure due to the

pore structural characteristics can take place. Hence the two main experimental

procedures of pore structural analysis and fractography were used in conjunction with

each other, on the same carbon electrode materials, with the intention of relating the

results from the two procedures. Results could have fallen into one of several

categories:

(1) Finding that there was no relationship between tensile strength and pore structural

factors would indicate that the experimental procedure used was unsuitable, the

wrong parameters were being examined, or the pore structural factors are of little

or no importance in a two-phase material such as an aluminium electrode. On the

basis of previous work, this would have been an unexpected result.

(2) Finding that the observed tensile strengths of the electrode materials could be

satisfactorily accounted for by the pore structural data would indicate that the

two-phase structure of the electrode material was irrelevant to its strength. This

would have meant that electrodes had similar tensile strength properties to

metallurgical cokes.

(3) Finding that the pore structural results were unrelated to the tensile strengths of the

electrodes, but that the fractography results satisfactorily accounted for the

117

observed strengths, would inclicate that the tensile strength of aluminium electrodes

was completely accounted for by the efficiency of the binder-filler interface, the

experimental procedure used was unsuitable, or the wrong parameters were being

examined.

(4) Fincling that the tensile strengths could be accounted for by a combination of the

pore structural results and fractography results would indicate that both the pore

characteristics and the efficiency of the interface between binder and filler

contributed to the mechanical strength of the electrode. If this were found to be the

case, it would remain to be determined what the relative contributions of the two

factors were in electrodes fabricated from different materials.

(5) Finding that the tensile strengths were unrelated either to the pore structural results

or the fractography results would indicate that the procedures used were

insufficiently sensitive to differentiate between the materials or were unsuitable

because they were provicling data about features of the carbons which were not

significant in influencing the properties being investigated here.

Which of these categories is correct will be discussed below. The first three

sub-sections of this cliscussion consider the results for tensile strength, tensile strength

and porosity, and tensile strength and fractography respectively. The next section

attempts to relate the results for tensile strength, porosity and fractography. The

following section considers the results for resistivity, and the final section draws some

overall conclusions.

5.1 Tensile Strength

The tensile strengths of the group of specimens for each carbon were found to follow

an approximately normal distribution. The tensile strengths of the two calcined

anthracite-based carbons were significantly clifferent, at 4.89 and 3.38 MPa, while their

mean apparent densities were similar, at 1470 and 1505 kgm-3 respectively. The mean

tensile strengths of the petroleum coke-based carbons were all higher than those of the

calcined anthracite-based carbons; the values ranged from 5.71 MPa (for carbon AL8)

to 6.62 MPa (carbon Y). This is a smaller range than that for the calcined

anthracite-based carbons, and indicates that less variation is likely to occur in the tensile

strengths of electrodes made- with a petroleum coke filler. Since only two calcined

anthracite-based carbons were available, it is clifficult to draw definite conclusions, but

on the basis of these results, petroleum coke is the preferred material for use as the

filler, where mechanical strength of the electrode is important

118

A considerable amount of work has been done on tensile strengths of metallurgical and

other cokes; for example, Patrick and Stacey [55] examined cokes with strengths from

2-6 MPa. This is smaller than the range for the petroleum coke-based materials

examined in this project. Less has been published on strengths of calcined

anthracite-based electrode materials, but recently some work has been done on such

materials with strengths ranging from 1-5 MPa [57, 58]. This range covers the

strengths of the two calcined anthracite-based carbons examined in this project.

However, the different pieces of work cannot be compared directly because different

specimen sizes were used in tensile strength testing; larger specimens will have lower

tensile strengths than smaller specimens made of the same material, because there is an

increased probability of a critically-sized flaw being present.

5.2 Strength/Porosity Results

The pore structural characteristics indicate significant differences between the calcined

anthracite-based and petroleum coke-based carbons. The latter all have much higher

mean tensile strengths than the former, and also have a much larger number of smaller

pores. The total porosity is also somewhat lower. Given that there is a greater disparity

between the pore sizes and number of pores than between the total porosities, it is a

reasonable assumption that the number of pores and size of the pores are more

significant parameters than the total volume porosity.

It is possible to draw some general conclusions about the relationship between the

strength of the carbons and their pore structural characteristics. For the materials used

in this work, those based on calcined anthracite as the filler resulted in relatively weak

electrodes with a coarse structure, with relatively large filler particles and pores (mean

pore size around 25J.Lm). This coarse structure is easily seen under polarised light or

using the image analysis system. For the materials used in this work, the use of

petroleum coke resulted in stronger electrodes with a much more fine-grained structure,

with relatively small filler particles and pores (mean pore size aroundI5-2DJ.Lm)_ Within

the group of petroleum coke-based electrodes, increasing tensile strength was asociated

with decreasing porosity, and to some extent with decreasing pore size. The latter

supports the view that the pores are the source of Griffith critical cracks; on this basis,

the smaller the pores, the more difficult will be the propagation of a critically-sized

crack, and the stronger will be the carbon.

For metallurgical cokes and some baked carbon electrode materials, the tensile strength,

number of pores, pore size and wall size have been related using an equation of the

form S.N = k + a.W/p2, with the values of the constant k and coefficient a depending

on the raw materials used [56, 57,58]. It was reasonable to try to determine if this

119

relationship was applicable to the materials being examined here.

Regression analysis of the pore sttuctural data yielded the equation:

S.N = -12209.W/p2 + 233630

R-sq 84.2%

S.E.E = 5992

This indicates a strong correlation between the quantities S.N and W/p2, which is

shown in Figure 5.1, using a single point to represent the data for all the specimens of

each carbon. The result shows that the pore sttuctural data obtained for the carbon

materials examined here can be successfully related to the tensile strengths of the

materials, using an equation of the form S.N = k + a.W/p2. Examination of the

distribution of points on the graph confirms that there are significant differences

between the two groups of electrodes, and that the raw materials used have a major

effect on the properties of the finished electrode.

On the basis of previous work, it was reasonable to suggest that plotting the values for

S.N against those for W /p2 for the individual specimens of the different carbons would

produce a series of straight lines, with a different gradient for each carbon. As shown

by Figure 5.2, this did not prove possible. The reason for not being able to obtain the

regression lines may be that a much larger number of specimens would be needed, the

error associated with a single specimen being considerable. An alternative explanation

is that the procedure used is insufficiently sensitive to differentiate between the

materials, possibly because of the difficulty of adequately reducing the level of human

error in the image analysis work.

Overall, it may therefore be concluded that the tensile strengths of the carbons examined

can be accounted for to a large extent by their pore sttuctural characteristics, but that

the strengths cannot be accounted for completely in this way.

120

5.3 Strength and Fractography

Carbon Tensile Percentage of Observations per Category

Strength

(MPa)

FF BF IFF PS

C 3.38 16.6 31.4 6.2 40

A4 4.89 32.2 30.8 5.8 24.8

AL8 5.71 18.6 22.4 15.6 30

Z 5.82 34.2 22 24 11.8

P 6.32 29.8 37.4 6.4 15.8

AL6 6.36 35.6 28.4 3.6 24.6

Y 6.62 33 21.4 13.6 19.8

Of the two calcined-anthracite based carbons, the one based on GCNgraphite (carbon

A4) shows higher tensile strength and a relatively large percentage of filler fracture,

indicating good binder-filler bonding, while the one based on ECA (carbon C) shows

a much lower percentage of filler fracture and a slightly higher percentage of interfacial

failure (fIller side); the relative weakness of this carbon may be partly due to the higher

percentage of interfacial failure and also panly to the higher percentage of pore

surface. It is possible that the shape of the pores is significant here; a material with

very flattened or elongated pores would be likely to show more pore surface on a

fracture than would a material with spherical pores .

. Of the five petroleum coke-based carbons, strength defmitely appears to be related to

the percentage of fIller fracture. The three strongest carbons, P, AL6 and Y, all show a

high percentage of fIller fracture, indicating good binder-filler bonding (since the filler

has failed, and not the binder-filler interface). Carbon Z is weaker than these carbons,

although it shows a comparable percentage of filler fracture; this may be accounted for

by the very large percentage of interfacial failure, so that binder-filler bonding is less

121

effective than in carbons P, AL6 and Y. The weakest petroleum coke-based carbon,

AL8, shows both a relatively low percentage of filler fracture and a high percentage of

interfacial failure. These data indicate that the percentages of filler fracture and

interfacial failure are the most important factors in correlating tensile strength and

fracto graphic features in carbon electrodes, since the other types of feature observed

show no significant correlation with the strengths of the different carbons. This may

be due to the difficulty of categorising some features; for example, it is often difficult

to distinguish between binder fracture and interfacial failure (binder side).

Looking at the results for both sets of carbons, the results for the calcined

anthracite-based carbons are largely explicable in terms of the percentages of filler

fracture and pore surface. This is because the calcined anthracite-based carbons are

relatively well-bonded at the binder-filler interface, so the controlling factors for

fracture are the strength of the filler particles and the amount of porosity. A high value

for the former clearly contributes to a strong electrode, while a high value for the latter

will tend to result in a weaker electrode, since however strong the filler particles are,

the electrode will be weak if much of its volume is made up of pores.

The results for the petroleum coke-based carbons are largely explicable in terms of the

percentages of filler fracture and interfacial failure. This is because the petroleum

coke-based carbons show less effective binder-filler bonding, so the controlling

factors for fracture are the strength of the mler particles (for the same reason as

described above) and the percentage of interfacial failure, which reflects the quality of

the binder-filler bonding. The smaller pore size and lower total porosity of the

petroleum coke-based carbons indicates that these factors have a less dominant effect

in determining the tensile strength.

122

5.4 Strength/Porosity Results and Fractography

Filler fracture was chosen as the single most reliable indicator of quality of

binder-filler interface bonding, a low percentage of filler fracture being observed in

poorly-bonded materials.

A table summarising tensile strength, fractographic and pore structural data is shown

below.

Carbon Mean Tensile % Filler % No. Pores Pore

Strength Fracture Porosity per mm1 Size

(Mh) (~m)

C 3.38 17 26 2455 25.18

A4 4.89 32 32 1681 27.08

AL8 5.71 19 24 6344 14.99

Z 5.82 34 23 5247 20.80

P 6.32 30 23 4972 16.31

AL6 6.36 36 23 6461 15.7

Y 6.62 33 22 5716 19.45

It is difficult to relate the results from strength/porosity measurements and fractography

on a simple basis, because many factors are involved, some of them mutually

dependent. However, some conclusions can be drawn by treating the carbons on the

basis of the type of filler used. The stronger of the calcined anthracite-based carbons

(A4) had a larger percentage of filler fracture and a larger total porosity; the pores were

larger and less numerous. These results indicate opposing effects of the propenies of

the carbon matrix and the pore strucrural characteristics on the strength of the material,

since it would be expected on the basis of previous work, and on other results in this

study, that higher strength would be associated with a larger number of smaller pores.

This may be observed where a larger number of carbons are used; the fact that only two

123

electrode materials made with calcined anthracite as filler were available makes it difficult to draw definite conclusions. For the petroleum coke-based carbons, higher

strengths are associated with a larger amount of filler fracTUre. The former is what

would be expected in a range of materials with progressively increasing tensile

strengths (provided that porosity of all the materials was comparable).

124

5.5 Comments on Resistivity

Of the five carbons tested. the resistivities of the GCA/graphite-based and ECA-based

materials were comparable. and lower than the values obtained for the petroleum

coke-based carbons. This may be because although ECA is associated with poor quality

bonding. this factor may be offset by the larger size of the filler particles. which

reduces the total area of interfaces. and thus enhances conductivity. The presence of the

graphite in carbon A4 may enhance conductivity. because of the good conductivity

along the graphite planes. It is also possible that the fillers in the anthracite-based

carbons may have been heated to higher temperatures. which may have lead to panial

graphitisation and consequently lower electrical resistivity.

The mean electrical resistivity for each carbon was compared with its mean tensile

strength. to see if low resistivity was associated with high strength. on the basis that a

well-bonded material might be expected to exhibit low resistivity. The strongest carbon

(Y) did show the lowest resistivity of the petroleum coke-based materials. but it is very

difficult to draw any definite conclusions because of the small number of different

materials available (two calcined-anthracite based and three petroleum coked-based) for

this test. However. on the basis of the results obtained. calcined anthracite materials

would appear to be the most suitable fIllers.

125

5.6 Conclusions

The results which have been obtained are encouraging in terms of relating the strength

of the electrode materials to aspects of the porous structure and fractographic features.

The results for the image analysis carried out appear to be fairly satisfactory, in giving

results for calculated strengths. Considering all the carbons together, attempts to relate

individual specimens for a particular carbon on the basis of an S.N against Wtp2 type

of equation were not very successful. This may have been due to the procedure used

being insufficiently sensitive to differentiate between the specimens. However, there

are clear differences between the carbons, and using the data for each carbon to produce

a single point supports the previously-reported work which suggests the general

applicability of an equation of the form S.N = k + a. W tp2.

In fractography, the number of sampling points used appears to give an adequate

number of points in the key categories ofmler fracture and interfacial failure. Also, the

number of cores looked at seems to be adequate, with little difference between results

for each core. The main means of producing accurate results using this method is

careful allocation of each sampling point.

The procedures for determination of electrical resistivity, wettability and real density

appear to be sound, and are ready for use on further carbon materials (although the

wettability test can only be used on materials where both the binder and filler are

available).

It has been demonstrated that different raw materials result in electrodes with widely

differing mechanical. pore structural and electrical properties. The differences in

strength between the different electrodes cannot be accounted for solely in terms of

either pore structural characteristics or fractography results. As described in the

literature review, a strong case has been made out for the pores in cokes being the

source of the Griffith critical cracks which lead to mechanical failure of the coke. This

may hold true for electrodes as well, but the present work has indicated that the

two-phase nature of electrode carbons has a significant effect on the mechanical

properties of the electrode. The results from this project suggest that in electrodes

which have a high volume porosity and large mean pore size, the porosity effects are

dominant, with the interfacial characteristics playing a less important role in determining

the electrode's strength. In other words, the electrode fails by pore structural weakness

before any interfacial weakness can have an effect Evidence for this conclusion may be

seen in the fractography results for the two calcined anthracite-based carbons; the

126

percentage of filler fracture in these materials is not greatly different to the values

obtained for the petroleum coke-based materials, but these electrodes are more porous

and considerably weaker than any of the petroleum coke-based electrodes.

Conversely, in electrode materials where the volume porosity is low and the mean pore

size is small, it is suggested by the present work that the interfacial effects play the

dominant role in determining the strength of the electrode; the electrode fails because of

a poor quality binder-filler interface before Griffith critical cracks can develop.

Evidence for this conclusion may be seen in the fractography results for the petroleum

coke-based electrodes, where the weakest showed the lowest percentage of interfacial

failures, while the strongest showed better interfacial bonding.

In practice, in nearly all commercial electrode materials it is difficult to distinguish

clearly between pore s~ctural and binder-filler bond quality effects. However,

experiments can be envisaged to study the two effects separately, as described below.

127

5.7 Recommendations for Further Work

The image analysis procedure should be examined critically, with a view to increasing

the reliability of setting the detection level. Also it may be necessary to use a larger

number of specimens, or further refine the procedure used, to increase the reliability of

the results.

In fractography, the procedure used appears to produce reliable results. More carbons,

especially calcined anthracite-based, need to be looked at

The effects of interfacial bond quality could be examined by fabricating electrodes with

a very low volume porosity and very small pores, to enable differences in bond quality

to be studied without having to take into account porosity effects. Similarly, effon

could be put into producing an electrode material with very high quality interfacial

bonding, so that the porosity effects could be studied. The overall plan would be to

reduce the number of variables involved.

The procedures for resistivity, wettability and real density appear to be sound, and are

ready for use on further carbon materials (although the wettability test can only be used

on materials where both the binder and filler are available). The area of relating

electrical resistivity to the microscopic structural characteristics of carbons appears to be

one in which very little work has been done; given that low resistivity is a desirable

feature of an electrode, further investigation in this field should be of both theoretical

interest and practical value.

Ultimately, it is likely that further work of the type undenaken in this project could lead

to the development of a more reliable means of manufacturing electrodes with both high

mechanical strength and low electrical resistivity, following further acquisition of data

about the microscopic structural features of electrode carbons.

128

~ N , E E

Figure 5.1 - Cumulative Results for all Carbons

45000-r---------------------.

40000

• 35000

• 30000 •

~ 25000

~ Z vi

20000

'5000

'0000

• 50001+---~-~--~--~-~--_r-~

0.08 0.10 0.12 0.14 0.16 0.18 Wjp2 (J.Lm -')

129

0.20 0.22

Figure 5.2 - Cumulative Results for all Specimens

70000 0

X

60000

0 X 0

/::,. 50000

f ~X

/::"0

• • ~ 40000 N

/::,.W 0 , E E • " "- X

:::E \l ~

~ 30000 'V • x Legend

20000 -A4

o C

• P

~ 0

10000 0 0 AL6 ... -0 /::,. ALB

o- X y

\l Z 0 , .

0.05 0.10 0.15 0.20 0.25 0.30

W/p2 Cum -')

130

6 References

1. lones, S.S., Anode Usage in the Aluminium Industry, in 'Petroleum-derived

Carbons', eds. 1.0. Bocha, I.W. Newman and 1.L. White, ACS Symposium

Series No. 303, ACS, Washington, USA, 1986, p234.

2. Ragan, S. and Marsh, H., Carbon, 1986,1, pI.

3. Marsh, H., 'The Production of Liquid Aluminium', 25th Annual Conference of

Metallurgists, 1986.

4. Taylor, G.H., Fuel, 1961, 40, p465.

5. Brooks,I.D. and Taylor, G.H., Carbon, 1965,2., p183.

6. Gibson, 1. and Gregory, D.H., 'The Carbonisation of Coal', Mills and Boon

Monographs in Chemical Engineering, London, 1971.

7. Marsh, H., Fuel, 1973,ll. p206.

8. Brooks, 1.0. and Taylor, G.H., 'The Chemistry and Physics of Carbon', Vol. 4,

Ed. P.L. Walker, Marcel Dekker, New York, 1968, pp243-272.

9. Brooks, 1.0. and Taylor, G.H, Nature, 1965,1Q2, p697.

10. Whittaker, M.P. and Grindstaff, L.I., Carbon, 1972,.ill, p165.

11. Ihnatowicz, M. ~ ai, Carbon, 1966, 1, p41.

12. Bradford, DJ. et ai, Proceedings of the Third Conference on Carbons and

Graphites, Society of the Chemical Industry, London, 1971, p520.

13. Dubois et aI, Metallography, 1970,2., p337.

14. Brooks, 1.0. and Taylor, G.H., 'The Chemistry and Physics of Carbon', Vol. 4,

Ed. P.L. Walker, Marcel Dekker, New York, 1968, p52.

15. Lewis, I.C., Carbon, 1982,20,.2, p519.

131

16. Honda. H .• Carbon. 1988. 26.~. p139.

17. Hays. D .• Patrick. I.W. and Walker. A .• Fuel. 1976.~. p299.

18 lones. S.S and Hildebrandt. R.D .• 'Light Metals 1975'. AlME. p291.

19. Honda. H. et al. Carbon. 1970. R. p 181.

20. Mochida. I.. Fuel. 1981. QQ. p405.

21. White. I.L.. 'Petroleum-derived Carbons'. American Chemical Symposium

Series. 1976. 21. p282.

22. Mochida. I.. Oyama. T. and Korai. Y .• Carbon. 1988. 26. 1. p49.

23. White. J.L. and Price. RJ .• Carbon. 1974.ll. p321.

24. Mason. C.R .• Proceedings of the Fifth London Carbon Conference. London.

1978. p344.

25. Wagner. P. and Dauelsberg. L.B .• Carbon. 1969. L p273.

26. Hays. D .• Patrick. I.W. and Walker. A.. Fuel. 1983.~. p946.

27. Auguie. D .• Oberlin. M. and Oberlin. A .• Carbon. 1981. 19.1. p277.

28. Dell. M.B .• Fuel. 1959 • .la. pl83.

29. Wagner. P. et al. Fuel. 1988.~. p946.

30. Briggs. D.K.H .• Fuel. 1980. ~. p201.

31. Hays. D .• Patrick. I.W. and Walker. A .• Fuel. 1983. 62. p1145,

32. Marsh, H. and Sherlock. I .• Fuel. 1981. 60. p434.

33. Greenhalgh. E. and Moyse. M.E .• Proceedings of the Third Conference on

Carbons and Graphites. Society of the Chemical Industry. London. 1971. p539.

34. Ehrburger. P. and Lahaye. J,. Fuel. 1984. 63. pI677.

132

35. Couderc, P., Hyvennet, P. and Lemarchand, J.L., Fuel, 1986,~, p281.

36. Heintz, E.A., Carbon, 1986, 24, ~, P 131.

37. Lahaye, J. and Ehrburger, P., Fuel, 1985, 64, p1187.

38. Hays, D., Patrick, J.W. and Walker, A., Fuel, 1982, ~ p232.

39. Marsh, H., Forrest, M. and Pacheco, L.A., Fuel, 1981, 2Q, p423.

40. Patrick, J.W., Shaw, F.S. and Wilmers, R. R., Fuel, 1977, ~,p8I.

41. Lewis, I.C. and Singer, L.S., The Chemistry and Physics of Carbon', Arnold,

London, 1968, Volume 17, pI.

42. Meers, lM., Encyclopaedia of Chemical Technology, Third Edition, John WHey

and Sons, New York, 1978, Vol. 4, p589.

43. Stadelhofer, J.W., Marrett, R. and Gemmeke, W., Fuel, 1981, @, p877.

44. Lang, I and Vaurecka, P., Fuel, 1981, QU, p1176.

45. Rallouch, R.W. and Fair, F.V., Carbon, 1980, il, p147.

46. Mantell, c.L.. Carbon and Graphite Handbook, WHey Interscience, New York,

1968, p156.

47. Ragan, S. and Marsh, H., Journal of Materials Science, 1983, ll., p3161.

48. Chard, W., Con away, M. and Niez, D., 'Petroleum-derived Carbons', in

American Chemical Society Symposia Series No. 21. Eds. M.L. Deviney and

T.M. O'Grady, ACS, Chapter 14, Washington, USA, 1976.

49. Gordon, J.E., 'The New Science of Strong Materials', Second Edition. Penguin,

London, 1974.

50. Inglis, C.E., Trans. Naval Archit., 1913,~, p219.

51. Griffiths, A.A., Phil. Trans. R. Soc. London. 1920, A221, p163.

133

52. Patrick, J.W. and Walker, A., J. Mat. Sci., 1987, 22, p3589.

53. Smith, M.C., USAEC Conference Report No.70115, 1972, p475.

54. Gilchrist, K.E. and Wells, D., Carbon, 1969,1, p627.

55. Patrick, J.W. and Stacey, A.E., Fuel, 1972, January,.ll, p81.

56. Patrick, J.W. and Stacey, A.E., ISS Transactions, 1983, J, pI.

57. Larsen, B., Patrick, J.W., Sprlie, M. and Walker, A, Proceedings Carbon '87,

p393.

58. Patrick, J.W., Sprlie, M. and Walker, A, Proceedings Carbon '88, p461.

59. Wood, J.L., Bradt, R.C. and Walker, P.L., Carbon, 1980, ll, p169.

60. Knudsen, F.P., J. Amer. Ceram. Soc., 1959, 42, p376.

61. Orowan, E., Progress in Physics, 1949,.ll, p185.

62. Espinola et aI, Carbon, 1986, 24, p337.

63. Kupperman, D.S., Chau, C.K. and Winstock, H., Carbon, 1973, ll, p171.

64. Huttinger, KJ., Carbon, 1971,2, p809.

65. Hemandez Sll J!!, Carbon, 1982, 2Q, p201.

66. Sakai, M., Hirota, S. and Nagaki, M., Carbon, 1984, 22, 187

67. Sharp, J.A, Fuel, 1987, November, 66, p1487.

68. Blayden, H.E., Yearbook of the Coke-Oven Managers' Association, 1966, p197.

134