Sie sind auf Seite 1von 12

FULL PAPER

www.steel-research.de

The Hot Strength of Industrial Cokes Evaluation


of Coke Properties that Affect Its HighTemperature
Strength
Juho A. Haapakangas, Juha A. Uusitalo, Olli J. Mattila, Stanislav S. Gornostayev,
David A. Porter, and Timo M. J. Fabritius

The strength of coke at high temperatures is of major importance for efficient blast furnace
operation. Despite this, there is little information regarding the hot strength of coke. In this
research, the hot strength of three industrial European coke grades was studied using a
Gleeble thermomechanical simulator. The hot strength was analyzed at three different
temperatures: room temperature, 1600 and 1750 8C by measuring the compressive strength
of roughly 50 coke samples at each temperature. A significant decrease in strength was
observed for all three coke grades at high temperatures. Notable differences between the
coke grades in compressive strength were observed at room temperature. At high
temperatures, differences were observed in strength and deformation behavior; however,
the order of magnitude of strength remained the same. The deformation behavior at high
temperatures was also studied and discussed based on stressstrain curves. Several
structural properties of coke such as total porosity, pore size distribution, pore shape factor,
amount of inerts, and degree of graphitization were determined in order to explain the
obtained strength results. Both hot and room temperature compressive strength values
were compared to industrial strength tests and discussed. The reliability of the results was
evaluated with statistical analyses.

1. Introduction
Coke as a raw material in the blast furnace provides fuel for
ironmaking, serves as a source of reductant gas, provides
structural support for the material bed and serves to
carburize hot metal. Its role as structural support is vital
and for this reason strength is one of the most important
properties of coke. Poor coke strength can cause many
operational problems, such as reduced permeability in the
shaft and hearth areas, undesirable gas and temperature
distribution, and the possibility of hanging of the burden.
In the near future, steel manufacturers worldwide are
faced with challenges regarding the quality of coke. The
availability of prime coking coals is decreasing and their

[] J. A. Haapakangas, S. S. Gornostayev, Prof. T. M. Fabritius


Laboratory of Process Metallurgy, University of Oulu, P. O. Box 4300,
FI-90014 Oulu, Finland
Email: juho.haapakangas@oulu.fi
J. A. Uusitalo, Prof. D. Porter
Materials Engineering Laboratory, University of Oulu, P. O. Box 4200,
FI-90014 Oulu, Finland
O. J. Mattila
Ruukki Metals Oy, Rautaruukintie 155, FI-92100 Raahe, Finland
DOI: 10.1002/srin.201300450

1608

steel research int. 85 (2014) No. 12

price is increasing, so coke manufacturers will have to


utilize coals of lower quality. One option could be the prepelletizing of weakly coking coals.[1] Also, the high price of
metallurgical coke is forcing companies to find alternative
fuels for the blast furnace. A common way to replace coke
is to increase the use of injectant fuels, commonly
pulverized coal. When less coke is charged, greater
chemical stresses are faced by individual lumps of coke.
Coke also undergoes higher load due to thicker iron ore
layers and rise of average burden weight. Another issue
is the rising price of CO2 emissions. One possibility to
reduce CO2 could be the addition of bio-material or
waste plastics into the coking mix, which can negatively
affect the strength of the coke. All of these factors
increase the importance of coke quality and its analysis
methods.
Stresses inside a blast furnace include mechanical
stresses such as shattering and abrasion, solution loss
reactions, alkaline attack, high temperature, and impact
shock by the high-speed hot blast. In industry, coke
strength is characterized by drum indices (Micum, Irsid)
or drum strength after reaction (CSR). Although these
methods are useful in assessing coke quality, they do not
measure coke strength at operational temperatures. Blast
furnace excavations made in Japan have shown that coke

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.steel-research.de

graphitization. Only visibly crack-free samples were


selected for strength testing. The method of sample
preparation is shown in Figure 1.
During strength testing, the Gleeble samples were held
for 30 s at their respective test temperatures to allow the
temperature to stabilize throughout the samples. After
heating, compressive forces were measured while the
samples were mechanically compressed 4 mm at the rate
of 1 mm s1. The use of a Gleeble device for testing the hot
strength of coke has been described in detail earlier.[7,8]
Industrial analyses provided by the coke manufacturers
are displayed in Table 1 and 2.
The XRD tests were performed with a Siemens D5000
X-ray diffractometer. The point counting of structural
properties was performed with a Pelcon Automatic Point
Counter.

3. Results and Discussion


3.1. Hot Strength
During previous work[7,8] coke hot strength was measured
for an individual coke grade and a significant decrease in
strength was observed at 1600 8C in comparison to 1000 8C.
Afterwards an even further decrease in strength was
observed at 1750 8C. Deformation of coke was also
reported to be partially plastic at 1600 8C and plasticity
further increased at 1750 8C. In this study, two new coke
grades were included in the testing.

2. Experimental
All of the coke samples used in the Gleeble experiments
were prepared beforehand and mixed in order to avoid
systematic errors caused by sample location. Coke lumps
were drilled with a hollow drill, cut to appropriate height
and polished to form cylindrical samples of 16 mm in
diameter and roughly 12 mm in height. This size and shape
has been found appropriate for use in the Gleeble device
using a standard sample holder. The largest coke lumps
were selected for sample preparation as small ones are
inadequate for producing a sufficient number of samples.
After machining, the coke samples were pre-graphitized
in a chamber furnace in order to speed up testing in the
Gleeble by reducing the required sample holding time. The
pre-graphitization was made at the sample temperature
that was subsequently used for testing in the Gleeble. It
was done in order to simulate the thermal stresses and
degree of coke graphitization brought about by the descent
of the coke in an actual blast furnace. The coke samples
were sealed in a graphite container and covered in graphite
powder to avoid gasification. In addition, the furnace was
operated in inert gas. The graphite container containing
the coke samples was placed in the furnace at 1200 8C after
which the temperature was raised up to the subsequent
test temperature and held for 60 min to ensure complete

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 1. A) A coke cylinder after drilling, B) a prepared coke


sample, C) the graphitization container, and D) coke sample
during a Gleeble test.

steel research int. 85 (2014) No. 12

1609

FULL PAPER

size remains mostly unchanged in the upper parts, but


significant degradation is observed in the lower parts when
the temperature exceeds 1400 8C.[2]
There are few published studies on the hot strength of
coke, possibly due to difficulties in experimental setup.
The reported hot strength results are somewhat contradictory: increases in compressive coke strength have been
reported at 1200,[3] 1300,[4] 1400,[5] and 1650 8C[6] while a
decrease in coke strength was reported at 2000 8C.[4] It
should be noted that most of the earlier results have not
been statistically reliable due to small sample sizes. The
properties of coke that define its hot strength are largely
unknown.
In this study, the study of compressive strength of
coke was further expanded to include three industrially
produced European coke grades at room temperature and
at elevated temperatures of 1600 and 1750 8C. Also in
this study several coke properties, such as total porosity,
inerts, and degree of graphitization were analyzed in
order to study how they relate to the strength of coke
both and room temperature and at elevated temperatures.
Both the room temperature and hot strength measurements were conducted with a Gleeble 3800 thermomechanical testing device. The obtained results were
intended to give information about the strength of
industrial coke in the lower part of a blast furnace, where
coke is the only solid material and its strength is vital for
good permeability.

FULL PAPER

www.steel-research.de

Coke 1

Irsid
> 20 mm

Irsid
< 10 mm

77.2

21.1

Micum
> 40 mm

Micum
< 10 mm

CRI

CSR

Volatiles
[%]

21.5

63.6

0.2

Coke 2

85.0

5.4

31.3

57.6

0.2

Coke 3

82.7

5.8

28.2

58.5

0.2

Table 1. Industrial analyses of the studied coke grades.

Coke 1

Total ash
[%]

SiO2

Al2O3

CaO

Fe

MgO

K2O

Ti

Na2O3

10.6

59.90

26.70

2.45

3.91

0.72

0.76

1.28

0.88

0.38

Coke 2

8.69

44.88

29.92

4.37

7.02

2.30

0.53

2.42

0.81

1.61

Coke 3

9.20

47.83

31.52

3.48

5.65

1.74

0.52

2.28

0.87

1.52

Table 2. Ash component contents.

The strength of three coke grades was measured at room


temperature and at elevated temperatures of 1600 and
1750 8C. Originally 50 coke samples were tested for each
grade at each temperature in order to obtain statistically
reliable results, however some of the results had to be
discarded due to equipment malfunction. The strength
results are listed in Table 3. In Table 3 the superscript
numbers 7 and 8 represent results obtained during
previous work. In the table, the term ultimate strength
refers to the highest obtained strength value during the

4 mm (33.3%) compression. In some cases, this was


obtained in the beginning of compression and in some
at the very end.
The mean strength and strain values at room temperature were similar to those reported elsewhere: Kim and
Sasaki[9] reported compressive strength of coke to vary
between 19.4 and 28.6 MPa, while strain at fracture ranged
from 2.2 to 3.0%. As observed in Table 3, there were notable
differences in the compressive strength of the coke grades
at room temperature; coke 2 was clearly the strongest and
Effect of heating
on the ultimate
strength [%]

Number of
samples

Mean
ultimate
strength [MPa]

Strain
[%]

Standard
deviation
[MPa]

Room

50

16.67)

2.5

7.57)

1600 8C

50

12.17)

3.67)

27.0

1750 8C

50

10.78)

3.38)

35.5

Room

50

23.8

1600 8C

40

15.8

5.2

33.6

1750 8C

50

17.5

6.2

26.5

Room

49

13.3

1600 8C

50

9.1

3.2

31.5

1750 8C

50

9.7

3.3

27.1

Temperature

Coke 1

Coke 2
3.0

8.1

Coke 3
2.5

6.7

Table 3. Room temperature and hot strength values.

1610

steel research int. 85 (2014) No. 12

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.steel-research.de

be the softening of bonding between carbon layers as a


result of increased heat energy, which would allow reorganization of the carbon matrix during compression.
It is noteworthy that none of the three coke grades
tested during this study suffered a catastrophic fall in
strength as a result of high temperatures. This result was
expected since all three cokes are used industrially and
known to function in actual blast furnace operation. The
significant fall in the strength of coke at high temperatures
is contradictory to some of the previous reports by other
researchers.[3,5,6] A possible explanation for this could be
the differences in the coke grades that were analyzed,
different methods of sample preparation (pre-graphitization) or the use of different testing equipment. Some of the
studies were also conducted at temperatures below the
melting point of cokes ash components, therefore not
taking into account their effect. The reported significant
fall in strength at 2000 8C and above, and the observed
plastic deformation behavior,[4] however, are congruent
with results of this study.
Statistical reliability of the hot strength results was
analyzed using a two-sample t-test assuming unequal
variances (Welchs t-test). A two-tailed test with p 0.05
was chosen as the level of significance. The following
hypotheses were set and evaluated:
H0: There is no statistical difference in coke hot strength
between the studied temperatures.
H1: There is a significant difference in the hot strength of
coke at the studied temperatures.
When the calculated value of t exceeds the critical value
of t and p is below the chosen significance level of 0.05, the

Degrees of
freedom

Critical t

p-Value

Statistically
significant

Room vs. 1600 8C

70

3.790

1.994

0.000

Yes

Room vs. 1750 8C

67

5.043

1.996

0.000

Yes

1600 vs. 1750 8C

97

2.019

1.985

0.046

Yes

Room vs. 1600 8C

84

5.648

1.989

0.000

Yes

Room vs. 1750 8C

91

4.328

1.986

0.000

Yes

1600 vs. 1750 8C

87

1.443

1.988

0.153

No

Room vs. 1600 8C

68

3.955

1.995

0.000

Yes

Room vs. 1750 8C

70

3.394

1.994

0.001

Yes

1600 vs. 1750 8C

97

0.872

1.985

0.386

No

Temperature

Coke 1

Coke 2

Coke 3

Table 4. Statistical reliability analysis of the hot strength values.

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

steel research int. 85 (2014) No. 12

1611

FULL PAPER

coke 3 the weakest. The difference in the reported drum


strength between cokes 2 and 3 in the Micum test, given in
Table 1, was much less pronounced. Unfortunately, only
Irsid values were available for coke 1.
At 1600 8C, the ultimate strength was significantly
decreased for all coke grades compared to room temperature. The decrease in strength was similar for all three
coke grades. An increase in test temperature had a varying
effect depending on the coke grade; the strength of
coke 1 suffered a further significant drop when heated
to 1750 8C. The ultimate strength of cokes 2 and 3, on the
other hand, seemed to slightly recover as a result of
increasing the temperature, however, as will be shown
in Table 4 this result was without statistical significance.
The possible recovery in strength could be explained by
the increasing plasticity of the coke at higher temperatures. As a result, the density and surface area of coke
resisting the compressive force are increased during
deformation resulting in a higher force required for
further deformation. Plastic deformation behavior of
coke at high temperatures has also been reported in a
previous study.[4] The results indicate that the development of coke hot strength as a function of temperature
may be material dependent. Despite the varying effects
of temperature, the relative strengths of the various coke
grades remained constant independent of temperature
(2 > 1 > 3).
The reason behind plastic deformation behavior of
coke at high temperatures is unknown. It is the belief of
the authors that melting of cokes ash components is not
responsible, since it has been shown that in tuyere coke the
melted ash has migrated into pores, therefore it should
not affect deformation behavior.[10] One explanation could

FULL PAPER

www.steel-research.de

difference in strength is concluded to be significant. Based


on the results presented in Table 4, the hypothesis H0
was accepted in two cases: both for cokes 2 and 3 when
comparing the hot strength between temperatures 1600
and 1750 8C. Therefore, the differences in strength between
these temperatures are statistically insignificant. In all of
the other cases, hypothesis H0 was rejected and the
alternative hypothesis H1 was accepted; the difference in
the other studied temperatures is statistically significant.
Digital stressstrain curves were obtained during the
strength tests both at room temperature and at high
temperatures. The curves can be used for analyzing the
deformation behavior of the cokes. Six general types of

stressstrain curves were identified during the tests and


they are presented in Figure 2. For clearer representation,
the actual stressstrain curves obtained from the
Gleeble experiments were re-calculated based on moving
average of 30 measurement points (one measurement
every 0.01 s). It should be noted that the strain values
in Figure 2 do not always precisely reflect the true strain
values since the graphs were re-calculated based on
moving average. The difference can be up to roughly 2%
during the linear elastic phase. These stressstrain type
categories were found to represent the vast majority of all
strength experiments both at room temperature and at
high temperatures.

Figure 2. General stressstrain curve types (AF) observed during the compressive strength tests for coke (see text for further details).

1612

steel research int. 85 (2014) No. 12

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.steel-research.de

Type A
[%]

Type B
[%]

was also regularly observed at room temperature, which


could be a sign of layered crushing as reported in a
previous study.[11]
The type D curve, like type C, had a steep linear rise in
the stress curve followed by fracture and a decrease in the
stress curve, however, at high temperatures toward the end
of compression the stress started to strongly increase.
Types E and F both had a very gentle rise in stress curve
during the linear elastic phase. Based on the shape of
the curve, this was followed by fracture half way through
the compression, i.e., about roughly 15% strain and a
decrease in stress required for further deformation. Type F
showed a second increase in the stress curve toward the
end of indicating plastic deformation during compression
and often the highest force value was obtained at the very
end. Plastic deformation behavior (types B, D, and F) could
be considered beneficial for blast furnace performance,
since the strength of coke would increase under compression at high temperatures.
As was reported in Table 3, the standard deviation of
strength showed a strong decrease at high temperatures.
This can also be observed from the distribution of all
results to specific strength regions, which are depicted in
Figure 3.
The distribution of coke strength becomes increasingly
narrower as the temperature is increased. This result
was especially pronounced for coke grades 1 and 3. For
coke 2, the standard deviation decreased at 1600 8C but
was slightly higher at 1750 8C compared to 1600 8C.
This result could be explained by the grades possibly
stronger tendency toward plastic deformation as indicated
by the mean strength values. It can be considered that
a narrow strength distribution would be desirable, since

Type C
[%]

Type D
[%]

Type E
[%]

22.0

14.0

Type F
[%]

Room temperature
Coke 1

76.0

24.0

Coke 2

88.0

12.0

Coke 3

72.0

28.0

1600 8C
Coke 1

2.0

Coke 2
Coke 3

40.0

22.0

15.0
18.0

67.5

17.5

46.0

16.0

16.0

4.0

1750 8C
Coke 1

42.0

18.0

4.0

30.0

6.0

Coke 2

4.0

10.0

4.0

62.0

20.0

56.0

12.0

22.0

6.0

Coke 3

4.0

Table 5. Distribution of stressstrain curves among the six general types.

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

steel research int. 85 (2014) No. 12

1613

FULL PAPER

The distribution of all stressstrain curves to the six


general types is presented in Table 5. As observed from the
table, each coke grade was unique in terms of the shape of
the stressstrain curves at high temperatures. At room
temperature, only curve types A and C were observed.
Meanwhile, types B, D, E, and F, which indicated plastic
deformation, were only found at high temperatures.
At room temperature, the ultimate strength values were
obtained at the very beginning of deformation for all three
grades. On a stressstrain curve this is reflected by an
initial almost linear rise in compressive force until fragile
fracture occurs at 04% strain (mean 2.53.0%) and in most
cases followed by a significant decrease in force required
for further deformation. This type of a stressstrain curve is
type A in Figure 2. Therefore, it can be stated that
deformation behavior was brittle in all room temperature
cases and the obtained stressstrain curves resembled
the deformation of a ceramic material, a result, which has
been reported in previous studies.
The type B curve was observed at high temperatures and
can be considered to indicate plastic deformation during
compression: as the coke sample is compressed, its density
and surface area are increased and force required for
deformation is increased. For type B curves, the ultimate
strength value was obtained at the very end of the 4 mm
compression. This type was observed for cokes 1 and 2 at
high temperatures.
Type C curve is similar to type A with a steep linear rise
in strength, but unlike for type A, there is no dramatic fall
in the force required for further deformation, instead
significant force is still required all the way up to 33%
compression. This type of a curve was common at
higher temperatures for coke 1 and especially coke 3. It

FULL PAPER

www.steel-research.de

blast furnace operators value consistent quality and


predictability. Strength of coke 3 is, however, clearly the
highest at all temperatures with a low number of weak
samples. Therefore, out of the tested cokes, the distribution of coke 3 can be considered the most desirable.
The industrial CSR (coke strength after reaction) test is
often called a hot strength test despite the fact that the
actual drum strength is measured at room temperature
after 2-h gasification at 1100 8C in 100% CO2 gas. Based on
the industrial tests for coke depicted in Table 1, coke 1 had
by far the highest CSR value, likely a direct result of its
lower chemical reactivity (CRI). The high CSR value is
usually interpreted as indicating good performance in a
blast furnace. However, it has been discussed that the
conditions during the CSR test are intensive and the
reaction mechanism differs from an actual blast furnace:
during the CSR test coke is gasified throughout the coke
whereas in a blast furnace reactions occur mainly on the
surface.[12] Despite the usefulness of CRI and CSR tests,
they do not include testing coke at blast furnace temperatures. Since no single analysis method is capable of
simulating blast furnace conditions, multiple different
tests are required. The use of a Gleeble could be a useful
addition to the coke testing protocol.
The most authentic strength test could be testing pregasified coke in a thermomechanical simulator, such as
the Gleeble. The pre-gasification could be performed in a
blast furnace gas phase simulator, such as presented in
literature.[13] The tested coke could be also taken from a
dissected blast furnace or an experimental blast furnace,
in which case it would also include the presence of
circulating elements. Including coke gasification to hot
strength testing, however, could not be fit to this article
and deserves a separate investigation.

3.2. The Effect of Heat Treatment on the Room


Temperature Strength
Heat treatment above the temperature of the coking
furnace is expected to have a twofold influence on the
strength of coke: (i) it promotes further graphitization,
which changes the atomic structure of coke into a more
highly ordered one and (ii) it causes significant weight loss
as displayed in Table 6.

Figure 3. Distribution of coke strengths.

1614

steel research int. 85 (2014) No. 12

Grade

1600 8C
[%]

1750 8C
[%]

Coke 1

7.2

10.1

Coke 2

4.9

7.6

Coke 3

6.2

8.7

Table 6. Weight loss of coke as a result of 1 h pre-graphitization in


a chamber furnace.

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.steel-research.de

FULL PAPER

As was shown in Table 2, coke 1 contained the highest


amount of SiO2, therefore its ash had the highest potential
to react with the carbon matrix (forming SiC and CO) and
to induce weight loss. The weight loss is a result of
chemical reactions, vaporization, and re-organization of
mineral matter in the coke (i.e., decarbonation and
desulphurization[14]). Another issue is the plastic deformation at elevated temperatures. For this reason, a test
was made to study how much merely the preceding
heat treatment affects the room temperature strength
of coke in comparison to it being in a hot state. This
study was done for coke 2 by testing the strength at
room temperature of 50 non-graphitized samples and
50 samples pre-graphitized at 1750 8C. The obtained
strength values are displayed in Table 7. Youngs modulus
was determined from the angular coefficient of linear
elastic part of the stressstrain curves.
As can be seen from Table 7, the mean room
temperature strength of coke is significantly decreased
by pre-graphitization. As with strength tests at elevated
temperatures, the standard deviation of strength is
much lower. From Figure 4, it can be concluded that
the structure of coke undergoes a process of homogenization during heat treatment. In particular, samples
showing high strengths are absent after graphitization.
The result of a decreased strength is likely a combination of several factors including weight loss and a
resultant increase in porosity, as well as changes in the
crystalline structure of the coke. It has been stated
that as the crystalline structure approaches that of
graphitic carbon, the cross-linking between carbon
layers is decreased and the coke matrix weakened.[15]
The above results support this theory, however the
weight of each factor is unknown and deserves separate
investigations.
For both graphitized and non-graphitized samples
tested at room temperature, the breakage mechanism
was the following: a short elastic phase followed by fragile
fracture. Heat treatment at 1750 8C did not have any
noticeable effect on the shape of stressstrain curves,
except for the angular coefficient of the elastic phase
(Youngs modulus). There was little to no change in the
strain value following heat treatment. Therefore, it is
expected that the plasticity of coke observed at high
temperatures is purely a result of the temperature during
the compression, not the preceding heat treatment.

Figure 4. Distribution of strength of coke 2 samples before and


after heat treatment at 1750 8C.

3.3. Graphitization Degree of Coke


Graphite is a crystalline structure of carbon that consists of
regular, vertical stacking of hexagonal aromatic layers.
When coke is heated to a sufficiently high temperature,
it begins to become more ordered and approach the
structure of graphite. Further graphitization of coke begins
to occur when its temperature exceeds the original coking
temperature; generally above 1100 8C. Graphitization is
made possible by the plastic phase of coking, during which
carbon layers are organized near parallel. As the temperature of coke is increased, it enables the continuous
rearrangement of the layer-planes to take place by small
stages.[16]
In the structure of graphite, each C atom within an
aromatic layer is linked through covalent bonds to three C
atoms. However, bonding between aromatic layers is
weak, easily broken by external forces. Non-graphitic
carbon, however, contains cross-linking between the
aromatic layers and much higher force is required to
dissociate them.[15,17] During heat treatment, non-organized carbon is presumably attached to the edge atoms
of the graphite like layers, which enables the growth of
organized layers but decreases the cross-linking between

Mean
strength
[MPa]

Standard
deviation
[MPa]

Strain
[%]

Youngs
modulus
[MPa]

Non-graphitized

23.75

8.09

3.0

1337.1

Pre-graphitized (1750 8C)

15.32

5.42

3.1

724.3

Change in
strength
[%]

35.5%

Table 7. Compressive strength values of coke 2 before and after heat treatment at 1750 8C.

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

steel research int. 85 (2014) No. 12

1615

FULL PAPER

www.steel-research.de

the layers. After heat treatment between temperatures of


1000 and 2000 8C non-graphitizing carbons become harder
and graphitizing carbons become softer.[16] Therefore, it is
important to know the degree of coke ordering, which is
unique for different coke grades.[18]
Some coals are easily graphitized, others are nongraphitizing and some, such as anthracite, begin to
graphitize at extremely high temperatures (above
2000 8C). Graphitization is inhibited by random orientation
of the graphite like layers as well as strong cross-linking
between the layers.[16] Coke grades have also been
observed to have varying degrees of graphitization even
in similar annealing temperatures.[18]
The size of graphite crystallites in coke can be
represented by the height of carbon net layer (Lc) and
the spread of carbon net plane (La). The growth of Lc during
graphitization is linear as a function of temperature and
is independent of the gas atmosphere and gasification
reaction. La also increases during heating, but is decreased
during gasification.[19] The graphitization degree of coke
can be evaluated by X-ray diffraction method. In this
work, graphitization degree was evaluated by calculating
the Lc values based on the 002 carbon peaks from the
XRD patterns. Calculating the La values is also possible,
but more difficult due to its smaller intensity in the
XRD spectra. The Lc value can be calculated from X-ray
diffraction profiles using Scherrers equation:[17,19]


Lc A

0:9l
B cos uB

1


where l is the wavelength of X-ray (A ), B is the angular


width in radians at half-maximum intensity of [002] peak,
and uB: is the reflection angle of [002] peak.
The calculation is based on the assumption that
crystallites are ideal with no strain or distortion. The
results can be considered precise only for ordered
materials. Images of the 002 peak in the XRD spectra
are displayed in Figure 5.
The Lc value illustrates the average height of graphite
crystals in coke, hence a high value indicates a high degree
of graphitization. The Lc values of the studied coke grades
are presented in Table 8.
The calculated Lc values indicate further graphitization
as a result of heat treatment for all three coke grades.
Similar observations have been made in previous studies.
The coke grades also showed similar Lc values at a given
temperature, therefore they are expected to be similar in
terms of graphitization behavior brought about by high
temperatures.
Strength and degree of graphitization of coke may
indeed be strongly linked: for all three coke grades strength
was significantly decreased at high temperatures as the
degree of graphitization was increased. This observation is
also supported by the fact that the room temperature
strength was significantly lower for coke heat-treated at
1750 8C. However, the significant weight loss and following

1616

steel research int. 85 (2014) No. 12

Figure 5. Shapes of the 002 peaks from the XRD spectra of the
studied coke grades. The background intensities of each curve are
scaled for clearer recognition of the shapes.

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.steel-research.de

Coke 2

[A ]

Coke 3

[A ]

Untreated

21.8

20.0

16.9

1600 8C

59.8

62.5

62.8

1750 8C

93.2

81.9

88.9

Porosity
[%]

Inerts
[%]

Reactive
texture
[%]

Coke 1

49.0

17.4

33.6

Coke 2

46.0

17.6

36.5

Coke 3

48.1

16.8

35.2

Table 8. Lc values of coke after heat treatment.


Table 9. Total porosity, contents of inert, and reactive texture.

increase of porosity during heat treatment should also be


kept in mind as a weakening factor. In this study, all of the
coke grades proved to be graphitizing and very close in
terms of degree of graphitization at each temperature.
Therefore, their differences in strength at a given
temperature are probably decided by other factors.

3.4. Effect of Coke Properties on Their Physical Strength


The strength of coke can be divided into two factors: the
adhesive strength of the solid matrix and the pore
structure. Lower total porosity is beneficial for coke
strength since high porosity indicates a reduction in the
amount of solid material. This result has been observed in
previous publications.[5] According to results of 3D
modeling, total porosity is dominant compared to carbon
matrix bonding strength regarding the mechanical
strength of coke.[9] Pore size distribution is also considered
important, since it affects the distribution of stresses. It has
been concluded that small pore size and high density of
pores produce higher strength compared to a small
number of large pores.[5,20] Thick coke cell walls in theory
should be beneficial for strength, however in practice a
high density of small pores tends to produce a thin cell
wall. Regular distribution of pores has also been suggested
to be beneficial for good strength as cracks can propagate
more easily through connected pores.[9,21]
A high amount of inerts in coke is another factor that has
been linked with good strength.[20] High amounts of inerts
have also been reported to decrease the mean pore size,
which may partially explain their positive effect on
strength.[20] A decrease of inert size can improve coke
drum strength up to a certain limit.[22] Inert macerals also
have higher hardness than reactive macerals. The hardness
of reactive macerals decreases with increasing mosaic
size.[20] Differences exist in the hardness of various coke
textures.[20] However, despite some attempts to link coke
texture with mechanical strength, correlations between the
two have been poor and it is possible that the effects of
texture distribution are overshadowed by other factors.[23]
In order to explain the differences in strength between
coke grades, total porosity and the contents of inert and
reactive texture were determined. The distribution of these
structural parameters was done by a standard point

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

counting method by counting at least 1000 points from


five polished sections for each coke grade. Structural data
obtained by these analyses is displayed in Table 9.
Differences in the content of inert material between the
coke grades proved to be small. The high strength of coke 2
reported in Table 3 could be at least partially explained by
its slightly lower total porosity. However, all of the studied
cokes were fairly close in measured structural characteristics and due to the heterogeneous nature of coke
substantial amounts of polished sections would have to
be analyzed in order to draw certain conclusions.
Pore size distribution and pore shape factors were
determined with a MATLAB-based image analysis software[24] using the same polished sections used for point
counting. Five images of size 3.5 mm  2.6 mm were taken
from each five polished sections per grade under 4
magnification. The pore shape factor was determined by
identifying pore edges from LOM images by subsequent
thresholding and filtering methods causing the smallest
pores to be excluded from the analysis as they could not be
separated reliably from the darker areas of the coke matrix
itself. After the pore edge detection, the measured pore
surface area was used to determine round-equivalent edge
length, which was divided by measured real edge length on
each pore. The software was programmed to ignore any
pore connected to the edge of an image in the calculations.
According to the image analysis results shown in
Table 10, there was practically no difference in the pore
shape factors in any size range. It can be concluded that in
this study no direct correlation was obtained between coke
strength and pore shape factors of pores below 1 mm in
diameter. It is thought by the authors that the effects of
pore shapes on coke strength are caused by flaws,
connected pores, and pores of larger size than the ones
measured during this study.

4. Conclusions
The compressive hot strength of three industrially
produced coke grades was analyzed by testing 50 coke
samples at room temperature, 1600 and 1750 8C. Statistical
reliability of the obtained strength results was analyzed.
The strength testing was performed with a Gleeble 3800

steel research int. 85 (2014) No. 12

1617

FULL PAPER

Coke 1

[A ]

FULL PAPER

www.steel-research.de

Pore edge
length [mm]

Number of pores

Pore shape factor

Coke 1

Coke 2

Coke 3

Coke 1

Coke 2

Coke 3

0.200.41

51.52

65.12

72.08

0.97

0.97

0.98

0.410.82

22.44

30.12

32.44

0.85

0.84

0.86

0.821.22

6.20

8.12

7.24

0.72

0.73

0.73

1.221.63

3.39

3.72

4.32

0.67

0.66

0.66

1.63

5.64

6.12

5.56

0.55

0.54

0.55

Table 10. Pore size distribution and pore shape factors obtained by image analysis.

thermomechanical simulator. Properties of coke, such as


total porosity, content of inerts, and reactive macerals as
well as degree of graphitization were determined in order
to explain the observed strengths. The measured strengths
were also compared to industrially obtained strength
analyses.
The following conclusions can be drawn:
1. Significant differences were observed in the compressive strength at room temperature. A large statistically
significant drop in ultimate strength (roughly 30%) was
observed for all three grades when tested at 1600 8C.
Coke 1 suffered a further statistically significant fall
in strength at 1750 8C. Despite differences in the hot
strength properties between the coke grades, the order
of strength (2 > 1 > 3) remained the same at high
temperatures.
2. The standard deviation of strength was significantly
decreased at high temperatures indicating homogenization of coke strength as a result of heating.
3. Differences in contents of inert material and reactive
macerals between the studied coke grades proved to be
small. Likewise, the graphitization behavior of all three
grades as a result of heat treatment proved to be similar
at the test temperatures. Therefore, the difference in
strength was likely governed by other factors such as
pore structure, flaws, or the adhesive strength of the
carbon matrix. No direct correlation was obtained
between coke strength and pore shape factors of pores
below 1 mm in diameter. It is thought by the authors
that the effects of pore shapes on coke strength are
caused by flaws, connected pores, and pores of larger
size than the ones measured during this study.
4. Preceding heat treatment of coke at 1750 8C caused a
large decrease in the mean strength of coke 2 at room
temperature. This may be explained by weight loss and
increase of porosity during heat treatment or by the
increased graphitization degree of coke.
5. The stressstrain curves were similar in shape at room
temperature for all three cokes; early linear elastic
phase followed by fragile fracture. In most cases, little

1618

steel research int. 85 (2014) No. 12

force was required after the initial fracture, although on


some occasions significant force had to be applied all
the way up to the 33% deformation, which may indicate
layered crushing.
6. At temperatures of 1600 and 1750 8C, all three coke
grades showed unique distributions of stressstrain
curve types. Although the stressstrain curves at high
temperatures were complex, certain general types were
found for each grade. The shape of the stressstrain
curves at high temperatures also indicated plastic
deformation.

Acknowledgments
This research is a part of the Energy Efficiency & Lifecycle
Efficient Metal Processes (ELEMET) research program
coordinated by the Finnish Metals and Engineering
Competence Cluster (FIMECC). The Finnish Funding
Agency for Technology and Innovation (TEKES) is
acknowledged for funding this work. The authors would
also like to thank Mr. Tommi Kokkonen for assisting with
sample preparation and experimental work.
Received: December 4, 2013;
Published online: April 28, 2014
Keywords: blast furnace; coke; high temperature;
strength; Gleeble

References
[1] V. Vasilev, Coke Chem. 2012, 55, 423.
[2] H. Haraguchi, T. Nishi, Y. Miura, M. Ushikubo,
T. Noda, Tetsu to Hagane 1985, 25, 190.
[3] J. Patrick, H. Wilkinson, Proc. of the 42nd Ironmaking
Conf., Atlanta, GA, USA 1983, p. 333.
[4] Y. Okuyama, T. Isoo, K. Matsubara, Fuel 1985, 64,
475.
[5] M. Grant, A. Chaklader, J. Price, Fuel 1991, 70, 181.
[6] M. Holowaty, C. Squarcy, JOM 1957, 4, 577.

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.steel-research.de

2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

[15] H. Sato, J. W. Patrick, A. Walker, Fuel 1998, 77, 1203.


[16] R. Franklin, Proc. R. Soc. Lond. 1951, 209, 196.
[17] V. Sahajwalla, M. Dubikova, R. Khanna, 10th Int.
Ferroalloys Congress, Cape Town, South Africa,
February 14, 2004, p. 351.
[18] S. Gupta, V. Sahajwalla, J. Burgo, P. Chaubal,
T. Youmans, Metall. Mater. Trans. B 2005, 36B, 385.
[19] Y. Kashiwaya, K. Ishii, ISIJ Int. 1991, 31, 440.
[20] N. Andriopoulos, C. Loo, R. Dukino, S. McGuire, ISIJ
Int. 2003, 43, 1528.
[21] Y. Kubota, S. Nomura, T. Arima, K. Kato, ISIJ Int.
2011, 51, 1800.
[22] Y. Kubota, S. Nomura, T. Arima, K. Kato, ISIJ Int.
2009, 48, 563.
[23] J. Patrick, A. Walker, J. Mater. Sci. 1987, 22, 3589.
[24] O. Mattila, P. Salmi, 3rd Int. Conf. on Process
Development in Iron and Steelmaking, Lulea, Sweden,
June 811, 2008, p. 237.

steel research int. 85 (2014) No. 12

1619

FULL PAPER

[7] J. Haapakangas, J. Uusitalo, O. Mattila, T. Kokkonen,


T. Fabritius, 4th Int. Conf. on Process Development in
Iron and Steelmaking, Lulea, Sweden, June 1013,
2012, p. 177.
[8] J. Haapakangas, J. Uusitalo, O. Mattila, T. Kokkonen,
D. Porter, T. Fabritius, Steel Res. Int. 2013, 84, 65.
[9] S. Kim, Y. Sasaki, ISIJ Int. 2010, 50, 813.
[10] S. Gornostayev, J. Harkki, Energy Fuel 2006, 20,
2632.
[11] N. Amanat, N. Tsafnat, B. Loo, A. Jones, Scr. Mater.
2009, 60, 92.
[12] M. Lundgren, L. Sundqvist, B. Bjorkman, Steel Res.
Int. 2009, 80, 396.
[13] O. Mattila, Coke Stressing Equipment, Blast Furnace
Seminar, University of Oulu, Finland, March 3031,
2004.
[14] S. Gornostayev, O. Kerkkonen, J. Harkki, Steel Res. Int.
2009, 80, 390.

Das könnte Ihnen auch gefallen