Physical and Chemical Properties of Metallurgical Coke and Its Evolution in

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Fuel 366 (2024) 131277

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Physical and chemical properties of metallurgical coke and its evolution in


the blast furnace ironmaking process
Wenquan Niu , Yan Li , Qiang Li , Jingsong Wang *, Guang Wang , Haibin Zuo , Xuefeng She ,
Qingguo Xue
State Key Laboratory of Advanced Metallurgy, University of Science and Technology Beijing, Beijing 100083, China

A R T I C L E I N F O A B S T R A C T

Keywords: Pursuing green, low-carbon ironmaking technology primarily aims to reduce fuel ratios, specifically coke ratios.
Metallurgical coke Concurrently, as coke ratios decrease, coke deterioration becomes more pronounced; this significantly affects the
Blast furnace ironmaking smelting process and increase the demand for high-quality coke. Coke plays an irreplaceable role in blast fur­
Coke properties
naces (BFs), acting as both the skeleton of the stock column and the permeable layer of the gas flow. Thus,
Multi-phase reactions
Alkali metals
examining the structural characteristics of coke and evolutionary behavior of its properties within the BF is
Graphitization crucial. To thoroughly comprehend the properties of coke and its evolution in BFs, this article divides the
structure and composition of coke into four aspects: pore, matrix, and microcrystalline structures and mineral
compositions. First, the behavior of coke in the BF is addressed, examining the evolution of its performance at
different positions and its involvement in multiphase reactions. Subsequently, the effects of alkali metals and zinc
accumulation in the furnace on the properties of coke are summarized in detail. Finally, the graphitization
phenomenon of coke within the furnace is introduced. As the temperature increases, the degree of graphitization
of coke deepens; this has a significant effect on its structure and composition. This review also identifies the
knowledge gaps and outlines potential future research opportunities in this domain.

replacement of coke with alternative reducing agents remains unfeasible


due to its irreplaceable role as both a structural component and a
1. Introduction permeable layer for gas flow [14]. In pursuit of the “double carbon”
objective, it is imperative to develop an innovative, intensive BF iron­
Blast furnace (BF) ironmaking continues to dominate the ironmaking making model to minimize coke consumption [15,16].
industry, accounting for approximately 70 % of the global share [1]. Accurately characterizing the transformation of coke’s macroscopic
Steel, a cornerstone of modern civilization, is extensively utilized in properties at the microscopic scale holds both theoretical and practical
sectors like transportation, construction, and infrastructure. As one of importance for precisely controlling coke quality. Beginning with an in-
the most carbon- and energy-intensive sectors, the steel industry faces a depth initial characterization of coke’s structure and composition, this
direct correlation between increased steel production and increased study meticulously explores the qualitative and quantitative relation­
greenhouse gas (GHG) emissions [2–5]. Notably, emissions from the ships between coke’s structure, composition, and properties. We sum­
steel sector constitute 7 % of total global emissions and 16 % of indus­ marize the mechanisms underlying coke quality deterioration in various
trial GHG emissions [6,7], placing a significant responsibility on this regions of the BF, along with the influencing factors. This is done to offer
industry regarding resource management, energy efficiency, and insights and perspectives to enhance coke quality and reduce its
pollution emission reduction. consumption.
The substitution of charcoal with coke in BF ironmaking marked a
revolutionary shift in the history of metallurgy. Coke, one of the most 2. Structure and composition of coke
costly raw materials [8], coke plays, fulfills four primary roles in BF: as a
reducing agent for iron ore, a heat source, a carburizing agent for hot When heated to (950 to 1050) ◦ C under air-isolated conditions,
melt, and as the skeleton of the stock column [9–13]. Despite centuries bituminous coal undergoes stages of drying, pyrolysis, melting, bonding,
of advancements in BF ironmaking technology, the complete

* Corresponding author.
E-mail address: wangjingsong@ustb.edu.cn (J. Wang).

https://doi.org/10.1016/j.fuel.2024.131277
Received 19 December 2023; Received in revised form 30 January 2024; Accepted 22 February 2024
Available online 4 March 2024
0016-2361/© 2024 Elsevier Ltd. All rights reserved.
W. Niu et al. Fuel 366 (2024) 131277

Nomenclature GHG Greenhouse gas


BET Brunauer-Emmett-Teller
BF Blast furnace DFT Density functional theory
BJH Barret-Joyner-Halenda OTI Optical anisotropy index
DH Dollimore-Heal CRS Coke strength after reaction
CRI Coke reactivity index La Average extension width
Lc Average stacking height d002 Interlayer spacing
g Graphitization degree TEM Transmission electron microscope
XRD X-ray diffraction HRTEM High-resolution TEM
HR High resolution XPS X-ray photoelectron spectroscopy
LTA Low-temperature ashing FTIR Fourier transform infrared
M10 Micum index 10 M40 Micum index 40
RAFT Raceway adiabatic flame temperature MCI Mineral catalytic index
MBI Mineral basicity index HTT Heat treatment temperature
DI Drum intensity PC Pulverized coal
NG Natural gas

curing, and shrinkage, and ultimately forms coke. Coke can be charac­ 500) ◦ C, where volatile components are actively decomposed, resulting
terized as a nonhomogeneous composite material comprising organic in a general porosity of 50 % to 60 % [26]. This structure ranges from
carbon, inorganic minerals, and numerous pores [17]. Coke burns with a macropores, with diameters exceeding several hundred micrometers, to
short blue flame and without smoke, releasing substantial heat; the ultra-micropores, with diameters less than 1 nm, dispersed throughout
combustion is slow and enduring, often forming an ash shell on its the coke. The pore structure serves as a focal point for stress concen­
surface. As showed in Fig. 1, macroscopically, coke is a hard, porous, tration, potentially impairing coke strength, and acts as a conduit for
cracked material, silvery-white or grayish black with a metallic luster. reaction gases to permeate the coke. Consequently, this structure
Microscopically, it comprises pores, microscopic cracks, and a solid significantly influences the coke’s mechanical strength and thermal
carbon matrix embedded with inorganic minerals and organic inert properties [27]. Hence, the pore structure is a critical indicator of its
materials [18]. mechanical strength and thermal characteristics. The primary tech­
Statistically, over 90 % of global coke production is utilized in BFs, niques for assessing coke’s pore structure encompass vacuum drainage,
rendering it a fundamental component of the steel industry. In the past microscopic image analysis, pressed mercury assays, and gas adsorption
six decades, the structure of coke has been extensively characterized and methods.
documented [19–22]. Recent fundamental research reveals that coke’s Different research tools are essential for examining the various sizes
performance primarily depends on its structure, encompassing aspects of coke pores. The mercury porosimeter can detect pore sizes ranging
like the ratio of isotropic carbon to inert matter, the size and shape of from (0.003 to 500) μm. Pores larger than 500 μm are measurable
anisotropic carbon units, the intersection of various structural organi­ through microscopic image analysis, while those smaller than 0.003 μm
zations, and the porosity and ash composition [23–25]. The forthcoming are detectable via the gas physical adsorption method [28–31]. Fig. 2
sections offer detailed insights into coke’s pore structure, matrix struc­ illustrates the capacity to characterize coke pores from the nanometer to
ture, microcrystalline structure, and mineral composition, thereby millimeter scale using this diverse detection [32,33].
enhancing the understanding of coke and contributing to producing The vacuum drainage method utilizes Archimedes’ principle to
higher-quality coke for BF operations. determine coke porosity. This is calculated based on the weight of the
coke before and after water absorption, a technique formalized as a
national standard (GB/T 4511.1–2008) [27]. Wu et al. [34] employed
2.1. Pore structure this method to measure coke’s apparent porosity. Their unidirectional
test revealed that the accuracy of the measurements is influenced by
The porous structure of coke primarily develops during the high- several process parameters, including the vacuum degree, resting time,
temperature dry distillation phase of coking coal, occurring at (350 to

Fig. 1. Samples of coke: (a)macroscopic sample; (b)microscopic sample.

2
W. Niu et al. Fuel 366 (2024) 131277

Fig. 2. Techniques for recognizing pores in solid materials [32].

and suction time, as depicted in Fig. 3. assessing pore size distribution. Using various theoretical models, this
The microscopic image analysis method employs an optical micro­ technique computes the specific surface area, pore volume, and pore size
scope to examine the polished surface of coke. This technique involves distributions. The Brunauer-Emmett-Teller (BET) model, for instance, is
setting a specific grayscale threshold for segmentation and identifica­ employed to ascertain the adsorption amount of a single molecular layer
tion, distinguishing between the pores and their walls. A computer then on the specimen, thereby facilitating the calculation of the specimen’s
statistically determines the pore structure parameters, such as pore specific surface area. Further, pore size distribution is derived using
diameter, wall thickness, and distribution across the entire measure­ Barret-Joyner-Halenda (BJH), Density functional theory (DFT), and
ment areacomputer [27]. Nyathi et al.[35] utilized optical microscopy to Dollimore-Heal (DH) models. In the gas-adsorption method applied to
analyze the pore structure of coke with varying grain sizes. An image coke’s pore structure analysis, N2 and CO2 are predominantly used. CO2,
processing program was employed to obtain parameters like pore area, under laboratory conditions, exhibits superior diffusion capability
wall thickness, and diameter. Fig. 4 showcases an example of how these compared to N2. This attribute enables CO2 to readily attain adsorption
pore parameters are determined. equilibrium and effectively penetrate slit-like micropores smaller than
In Japan, 3D imaging of coke’s internal structure was achieved using 0.7 nm [40]. Grigore [41] measured the specific surface area of nine
X-ray transmission scanning. This method allows for reconstructing coke types using N2 and CO2 as adsorbents. The study revealed that the
arbitrary tomographic images, thereby completing the 3D internal [36] specific surface area determined through CO2 adsorption was markedly
coke structure. The internal structure data obtained through comput­ greater than that measured via N2 adsorption. A detailed comparison of
erized tomography was then used to model coke, enhancing the simu­ these methods, including the sizes of pores they can detect and their
lation of the model [37]. However, the effectiveness of this approach respective advantages and disadvantages, is provided in Table 1.
depends on the scanning equipment and image processing software,
necessitating further research for its widespread application. 2.2. Matrix structure
The pressed mercury assay operates on the principle of mercury’s
surface tension, assuming that the internal pores of porous materials are The coke matrix, or pore wall, constitutes the primary component of
cylindrical of varying sizes and extend to the sample’s outer surface. coke, with its structure critically influencing coke’s most essential
This method analyzes the distribution of pores in porous materials by properties. This solid portion of coal remains after parts of the coke oven
calculating the amount of mercury pressed at different pressures. The gas and coal tar are released during dry distillation. During the coking
pore distribution in coke is determined based on the varying pressures process, the active components of coal such as vitrinite and lipidome
applied during mercury pressing [38]. Yamamoto et al.[39] explored the soften and melt, forming a continuous pore wall. In contrast, inert
relationship between coke pore configuration and tensile strength using components like inertinite, opaque, and semi-vitrinite are retained in
mercuric pressure. They found that an increase in stomata below 100 μm the pore wall alongside the softened active components [42]. The optical
helped maintain the tensile strength of coke at consistent reaction rates. organization composition of coke can be ascertained from the varied
The gas adsorption method leverages the phenomenon of capillary morphological characteristics of the coke matrix structure, observable
condensation and the principle of volume equivalent exchange for through polarized light microscopy [43]. The optical organization of
coke is investigated by isolating the matrix portion remaining in the pore
structure, classifying it based on microscopic characteristics (shape, size,
color) using polarized light microscopy.
International optical organization classification schemes, including
those by Patrick, Sugimura Hidehiko, and Marsh, are widely recognized.
Reflecting the characteristics of China’s coking coal and metallurgical
coke, Chinese researchers revised the national metallurgical industry
standard YB/T077–1995, “Determination Methods of Optical Tissue of
Coke,” in 1994. Coin et al. [44] identified typical visual textures of coke,
such as isotropy, mosaic structure, schistose, fibrous, banding, anisot­
ropy, and pyrolytic carbon. To streamline matrix structure character­
Fig. 3. Orthogonal test results: (a)measurement results of apparent porosity of ization, the optical organization of coke was simplified to classifications
coke under different vacuum levels; (b)coke water absorption as a function of of isotropic or anisotropic. All optical textures are considered aniso­
time [34]. tropic except for isotropic and inertinite [32]. The degree of optical

3
W. Niu et al. Fuel 366 (2024) 131277

Fig. 4. An example of the determination of pore parameters in a coke matrix: (a)original image; (b)measurement procedure [35].

Table 1
Comparison of coke pore structure characterization methods.
Method Sample size Pore size Parameters Type Advantages Disadvantages

Vacuum drainage 15–60 mm >2.2um Porosity Open pores Convenient and fast Parameter less
Microscopic image 30 × 30 mm >100 nm Porosity, shape, size Open and close High resolution Broken samples, plane
pores representation issues
3D scanning Cylindricity >500um Porosity, shape, size Open and close 3D structure without breaking Complex process, imperfect
pores the sample methodology
Pressed mercury 0.02 mm 0.3um- Size distribution Open pores Fast measuring speed Possible collapse
assay 400um
Gas adsorption 0.02 mm 0.4 nm-100 Size distribution, fractal Open pores Avoiding interference from Small detection range
nm dimension chemisorption

anisotropy varies among different optical tissues. As coal deterioration


intensifies, the isotropic organization in the resultant coke diminishes,
the anisotropic organization and size incrementally increase, and the
microcrystalline large-particle structure becomes more prevalent.
The various optical organization structures in coke are assigned
distinct optical anisotropy values. The coke optical anisotropy index
(OTI) is derived by weighing the coke according to its content. This
index reflects the degree of anisotropy in coke, indicating the level of
order of the carbon atoms within it. Zhang et al. [45] evaluated the
optical structure of coke from six different manufacturers, adhering to
the Chinese standard YB/T 077–2017 (determination of the optical or­
ganization of coke). The study assessed the matrix structure of coke from
these manufacturers, as delineated in Table 2. It was observed that coke
with a higher content of isotropic structure exhibited increased reac­
tivity. All isotropic textures appear black due to extinction, whereas
anisotropic weaves exhibit brightness as the reflected light partially
penetrates the bias detector. The results of this study, illustrated in
Fig. 5, indicate that coke with a higher degree of anisotropy corre­
sponded to a lower Coke Reactivity Index (CRI).
In conclusion, various scholars have employed many methods to
investigate the coke matrix structure, significantly enhancing our un­ Fig. 5. Content of eight coke anisotropic structure types [25].
derstanding of coke’s microstructure. These studies have established a
qualitative and quantitative relationship between the structure of the multiple dimensions.
coke matrix and its macro properties. Nonetheless, given the complexity
of the coke matrix structure, continued development of diverse meth­
odologies is essential to comprehend the structure of the coke matrix in 2.3. Microcrystalline structure

During the high-temperature pyrolysis of coking coal, an increase in


Table 2
temperature leads to the continuous shedding and decomposition of side
Optical tissue content and CRI values [45].
∑ ∑ chains on aromatic rings. This process results in the condensation and
Coke ISO a/% CRI/% Coke ISO/% CRI/% thickening of these aromatic rings, forming a microcrystalline structure.
C1 28.8 18.4 C4 47.4 34.4 This structure is the most fundamental structural unit of the coke matrix
C2 31.0 21.3 C5 45.3 36.5 (stoma wall) and is akin to the structure of graphite crystals. The
C3 23.0 20.6 C6 79.9 46.2
investigation of the microcrystalline structure of coke often involves
a
Σ ISO=%Isotropic+%Inert. comparisons with graphite crystal structures, as schematically

4
W. Niu et al. Fuel 366 (2024) 131277

illustrated in Fig. 6. Carbon, the predominant component of the coke bonds, C = O bonds, and COO bonds. It is also used for fitting various
matrix constitutes over 80 % of coke’s mass and exists in various forms ratios and determining the graphitized carbon content, which serves as a
microcrystalline [11]. The carbon structure within coke is primarily parameter for assessing the degree of graphitization [48]. Raman spec­
discerned through its microcrystalline structure [46]. troscopy, with its focus on G (graphite) and D (defect) bands, is typically
Among the different dimensions of coke structure, the microcrys­ found near (1580 to 1600) cm− 1 and 1350 cm− 1, respectively [49]. For
talline structure of the matrix holds significant importance. It is instru­ detailed information on the carbon structure of coke, the Raman spectra
mental in guiding coke production, understanding coke performance, are curve-fitted in the range of (800 to 1800) cm− 1, as shown in Fig. 8 for
and promoting efficient utilization in the BF. Therefore, obtaining the coal samples post-carbonization at 700 ◦ C [50]. FTIR analysis primarily
parameters of the microcrystalline structure through characterization involves identifying and attributing absorption peaks [33], which helps
methods and establishing its correlation with coke performance be­ to discern the various functional groups and their relative contents
comes a crucial indicator for evaluating coke. within the coke structure.
Utilizing Scherrer’s formula and Bragg’s equation enables the The parameters of coke’s microcrystalline structure indicate the size
computation of various microcrystalline parameters. These calculations and ordering degree of coke microcrystals and their influence on coke
facilitate the determination of vital microcrystalline attributes, reactivity. Researchers have garnered significant insights into its struc­
including the average stacking height (Lc), average extension width (La), ture through qualitative and quantitative studies of coke’s microcrys­
the graphitization degree (g), and interlayer spacing (d002) as follows: talline structure using these characterization methods. This knowledge
is vital for understanding the nuances of coke’s microcrystalline struc­

L= (1) ture, guiding coke production processes, and optimizing BF ironmaking
βcosθ
techniques.
λ
d002 = (2) 2.4. Mineral composition
2sinθ002

0.3340 − d002 Coke minerals originate from the ash composition of coking coal and
g= (3)
0.3340 − 0.3354 exist primarily as minerals expressed in oxide forms. The type and
content of these minerals not only influence the dissolution loss reaction
where K represents the constant specific to the given reflection plane of coke, which subsequently affects its thermal properties but also, as
(L002: A = 0.91; L10: A = 1.84); λ is the wavelength of the X-rays, nm; β is domestic and international studies have indicated, the distribution
the half-peak full width; θ signifies the Bragg angle; and 0.3440 nm is the mode of these minerals [51]. Although the mineral content in coke is
layer spacing for completely non-graphitized carbon. The g value of the typically less than 15 % by mass, its distribution within coke can be
coke is determined by comparing the relative difference between the categorized into three primary forms: (a) typical discrete distribution,
calculated d002 value and the layer spacing of ideal graphite, which is (b) decentralized distribution, and (c) stomatal inclusions, as illustrated
0.3354 nm. in Fig. 9. Despite their small proportion, these minerals significantly
The microcrystalline structure of coke can be characterized by impact the interaction between coke and the gas–solid–liquid phases in
several methods, including X-ray diffraction (XRD), transmission elec­ the BF. Their roles in the coke dissolution loss reaction can be catego­
tron microscopy (TEM), high-resolution TEM (HRTEM), X-ray photo­ rized as increasing the dissolution loss reactivity (positive catalysis),
electron spectroscopy (XPS), and Raman and Fourier transform infrared decreasing it (negative catalysis), or having no effect.
(FTIR) spectroscopies. XRD and TEM are the most commonly used for Alkaline elements such as Fe, Ca, and Mg, and alkali metals like K
studying coke’s microcrystalline structure. The geometry of ideal mi­ and Na are highly reactive at elevated temperatures, affecting coke’s
crocrystals is depicted in Fig. 7. XRD allows for the quantitative analysis structure and catalyzing its dissolution reaction [52]. Boron, conversely,
of layer spacing, stacking height, and lamella size in coke microcrystals. exerts a negative catalytic effect on coke solvation and is utilized for
Images of coke microcrystals obtained through TEM and HRTEM un­ coke passivation. Spraying boron modifiers onto coke has been shown to
dergo Fourier transformation, noise reduction, and mathematical and reduce coke reactivity and enhance its strength post-reaction [53]; the
statistical computation to derive parameters characterizing their struc­ addition of TiO2 decreases the proportion of active carbon sites in coke,
ture. These parameters include microcrystalline stripe length, curvature, leading to reduced reactivity during gasification and increased strength
and inclination angle, collectively providing insights into the size, post-gasification [54]. Aluminosilicates in coke, being stable at high
ordering degree of coke microcrystals, and their impact on coke reac­ temperatures, exhibit no catalytic effect [11]. Certain metals and alkali
tivity. XPS technology is instrumental in characterizing the bonding metals are recognized as effective catalysts for coke gasification,
states of carbon and oxygen in coke, including graphitized carbon, C-O particularly at temperatures below 900 ◦ C, accelerating the reaction

Fig. 6. Schematic crystal structure of graphite [47].

5
W. Niu et al. Fuel 366 (2024) 131277

Fig. 7. Geometry of ideal microcrystals: (a)top view; (b)side view [32].

effectiveness was K > Na > Ca > Fe > Mg.


Researchers [57,58] have introduced the concept of the coke reac­
tivity catalytic index, quantifying the role of minerals in catalyzing the
coke dissolution reaction. This index serves as a predictive measure for
coke reactivity. Presently, domestic and international research recog­
nizes the MBI and MCI as catalytic indices. These two indices exhibit a
strong correlation and can be interchangeably applied under specific
conditions. Furthermore, Sakurovs et al. [59] employed the low-
temperature ashing (LTA) method [60,61] to quantitatively charac­
terize the mineral structure in commercial coke. This approach revealed
that the minerals in coke predominantly consist of amorphous alumi­
nosilicates and other amorphous minerals formed due to the removal of
bound water or volatile components.

3. Evolution of coke in the BF

3.1. Evolution of coke properties

Fig. 8. Fitted Raman spectra of carbonized samples [50]. Coke properties typically encompass particle size, strength, and
chemical composition. The primary indicators of coke quality are par­
rate. However, the presence of SiO2 and Al2O3 in coke reduces its ticle size, mechanical strength (M40, M10), thermal strength (CRI, CSR)
reactivity[55]. Huang et al.[56] investigated conducted research on the and composition. Performance indices for coke vary depending on the
reactivity of alkali, alkaline earth, and transition metals in semi-coke size of the BF, and the ’Design Code for BF Ironmaking Processes’ stip­
gasification, finding that the order of metals in terms of catalytic ulates these indices for large BFs, as detailed in Table 3.

Fig. 9. Forms of mineral distribution within coke: (a)typical discrete distribution; (b)decentralized distribution; (c)stomatal inclusions[17].

6
W. Niu et al. Fuel 366 (2024) 131277

BFs generally require a coke particle size range of (40 to 50) mm. The quicker charge renewal and, consequently, shorter residence times for
coke particle size must not be significantly larger than the ore size, coke in the lumpy zone. In this zone, coke and ore display a layered
aiming for close similarity between the two. Huge average particle sizes alternating distribution. The charge remains granular, the coke is not
in coke can lead to easy breakage, excessive powder production, softened and melted, and there is no adhesion [64,65].
impaired gas and liquid permeability, and an elevated differential Owing to the lower temperatures, which do not reach the threshold
pressure in the BF. This may necessitate a reduced air volume, leading to for the dissolution reaction, in the lumpy zone, coke undergoes physical
decreased output and fuel ratio [45]. According to contemporary BF changes. It is subjected to friction and compression from the charge,
refining technology standards, the coke powder content less than 5 mm leading to a reduction in coke particle size and the production of pow­
should be below 5 %, and the proportion of (5 to 15) mm coke should be der. In this zone, coke experiences extrusion, mutual collision, and
under 30 % [62]. abrasion; however, the bulk layer’s static pressure is significantly lower
The mechanical strength of coke is a vital metric for assessing its than coke’s compressive strength. As a result, the forces of impact and
ability to support the BF’s framework and ensure smooth operation. abrasion are also diminished. Consequently, there are no significant
Coke’s thermal strength reflects its capacity to resist chemical erosion changes in coke strength reduction, lump size reduction, chemical
and protect the charge skeleton in the BF, serving as a critical index for composition, and deterioration of material column permeability.
evaluating its thermal stability. M10 is commonly used to indicate
coke’s wear resistance, while M40 denotes its fracture resistance [63]; 3.1.2. Cohesive zone
CRI and CSR are employed to characterize coke’s high-temperature re­ The cohesive zone, situated in the shaft and bosh of the BF, has a
action properties, with numerous studies indicating a negative correla­ residence time influenced by the type of BF, productivity, and the soft
tion between them [43]. The chemical composition includes ash, sulfur, melting characteristics of the charge. A higher softening temperature of
volatiles, and phosphorus. Increased ash content can deteriorate the the charge broadens the cohesive zone and extends the residence time of
internal structure of coke, leading to more cracks and facilitating CO2 the coke [66]. The charge begins to soften and deform within this zone,
diffusion into its interior, thus exacerbating thermal performance creating a soft-melting material that bonds the coke particles together.
degradation. High sulfur content can raise the sulfur level in pig iron, This close combination increases friction and extrusion among particles,
reducing its quality or increasing slag alkalinity, adversely affecting BF enhancing both chemical action and mechanical damage. Consequently,
operation indices. Elevated volatile content suggests under-processed the particle size of the coke significantly decreases, and its abrasion
coke. High phosphorus levels in coke can increase the brittleness of resistance is reduced.
pig iron when cold. Recent research indicates that coke deterioration is exacerbated in
The evolution of coke within the BF is a gradual and continuous large-scale BFs and with oxygen-rich blowing. This leads to an increased
process. In the upper part of the furnace, changes are observed in the coke load, heightened coke gasification reactions, destruction of the
coke’s particle size, which decreases; its strength, which worsens; its coke surface structure, and reduced coke particle size [67]; Hydrogen-
reactivity, which increases; and its chemical composition, which alters. rich smelting in the BF weakens the assimilation of the burden and de­
Dissecting BFs has proven to be the most effective method for studying creases the erosion of the coke during smelting reduction. This process
material properties inside the furnace, and this approach has been results in the downward movement and narrowing of the cohesive zone,
employed in dozens of BFs worldwide. Notably, in the 1970 s, three effectively improving the BF’s permeability [68–70].
Japanese BFs underwent dissection. Based on these studies, researchers As the charge descends, the coke endures not only extrusion and
analyzed changes in coke properties along the height inside the BF and friction but also high-temperature thermal stress. This leads to a marked
examined the effects of variations in coke properties and BF operational increase in porosity, thinning of pore walls, and a reduction in the
methods on furnace conditions, as depicted in Fig. 10. In the upper strength of the coke matrix, resulting in chalking that further reduces
section of the BF, where temperatures remain below 1000 ◦ C, the mean coke particle size and impairs the BF’s gas permeability. Coke undergoes
size, drum intensity, reactivity, and CSR of the coke do not exhibit sig­ a vigorous gasification reaction with CO2 produced by ore reduction.
nificant changes. However, in the middle part of the BF, at temperatures This reaction starts at the coke’s surface, causing the surface pores to
ranging from (1000 to 1400) ◦ C, the surface of the coke begins to enlarge and weakening the coke’s strength. The gasification reactivity of
deteriorate, and all indices start to degrade gradually. Above the tuyere coke in the cohesive zone is heightened due to the adhesion of iron-
level, where temperatures increase progressively, the deterioration of containing dust and the chemical composition of zinc and alkali
coke properties becomes markedly pronounced. metals compared to the incoming coke [28]. Alkali cycling in the BF
increases the potassium and sodium mass fraction, which catalytically
3.1.1. Lumpy zone accelerates coke gasification. Post-reaction, the strength of the coke
The lumpy zone between the charge level and the cohesive zone is diminishes sharply, leading to pulverization. This deterioration in coke
characterized by temperatures below 1000 ◦ C. The residence time of integrity compromises the BF’s permeability and seriously threatens its
coke and iron-containing charges in this area is contingent on the smooth operation [71].
furnace capacity and the utilization factor. Larger capacities entail
longer residence times, whereas the higher utilization factors lead to 3.1.3. Dripping zone
The dropping zone, situated below the cohesive zone, is where the
fully melted liquid slag and iron pass through the solid coke layer in
Table 3 droplet form before entering the furnace hearth. In this zone, the coke
Quality regulations of coke with different volumes in code for design of the BF
undergoes scouring by iron and slag and is impacted by high-
ironmaking process.
temperature furnace gases. This interaction leads to the evaporation of
Furnace melting level/m3 1000 2000 3000 4000 5000 some ash components in the coke, further increasing its porosity and
M40/% ≥ 78 ≥ 82 ≥ 84 ≥ 85 ≥ 86 continuing the decline in strength. However, it remains solid due to
M10/% ≤ 7.5 ≤ 7.0 ≤ 6.5 ≤ 6.0 ≤ 6.0 coke’s high melting point (3000 ◦ C) and its unburned state. The evolu­
CSR/% ≥ 58 ≥ 60 ≥ 62 ≥ 64 ≥ 65
tion of coke properties in the cohesive zone has been examined through
CRI/% ≤ 28 ≤ 26 ≤ 25 ≤ 25 ≤ 25
Ash/% ≤ 13 ≤ 13 ≤ 12.5 ≤ 12 ≤ 12
core drilling and sieving. Analyses of coke samples collected from the
St,d /% ≤ 0.85 ≤ 0.85 ≤ 0.7 ≤ 0.6 ≤ 0.6 tuyere reveal that the gasification reaction predominantly occurs on the
Size range/mm 25–75 25–75 25–75 25–75 30–75 surface of the coke. Internally, the coke remains largely undamaged by
Size overtakes upper limit ≤ 10 ≤ 10 ≤ 10 ≤ 10 ≤ 10 the chemical reactions inside the BF, especially when compared to the
Size overtakes lower limit ≤8 ≤8 ≤8 ≤8 ≤8
reducing effects of minerals in the ash.

7
W. Niu et al. Fuel 366 (2024) 131277

Fig. 10. Variation of coke properties in the direction of height in the BF [64].

In the dripping zone, the iron melt undergoes carburization, and the
coke experiences dissolution and particle size reduction. The carburi­
zation mechanism of the iron melt is influenced by various factors,
including hydrogen-rich conditions, which are known to enhance the
permeability of the layers [70]. In the hydrogen-rich upper part and
belly of the BF, there is a significant decrease in temperature, and the
high-temperature zone becomes narrower. Consequently, the direct
reduction of the iron-bearing burden and the solution loss of coke are
reduced, favoring an indirect reduction process [72].

3.1.4. Raceway and deadman


The hot blast emitted from each tuyere in the BF creates a raceway in
front of it, where coke particles and blowing materials interact at high
speed convolutedly. Around this raceway, coke is rapidly consumed due
to the gasification reaction, moving slowly towards the center of the BF Fig. 11. Gas temperature distribution in the center section of the raceway: NG
to form a “deadman.” In the BF raceway, the exothermic combustion and PC injection (top); only PC (bottom) [74].
reactions of coke and blowing fuel cause the raceway adiabatic flame
temperature to surge rapidly, exceeding 2273 K. When employing a
This temperature significantly exceeds the coking temperature. As a
process of blowing pulverized coal in the BF, the temperature at the
result, the ash in the coke begins to decompose. Sulfides and oxides
flame’s center in the tuyere’s raceway can rise substantially, reaching
within the ash react with the coke’s carbon. Concurrently, volatiles
approximately 2373 K [73]. Fig. 11 presents the temperature variation
precipitate from the pore walls under high-temperature conditions,
in the raceway during the numerical simulation of scenarios where
leading to a further weight loss of approximately 4 % to 5 %. This
natural gas and pulverized coal (PC) are mixed, as well as when only
process alters the lattice height and length of the carbonaceous matter,
pulverized coal is blown.
reducing the coke’s reactivity. The coke experiences substantial thermal

8
W. Niu et al. Fuel 366 (2024) 131277

stresses due to the temperature gradient between its surface and center crystals. As the sampling position progresses, the slag and iron contents
under high heat, forming numerous microcracks [75]. in the samples initially increase before stabilizing at a relatively constant
CO2 molecules and slag iron penetrate the coke through surface level [85,86]. Similarly, the particle size of coke diminishes, then levels
pores, eroding its surface layer. Subsequently, the coke’s outer layer off as the location moves closer to the center. The alkali content
undergoes several processes, including slag iron scouring, carburization, significantly increases with greater distance from the tuyere entrance.
and combustion. This leads to the layer-by-layer powdering and peeling This suggests that alkali vapors predominantly circulate and accumulate
off the outer layer, progressing the reaction towards the coke’s core. in the central part of the BF, with the high-temperature conditions at the
Consequently, the coke’s size gradually reduces until it completely tuyere being a primary factor in coke graphitization [46,87].
disappears. Meng et al. [88] used the drilling method to collect four sets of
In the raceway, coke gasification creates a deteriorated layer on the samples at different hearth heights, as depicted in Fig. 13. Their findings
coke surface. This layer is continuously shed during the rotating friction indicate an increase in coke size with greater distance from the sidewall.
process, generating a substantial amount of coke fines. These fines enter Conversely, the closer the proximity to the hearth’s bottom, the smaller
the birdnest and deadman, impairing the gas and liquid permeability of the coke size. Enriching S and K elements at the interface contributes to
the BF, thereby affecting the BF’s stability [76]. coke cracking and impedes the carburization reaction. Additionally, the
Experimental results [77–79] indicate that the deadman is situated presence of slag and alkali metals in the coke’s pores leads to the
in the central lower part of the BF, having a conical shape and a rela­ volumetric expansion of the coke.
tively dense structure. The edge of the deadman represents the coke-
consuming area, where the internal coke moves towards the edge to
renew its state. The carbon is gradually eroded by the surrounding iron 3.2. Multiphase reactions of coke
melt and eventually disappears, with the ash integrating into the slag.
3.2.1. Gas–solid reaction
3.1.5. Hearth The high-speed hot blast in front of the tuyere forms a cylindrical jet
In the lower central region of the BF, coke is less influenced by space comparable in diameter to the tuyere, known as the combustion
combustion and raceway activity, resulting in larger coke sizes in this zone. In this zone, coke is burned and gasified by the incoming hot blast,
area of the hearth. As the coke gradually descends to the furnace bottom, leading to gasification and dissolution reactions that produce reducing
its size decreases correspondingly—the closer to the bottom, the smaller gases such as CO and H2. The ascending hot gas stream transfers heat to
the coke size. Tuyere sampling and BF dissection studies have reported a the raw materials and engages in a reduction reaction with iron ore,
reduction of at least 50 % in the size of feed coke by the time it reaches generating a significant amount of CO2. This process results in a pro­
the tuyere level [80,81]. Studies at the tuyere level have revealed that nounced gasification reaction with coke [67]. Thus, in the ironmaking
little or no slag infiltrates the macroscopic pores of the coke, suggesting process, coke participates in gas–solid reactions, primarily the gasifi­
that the slag does not penetrate the internal structure of the coke. cation reaction, the indirect reduction reaction of iron oxides, and the
Consequently, there is no reaction between the coke ash and slag, combustion reaction.
leaving most of the coke intact at this level [28].
The extended contact duration between coke and slag at the slag (1) Gasification reaction
liquid level exceeds that at the tuyere level. This prolonged interaction
allows slag ample time to permeate the coke pores fully. Significant The gasification reaction of coke is a heterogeneous gas–solid reac­
reactions occur between the coke ash and slag, yet the coke’s structure tion, predominantly a chemical process occurring via the continuous
remains unchanged [82]. In contrast, at the iron melt level, the iron melt contact of CO2 produced by the indirect reduction of iron oxides, with
erodes the coke, dissolving it. The irregular contact surface between iron activation sites on the surface of solid coke. BF researchers often
and coke evidences a reaction attributed to the robust impact of the iron describe carbon consumption in terms of a carbon dissolution reaction
melt on the coke structure. This interaction ultimately leads to the [89]. Boudouard established the stoichiometry and reversibility of this
destruction and complete dissolution of coke [63]. Coke samples from reaction in 1899 [90]. The reaction equation is as follows:
various depths and heights are commonly extracted through drilling to
C + CO2 = 2COΔr Hmθ (298.15 K) = + 172.7 kJ⋅mol− 1 . (4)
analyze the evolution of coke properties in hearths. These morpholog­
ical changes are illustrated in Fig. 12. The study’s findings indicate that The experimental methods for investigating the behavior and
the coke particle size diminishes with increasing sampling depth. This mechanism of coke gasification reactions in BF encompass techniques
size reduction is inversely correlated with the degree of graphitization such as fixed beds, thermogravimetry, micro fluidized bed, and gas
and the adsorption amount of alkali metals [83,84]. analyzer combinations. The fixed-bed approach effectively simulates the
Coke graphitization facilitates the production of fine coke particles. interaction of reacting gases traversing through the coke layer and the
It has been observed that the finest coke particles exhibit the highest interplay between gas and solid phases within the BF. Notably, the CRI
degree of graphitization and contain the largest proportion of graphite and CSR tests employ fixed-bed reactors. The reactivity and kinetic
characterization of coke gasification are primarily conducted using
thermogravimetry and micro-fluidized beds, enabling precise analysis of
the impact of diverse experimental parameters on the coke gasification
reaction behavior[32].
Coke dissolution is a gas–solid reaction, depicted in Fig. 14, and
predominantly proceeds through the following stages [91]: ① Diffusion
of CO2 and CO through the stagnant boundary layer to the coke’s outer
surface; ② Transfer of CO2 to the surface of reacting particles via the
outer layer, coupled with the movement of the carbon surface desorp­
tion product, CO, from reactive sites to the particle surface through
pores; ③Adsorption of the reactant CO2 onto the active sites on the
carbon surface, ensuing in a chemical reaction to produce CO. This re­
action encompasses three distinct parts: chemical control, transition,
and diffusion control [92] zones. Transition temperatures for these
Fig. 12. Morphology of coke at different depths in core samples [83]. processes vary across different coke types, with the three stages not

9
W. Niu et al. Fuel 366 (2024) 131277

Fig. 13. Schematic diagram of the hearth drilling of BF sampling positions [88].

Table 4
Thermodynamic data for the indirect reduction reaction of iron oxides.
Chemical reaction ΔH(kJ/mol) ΔS(J/Kmol) ΔG(kJ/mol) T
(K)

Hydrogen reduction
3Fe2O3 þ H2 ¼ 2Fe3O4 þ –6 +89 –33 298
H2O
Fe3O4 þ H2 ¼ 3FeO þ H2O +61 +95 –22 873
FeO þ H2 ¼ Fe þ H2O +31 +24 +10 873
Carbon monoxide
reduction
3Fe2O3 þ CO ¼ 2Fe3O4 þ –47 +47 –61 298
CO2
Fe3O4 þ 4CO ¼ 3Fe þ +19 +53 –26 873
4CO2
FeO þ CO ¼ Fe þ CO2 –12 –18 +4 873
Fe3O4 þ 6CO ¼ Fe3C þ –161 –159 –114 298
5CO2
Fig. 14. Schematic diagram of the coke dissolution reaction process [91].

(3) Combustion reaction


being sharply demarcated. Studies indicate that at temperatures below
900 ℃, the reaction is predominantly governed by chemical processes. In the combustion zone, located in front of the tuyere, carbon reacts
In the range of (900 to 1200) ℃, it enters a transition zone where control with oxygen and undergoes gasification. This zone can be characterized
shifts from chemical to diffusion processes. At temperatures between as a packed bed consisting of numerous coke particles and exhibiting a
(1200 to 1300) ℃, diffusion predominantly controls the reaction, and certain degree of porosity. Here, airflow mobilizes coke and engages in
above 1300 ℃, the reaction is entirely governed by the diffusion high-temperature reactions with oxygen. While unburned PC combus­
process. tion occurs in this zone, coke remains the primary combustible material.
The combustion of coke principally yields CO, a vital component in BF
(2) Indirect reduction of iron oxides ironmaking, providing both heat and a reducing agent for BF [88,97].
The combustion zone serves as the origin of the ascending hot gas stream
Thermodynamic analysis of iron oxide reduction reveals that CO and and facilitates a continuous reduction in charge, a result of the space
H2-driven reduction of hematite (Fe2O3) by CO and H2 influences the created by coke gasification. Coke oxidation within this zone is pri­
microstructural evolution of iron ore. Table 4 details the thermodynamic marily categorized into two processes [98]:
data for this reduction reaction. This process involves multiple steps,
including the formation of various structural modifications and complex (i) Near the raceway front, adjacent to the tuyere, carbon undergoes
compositions in intermediate oxides, such as the transformation a complete combustion reaction with oxygen, facilitated by an
sequence from hematite (Fe2O3) ⟶ to magnetite (Fe3O4) ⟶ floater ample supply of O2.
(Fex O) ⟶ and finally to iron (Fe). Typically, these reactions occur
simultaneously [93]. The reduction dynamics of iron oxides are C + O2 = CO2 Δr Hθm (298.15 K) = − 393.5 kJ⋅mol− 1
(5)
temperature-dependent, impacting the formation of Fex O [94]. Kinetic
studies have demonstrated that the reduction of Fe2O3 to metallic iron
(ii) At the back end of the raceway near the deadman, a carbon
follows either a two-step mechanism below 570 ◦ C or a three-step
dissolution loss reaction occurs because of insufficient oxygen
(>570 ◦ C) mechanism. Initially, Fe2O3 is converted to Fe3O4, which
and excess coke.
then forms Fex O at temperatures above 570 ◦ C [95]. Below 570 ◦ C,
magnetite directly to Fe [96], as Fex O is unstable below this threshold. C + CO2 = 2CO Δr Hθm (298.15 K) = + 157.8 kJ⋅mol− 1
(6)

10
W. Niu et al. Fuel 366 (2024) 131277

The oxidation and reduction zones are the two distinct areas within Pavlov believed that indirect reactions accomplished all of the re­
the combustion zone. The disappearance of CO₂ serves as a demarcation action Fe2O3⟶Fe3O4⟶FeO in the BF and that the reduction of iron
for the boundary of the combustion zone. In practical production sce­ from FeO depended partly on the indirect reduction of CO and H2, and
narios, pinpointing the exact point of CO2’s complete disappearance is the rest on the direct reduction of carbon. In his view, the reduction of
challenging; thus, the CO2 content is often reduced to as low as 1 % at iron from FeO partially relies on the indirect reduction by CO and H2,
the boundary of the combustion zone. The reaction process unfolds in with the remainder attributable to the direct reduction by carbon. He
the following manner: oxygen molecules initially adhere to the carbon quantified the directness of iron reduction as the ratio of iron produced
surface. As the temperature rises, carbon atoms and oxygen adsorption through the direct reduction of FeO to the total iron produced by all
intensifies, transitioning from physical to chemical adsorption. This reduction processes. In cases where solid carbon is the reducing agent
change weakens the bond between oxygen atoms, causing the oxygen and the resultant product is CO, the process is termed “direct reduction.”
bond to elongate and ultimately break. Consequently, carbon atoms on This reaction, characterized by strong heat absorption, involves the
the surface form complexes. Due to the influence of surrounding airflow direct consumption of solid carbon. However, the contact conditions
and high temperatures, these surface complexes decompose into CO and between the two solid phases are typically inadequate to sustain a
CO2. This phenomenon is referred to as the primary reaction of com­ perceptible reaction rate. In reality, the direct reduction reaction is
bustion or the main reaction. The reaction formulas are as follows: achieved through the combined effects of carbon dissolution and the
water–gas shift reaction:
<1300 ◦ C 4C + 2O2 = (4C) ⋅ (2O2) (7)
Indirect reduction FeO + C = Fe + CO2 ΔHθ298 = − 13190 J⋅mol− 1
(13)
(4C) ⋅ (2O2) + O2 = 2CO + 2CO2 (8)
Carbon Dissolution Reaction C + CO2 = 2CO ΔHθ298 = − 165390
>1600 ◦ C 3C + 2O2 = (3C) ⋅ (2O2) (9) J⋅mol− 1 (14)
(3C) ⋅ (2O2) = 2CO + CO2 (10)
Indirect reduction FeO + H2 = Fe + H2O ΔHθ298 = − 13190 J⋅mol − 1
(15)
At temperatures ranging from 1300 ◦ C to 1600 ◦ C, two reactions Water gas reaction H2O + C = H2 + CO ΔHθ298 = − 124190 J⋅mol− 1
(16)
occur concurrently, with “complex decomposition” as the prevalent
control mechanism. CO and CO2 produced in the primary reaction Direct restore FeO + C = Fe + CO ΔHθ298 = − 152200 J⋅mol− 1
(17)
persist in reacting either with oxygen or carbon in what constitutes the
secondary or side reactions of the combustion process. (2) Reduction of silicon oxides

2CO + O2 = 2CO2 (11) Since the 1970 s, perspectives on the reduction process of SiO2 in BF
C + CO2 = 2CO (12) have evolved significantly. It is now widely accepted that the silicon in
BF iron melts is derived from SiO2 in liquid slag and gaseous SiO [89]. In
Referring to the CO-CO2-C equilibrium system illustrated in the gas- the high-temperature region of the BF, conditions are conducive to
component change diagram (Fig. 15), it can be inferred that the ther­ forming “SiO.” This gaseous “SiO” is primarily produced through two
modynamic and kinetic conditions of the coke in the tuyere area are key reactions.
conducive to its complete reaction within this region. This reaction ul­
timately produces CO, which serves two critical roles: providing energy (SiO2)slag + Ccoke = {SiO} + {CO} (18)
and generating reducing gas for the process. (SiO2)ash + Ccoke = {SiO} + {CO} (19)

3.2.2. Solid–solid reactions Due to the higher activity of SiO2 in coke ash than in slag [99,100], it
is commonly acknowledged that SiO gas mainly results from reduced
(1) Direct reduction of iron oxides SiO2 in coke ash. As illustrated in Eq. (18), this process is more likely to
occur than the one described in Eq. (19), leading to a greater production
of SiO gas. The SiO thus produced is ultimately reduced by solid coke, as
per the following reaction equation:

(SiO) + Ccoke = [Si] + {CO} (20)

It has been observed that the silicon content in iron droplets pre­
dominantly originates from the ash in coke [99]. Additionally, dissec­
tions of BFs have revealed that temperature significantly influences the
reduction of SiO2 [100]. The variation in SiO2 content in coke within the
BF as a function of temperature increase is depicted in Fig. 16. Notably, a
marked decrease in SiO2 content in coke begins from the soft molten
droplet fallout zone, coinciding with an increase in silicon content in the
iron droplets [101]. Therefore, controlling the SiO2 content in coke and
adjusting the operating temperature of the BF are vital strategies for
managing the silicon content in the iron metal.

3.2.3. Solid–liquid reaction

(1) Carburization reaction

Incorporating carbon into molten iron, known as carburization, is a


critical process in ironmaking. During the refining process, the carbon
content of coke is approximately 85 % to 90 %, influenced by the type of
Fig. 15. Variation of CO and CO2 components in CO-CO2-C system under coking coal used. Apart from carbon, coke contains about 10 % ash and a
different temperature and pressure conditions [16].

11
W. Niu et al. Fuel 366 (2024) 131277

concentration of carbon at the moment of the reaction time (wt.%).


Notably, the saturation concentration of carbon in the iron melt is pre­
dominantly influenced by the temperature and the presence of alloying
elements within the iron melt [104–106]:

CS = 1.3 − 0.00257T(◦ C) − 0.31 Si − 0.33 P − 0.45 S + 0.28 Mn (22)


The primary dissolution reaction rate was determined by integrating
Eq. (21), assuming that A, V, and K are independent variables, which can
be obtained as
∫ Ct ∫
dCt A t
=k dt (23)
C0 Cs − Ct V 0

Cs − Ct A
ln = − k t (24)
Cs − C0 V

where C denotes the initial carbon content of the iron melt.


The Fe-C phase diagram reveals that even trace amounts of carbon
significantly impact the melting point of iron. Enhancing the carburi­
zation rate of coke is crucial for facilitating the melting and reduction of
iron ore. The carburizing capacity of coke influences the saturation
Fig. 16. Variation of SiO2 content in coke with increasing temperature in the degree of the iron melt in the furnace hearth and serves as an indicator of
BF [89]. the cleanliness of the deadman [107]. Theoretical. Both theoretical
calculations and experimental findings indicate that the saturated sol­
minor proportion of volatile substances. Notably, the carburization re­ ubility of carbon in the iron melt is approximately 5.0 %. In practice, the
action dissolves around 7 % to 10 % of this carbon into the pig iron. In carbon content of the iron melt in the furnace hearth typically reaches
the lower region of the BF hearth, coke is the sole solid material. As the about 4.5 % after undergoing a series of complex carburization
iron descends from the dripping zone, it encounters coke and undergoes reactions.
the contact carburization reaction; this occurs at the base of the ’dead Understanding the carburization behavior of coke in iron melts is
man’ in the hearth and eventually leads to the dissolution of carbon into vital for maintaining stable BF operations and making informed ad­
the molten iron [89]. justments to the charge structure. Furthermore, this knowledge signifi­
The carburization process comprises three primary stages, as illus­ cantly reduces fuel costs and is integral to the GHG emission reduction
trated in Fig. 17 [102,103]: ① Carbon atoms detach from the carbo­ strategies. By optimizing the carburization process, ironmaking can be
naceous material matrix and form a reactive interfacial layer on the more efficient and environmentally sustainable [108].
surface of the carburizer. ② These carbon atoms then diffuse into the
molten iron through the interfacial layer. ③ Finally, the carbon atoms (2) Direct reduction of iron oxides
dissolve in the liquid iron. Studies on the carburization process of
graphite have led to the formulation of a specific rate expression for its In another form of direct FeO reduction, the FeO-containing liquid
carburization, which can be articulated as follows: slag directly contacts solid coke or reacts with the saturated carbon in
dCt A the iron melt. Ore, not fully reduced, falls into the high-temperature
= k (Cs − Ct ) (21) zone, softens, melts, and results in an initial slag with high FeO con­
dt V
tent, which then flows downward through the coke voids. Due to
where k denotes the first dissolution reaction rate constant (m/s); t is the effective contact between the liquid slag and coke surface and the lesser
reaction time (s); A is the reaction contact area (m2); V represents the diffusion resistance compared to gas diffusion in zigzagging micropo­
volume of liquid ferric iron (m3); Cs represents the saturation concen­ rous spaces, coupled with high temperatures, the reaction rate constant
tration of carbon in the ferric iron (wt.%); and Ct represents the is considerably large. Consequently, this type of reaction proceeds
rapidly. By the end of the process, the slag’s FeO content is reduced to
less than 0.5 %, and the total iron recovery exceeds 99.7 %.
Thermodynamic studies indicate that the direct reduction reaction
yields a single gas-phase product, CO, as per the following equation:
(FeO) + C = Fe + CO (25)

The equilibrium constant of the reaction can be expressed as:


PCO Ptotal ⋅φ(CO)
K= = (26)
α(FeO) α(FeO)

where K is a function of both temperature, total system pressure (Ptotal ),


and the activity of FeO in the slag.
Kinetic testing of this reduction process has shown that if the
composition of the primary slag is ω(FeO) = 70 %, ω(CaO) =
15 %, ω(SiO2 ) = 15 %, and this slag flows through a red-hot coke
layer with a (20 to 25) mm height. The FeO in the primary slag can be
completely reduced within 3 min when flowing through the coke layer
at system temperatures of 1400 ◦ C and 1500 ◦ C.

Fig. 17. Schematic diagram of the carburizing reaction process [102].

12
W. Niu et al. Fuel 366 (2024) 131277

3.3. Influence of alkali metals and zinc on coke properties engage with carbon. The theory of interlayer compounds posits that
graphitic interlayer compounds form during reactions between alkali
3.3.1. Effect of alkali metals on coke properties metal carbonates and CO2. The electrochemical mechanism indicates
Alkali metals, specifically K and Na, are introduced into the BF that metal ions alter their electronic structure upon contact with the
through the ore. Due to their chemical stability, these metals do not carbon surface and act as electron donors, providing electrons to the
begin to reduce below the cohesive zone. A significant amount of alkali carbon surface. This process speeds up specific reactions, thus enhancing
metals cannot be expelled from the BF; instead, they accumulate cycli­ the overall reaction rate. “Electron transfer” bears close resemblance to
cally in various forms throughout different regions of the BF, as depicted the oxygen transfer mechanism. In the absence of electron transfer
in Fig. 18. In the different areas of the BF, alkali metals exist in distinct during the catalytic process of gasification reactions, O2 is required for
forms: silicates are predominant in the hearth; elemental vapor is the the transfer, similar to the oxygen transfer mechanism [113].
main form in the furnace shaft and bosh, as well as the lower part of the
hearth; and carbonates are prevalent in the upper and middle parts of 3.3.2. Effect of zinc on coke properties
the shaft [109]. K and Na in the process of rising at different. They can The primary source of Zn in BF is ore, with a small portion origi­
interact with other substances as temperatures increase, forming K2O, nating from coke [114,115]. It enters the charge pores through Zn va­
Na2O, or even further reactions that produce silicates and carbonates pors oxidized to gaseous oxides in the gas stream to form fine oxide
[71]. This indicates that alkali metal carbonates are primarily formed particles or condenses on the surface of the charge to form a thin film of
from alkali vapors that rise with the gas flow and circulate in the upper metal or oxide of Zn that accumulates in the furnace. The practice has
part of the BF. shown that the Zn content present in the charge is higher in the range of
The presence of alkali metals in the BF influences the reaction rates (1000 to 1200) ◦ C in the BF, so Zn has a significant effect on the gasi­
of its components. Specifically, the reaction rate of the isotropic fication reaction of coke; the larger the Zn content adsorbed by coke, the
component diminishes, while that of the anisotropic component in­ worse its thermal state performance.
creases. This could be due to alkali metals embedding within the Experimental studies showed that the best adsorption effect of Zn
anisotropic structure, forming a looser structure with the carbon atoms was achieved at 1000 ◦ C; Zn played a positive catalytic role in the
of the carbon layer, thereby facilitating the reaction of carbon dioxide dissolution reaction of coke, and with the increase of Zn adsorption, the
with it [97]. The reactivity of coke notably escalates when the gas–solid reactivity of coke was subsequently enhanced [116]. ZnO and Zn have
mass ratio of K vapor to coke exceeds 3 %. At 1100 ◦ C, the adsorption, no direct catalytic effect on coke, and only under a CO2 atmosphere do
degradation, and catalytic gasification reactions of K vapor in coke are both of them catalyze the coke solvation reaction, leading to poor per­
significantly more pronounced than those of sodium [109]. The struc­ formance of coke in the thermal state [117]. The coke adsorption of Zn
ture of coke is compromised when the proportion of alkali metals rea­ produced more cracks on the substrate, which accelerated the coke
ches 3 %; notably, when the mass ratio of K to Na is 3:7, the CRI peaks, gasification reaction, and the Zn content at the edges of the cracks
and the CSR hits its lowest [111]. increased [118].
Circulating enriched alkali metals in the BF act as catalysts in the
coke dissolution reaction. They increase the rate of this reaction,
accelerating coke gasification. Consequently, coke strength significantly 3.4. Graphitization of coke in the BF
declines after reacting, leading to increased coke pulverization and
decreased permeability in the BF. Alkali metal-catalyzed coke dissolu­ Coke, a porous carbonaceous material, begins to undergo significant
tion mechanisms include the oxygen transfer mechanism, the theory of changes upon entering the cohesive zone of the BF. As it descends, coke
interlayer compounds, and the electrochemical mechanism. Mckee et al. is exposed to high temperatures, causing its microcrystalline structure to
[112] have proposed the oxygen transfer mechanism based on graphite- grow gradually. This process eliminates aberrations and defects in the
CO2 reaction experiments. This mechanism suggests that the catalytic microcrystals, leading to a transformation into graphitized carbon. The
action of alkali metals primarily involves their oxidation–reduction physical phase transformation of ash minerals within the coke and melt
cycle, in which both metals and metal salts facilitate oxygen transfer migration further facilitates this graphitization. The direct interaction
through their oxides and peroxides, allowing reactive oxygen atoms to between coke and the liquid slag iron in the BF’s lower region can also
influence graphitization [119]. It is crucial to understand the factors that

Fig. 18. Schematic diagram of alkali metal circulation in a BF [110].

13
W. Niu et al. Fuel 366 (2024) 131277

affect coke graphitization in the BF and the implications of high in constant contact with coke in various forms, undergoing diverse re­
graphitization on its structure and strength. This knowledge contributes actions [85]. XRD and Raman spectroscopy reveal that both iron and
to the BF’s smooth operation and supports the goals of green and low- slag catalyze coke graphitization, with iron’s effect being more pro­
carbon production. nounced. Metallic iron catalyzes the transformation of SP3 to SP2 carbon
bonds in coke, promoting graphitization. Thermogravimetric analysis
3.4.1. Influencing factors of graphitization indicates that coke analogs containing metallic iron react more swiftly
Temperature is a pivotal factor influencing the degree of graphiti­ with CO2.
zation in coke [48,120]. A plethora of experimental data indicates that a
gradual heating regime enhances the final graphitization quality; higher 3.4.2. Effect of graphitization on coke properties
temperatures improve graphitization and reduce the time needed to The graphitization of coke significantly impacts its particle size, pore
achieve equilibrium. As a result, coke graphitization escalates markedly, structure, strength, and microcrystalline architecture. Gornostayev et al.
transitioning its structure towards an orderly, regular graphite form [127], Gupta et al. [83] and Li et al. [86] have shown that coke’s
[120]. With every 100 ◦ C rise in temperature, the graphitization degree graphitization in high-temperature regions commences at the surface,
deepens approximately 1.8-fold, the d002 value shows a reduction of leading to the formation of coke fines. As graphitization intensifies, the
about 2 %, the La value sees an approximate 3 % increase, and the Lc spacing (d002) between the levels of the coke’s microcrystalline structure
value escalates by around 15 % [121]. The temperature-dependent diminishes, augmenting the interatomic forces within the same layer
graphitization of various coke types is depicted in Fig. 19. Notably, and stabilizing the crystal structure [128,129]. This stabilization man­
the Lc value escalates substantially between 1500 and 1700 ◦ C, under­ ifests as a decrease in the CRI and an elevation in CSR. Additionally, the
scoring the profound impact of temperature on graphitization across the reduced reactivity of graphitized coke and its heightened post-reaction
entire coke mass. strength suggest enhanced structural stability during graphitization
Certain substances in coke ash can act as catalysts for graphitization, [121]. The smaller the particle size of tuyere coke, the higher the degree
enabling coke to graphitize at lower temperatures or to achieve higher of graphitization. Therefore, the graphitization of coke leads to the
graphitization at the same temperature. The ash of coke, which is the flaking of its surface and the formation of fine powdery particles, so the
solid residue after ashing, typically contains various oxides and has a pulverization rate is elevated [85,86]. Moreover, the coke’s surface
mass fraction generally less than 15 %. As Fig. 20(a) illustrates, the stack pores undergo changes: large pores shrink to small or micropores, the
height (Lc) of coke microcrystals incrementally increases with the heat pore walls become thinner, and numerous pits form. These alterations
treatment temperature. This increase becomes more pronounced at could result from the transformation of isotropic substances to aniso­
temperatures ranging from (1600 to 1800) ◦ C. Additionally, prolonging tropic ones within the coke during heat treatment, coupled with an in­
the heat treatment time at constant temperatures enhances the final crease in the coarse-grained mosaic structure on the surface [121].
graphitization degree of coke [122]. As shown in Fig. 20(b), the
graphitization degree of coke rapidly escalates between 1200 and 4. Coke quality requirements and evaluation for modernized BF
1300 ◦ C [120]. The XRD patterns reveal that coke’s ash content trans­
forms at temperatures above 1573 K, evidenced by the gradual dimin­ Addressing the quality of coke entails defining and measuring its
ishing of quartz peaks [123]. properties to meet the requirements of the BF process. This involves
Numerous scholars[85,123–126] have indicated that slag iron cata­ setting production targets for coke based on its property determination.
lyzes coke’s graphitization with a mechanism akin to graphite forma­ The preceding analysis consolidates the role of coke in the BF, its
tion. The catalytic component in slag iron utilizes the free energy degradation mechanisms, and the requisite quality standards, as out­
disparity between ordered and disordered carbon arrangements to lined in Table 5.
dissolve carbon and subsequently crystallize it through deposition. In In various regions of the BF, coke encounters diverse conditions such
the BF’s lower cohesive zone, liquid slag iron forms and, while as pressure, abrasion, gasification, exposure to alkali metals, combus­
descending, traverses the dripping zone and combustion belt before tion, graphitization, and carburization. Consequently, coke must serve
finally reaching the furnace hearth. Throughout this journey, it remains as a structural support, facilitating gas and liquid permeability. Specific
quality indices are essential to mitigate coke degradation and ensure
optimal BF performance. From the lumpy zone to the hearth, coke
should possess strong crushing strength, high abrasion resistance, and
robustness against varying atmospheric conditions, alkali metals, and
other chemical assaults. Additionally, it should maintain effective
carburization with the iron melt.
Highly reactive coke in BFs is advocated to lower the thermal reserve
zone temperature, thereby enhancing the gas utilization factor. This
suggests that the ore properties possess potent reducing capabilities
[130]. Increasing coke’s reactivity can improve the permeability of the
stock column and the high-temperature efficacy of the iron-bearing
charge, reducing the reductant rate and consumption [131]. There­
fore, employing highly reactive coke is recommended for achieving
efficient, low-reductant rate BF operations [132].
Coke performance indexes include microscopic parameters and
macroscopic performance, the former is fundamental to the evolution of
the latter. Microscopic parameters determines macroscopic perfor­
mance, and macroscopic performance is the external expression of
microscopic parameters. Through the discussion in the previous chap­
ters, taking the coke macroscopic performance as the target and coke
microscopic parameters as the variables, the main relevant laws are
obtained, as shown in Fig. 21. In order to pursue high strength and low
Fig. 19. Variation of graphitization with temperature for different coke after fine rate, it is necessary to ensure firstly that the coke matrix structure is
heat treatment [82]. strong enough, secondly that the coke pore structure is uniformly

14
W. Niu et al. Fuel 366 (2024) 131277

Fig. 20. Variation of graphitization degree of coke under different heat treatment conditions [120,122].

influenced by its structure and composition. In BF production, the crit­


Table 5
ical factors impacting coke’s cold strength (M40 and M10) and hot
Coke function, degradation mechanism and quality requirements.
strength (CRI and CSR) include its pore structure, matrix structure,
Position Functions Degradation Quality microcrystalline structure, and mineral composition. Metallurgical
mechanisms requirements
coke, or graphitized carbon, demonstrates decreasing reactivity with
Lumpy Air permeability Alkali deposits, Size, stability, increased Lc and La values. Coke, characterized by highly rounded pores
zone mechanical stress, mechanical strength, and substantial pore wall thickness, exhibits greater strength. The pore
abrasion abrasion resistance
Cohesive Strut, air Gasification reaction, Low carbon dioxide
structure has been analyzed using methods such as drainage, micro­
zone permeability, abrasion reaction, high CRS scopic image analysis, mercury pressure, and gas physisorption,
slag-iron revealing that cokes with a higher proportion of anisotropic structures
discharge tend to be less reactive. The microcrystalline structure of coke has been
Dripping Strut, air Gasification reaction, Size distribution,
characterized through XRD, TEM, HRTEM, XPS, Raman, and FTIR; the
zone permeability, abrasion, alkali metal low carbon dioxide
slag-iron erosion, ash reactions reaction, abrasion mineral composition and distribution within the coke and the detection
discharge resistance methods were also described.
Raceway Production of CO Combustion, Resistant to thermal The evolution of coke within a BF is a progressive and continuous
graphitization, shock, mechanical process, experiencing varying degrees of degradation in different zones.
stress, abrasion stress and abrasion
Hearth Strut, slag-iron Graphitization, Mechanical and
In the BF’s upper region, where temperatures are lower, coke deterio­
discharge, mechanical stress, abrasion strength, ration is mainly physical, attributed to friction, extrusion, and charge
carburization carburization carbon dissolution expansion. Conversely, rapid degradation occurs in the lower high-
temperature zone due to coke’s involvement in multiphase reactions.
As the charge descends, coke reaches the hearth, eventually dissipating
distributed, and lastly that it has appropriate reactivity.
under the influence of slag iron and thermal stress. The cyclical
enrichment of alkali metals and zinc in the BF significantly impacts
5. Conclusions and outlook
coke’s performance. Alkali metals can lower the activation energy
needed for gasification reactions, disrupt the carbon matrix structure,
This paper comprehensively reviews the structure and composition
weaken the carbon boundary connections in graphite microcrystals, and
of coke, its various characterization methods, and summarizes coke’s
catalyze coke reactions. Increased adsorbed Zn content results in
behavior in the BF. The metallurgical properties of coke are primarily
heightened coke reactivity, reduced post-reaction strength, and an
amplified pulverization rate. The increasing heating temperature grad­
ually enhances graphitization as coke moves from top to bottom in the
furnace. This results in a decrease in CRI, an increase in CSR, and a
significant reduction in both gas and liquid permeabilities within the
furnace.
Since the onset of the 21st century, the advancement of large-scale
and intelligent BF has significantly propelled ironmaking technology.
However, challenges such as energy supply constraints and environ­
mental protection requirements have imposed greater developmental
hurdles on the BF ironmaking process, necessitating higher quality
standards for coke. The emergence of new ironmaking technologies, like
hydrogen-rich and oxygen-rich BFs, has led to a notable decrease in the
coke ratio used in BFs, edging closer to the theoretical limit of coke
usage.
Researchers have explored alternatives to coke in BFs in pursuit of
Fig. 21. Relationship between coke microscopic parameters and macro­
green and low-carbon development. Despite these efforts, coke’s role as
scopic properties.

15
W. Niu et al. Fuel 366 (2024) 131277

a structural support for the material column remains indispensable. It [10] Gupta S, Ye ZZ, Kanniala R, Kerkkonen O, Sahajwalla V. Coke graphitization and
degradation across the tuyere regions in a blast furnace. Fuel 2013;113:77–85.
ensures the smooth flow of BF gases and effective penetration of liquid
[11] Li KJ, Khanna R, Zhang JL, Liu ZJ, Sahajwalla V, Yang TJ, et al. The evolution of
slag and iron melt, a function yet to be replicated by other materials. structural order, microstructure and mineral matter of metallurgical coke in a
Research on coke must evolve to foster cleaner production in the steel blast furnace: a review. Fuel 2014;133:194–215.
industry. New techniques should be developed to elucidate the coke [12] Lomas L, Roest R, Gupta S, Pearson RA, Fetscher R, Jenkins DR, et al.
Petrographic analysis and characterization of a blast furnace coke and its wear
microstructure, building upon traditional kinetic methods for studying mechanisms. Fuel 2017;200:89–99.
its macroscopic properties. This approach will enable a comprehensive [13] Chang ZY, Zhang JL, Ning XJ. Phase and mineral behavior of coke in cohesive
systematization of findings related to coke structure and its behavior zone. Fuel 2019;253:32–9.
[14] Lomas H, Roest R, Thorley T, Wells A, Wu H, Jiang Z, et al. Tribological testing of
inside the furnace. metallurgical coke: coefficient of friction and relation to coal properties. Energy
Furthermore, the introduction of a coke-quality evaluation system is Fuel 2018;32:12021–9.
imperative. This system should define quality evaluation indices for [15] Li KJ, Zhang JL, Liu ZJ, Jiang X. Critical analyses about the development of
ironmaking process based on the principle of energy-saving and emission
various types of coke tailored to modernized BF ironmaking processes. reduction. Chinese J Proc Eng 2014;14(1):162–72.
These indices would regulate coke allocation, ensuring that the most [16] Li KJ, Zhang JL, Liu ZJ, Ning XJ, Wang TQ. Gasification of graphite and coke in
suitable types of coke are used in specific processes. This strategic carbon-carbon dioxide-sodium or potassium carbonates system. Ind Eng Chem
Res 2014;53(14):5737.
approach will provide vital guidance for achieving efficient and low- [17] Gupta S, French D, Sokurov R, Grigore M, Sun H, Cham T, et al. Minerals and
consumption BF ironmaking, aligning with the industry’s broader sus­ iron-making reactions in blast furnaces. Prog Energy Combust 2008;34(2):
tainability goals. 155–97.
[18] Vandezande J A. Correlation of coke microstructure and properties, Proceedings
of Ironmaking Conference, Inland Steel Co., East Chicago, IN, Pittsburgh, USA,
CRediT authorship contribution statement 1982;pp.12–23.
[19] Geng W, Nakajima T, Takanashi H, Ohki A. Analysis of carboxyl group in coal and
coal aromaticity by fourier transform infrared (FT-IR) spectrometry. Fuel 2009;
Wenquan Niu: Writing – original draft, Methodology, Investigation,
88:139–44.
Formal analysis, Conceptualization. Yan Li: Writing – review & editing, [20] Pusz S, Krzesinska M, Smedowski L, Majewska J, Pilawa B, Kwiecinska B. Changes
Investigation. Qiang Li: Resources, Methodology. Jingsong Wang: in a coke structure due to reaction with carbon dioxide. Int J Coal Geol 2010;81:
287–92.
Supervision, Resources, Formal analysis, Conceptualization. Guang
[21] Feng L, Zhao GY, Zhao YY, Zhao MS, Tang JW. Construction of the molecular
Wang: Supervision, Investigation. Haibin Zuo: Methodology, Formal structure model of the shengli lignite using TG-GC/MS and FTIR spectrometry
analysis. Xuefeng She: Methodology, Investigation. Qingguo Xue: data. Fuel 2017;203:924–31.
Funding acquisition. [22] Rantitsch G, Bhattacharyya A, Günbati A, Schulten MA, Schenk J, Letofsky-
Papst I, et al. Microstructural evolution of metallurgical coke: evidence from
raman spectroscopy. Int J Coal Geol 2020;227:103546.
Declaration of competing interest [23] Marsh H, Clarke D. Mechanisms of formation of structure within metallurgical
coke and its effect on coke properties. In: First international meeting on coal and
coke applied to ironmaking. Rio de Janeiro, Brazil; 1987.
The authors declare that they have no known competing financial [24] Fortin F, Rouzau JN. Different mechanisms of coke microtexture formation during
interests or personal relationships that could have appeared to influence coking coal carbonization. Fuel 1994;73:795–809.
[25] Sun WZ, Zheng MD, Cui P, Hu DS, Kan X. Novel method for coke optical texture
the work reported in this paper. measurement and application on coke derived from different ranks of coking coal.
Asia-Pac J Chem Eng 2021:e2635.
Data availability [26] Saito Y, Kanai T, Yamazaki Y. Effect of mesoscale pore structure on coke strength.
J Jpn Inst Energy 2017;96(4):93–101.
[27] Liang L, Sun Z, Wei ZK, Guo R. Progress in research of structure and composition
Data will be made available on request. of metallurgical coke. Fuel & Chemical Processes 2019;50(02):18–22.
[28] Fu YN. Coking chemistry. Beijing: Metallurgical Industry Press; 1982.
[29] Zhang YL, Wu WG, Zhao SH, Long YF, Luo YH. Experimental pyrolysis tar
Acknowledgments removal over rice straw char and inner pore structure evolution of char. Fuel
Process Technol 2015;134(1):333–44.
This work was supported by the National Natural Science Foundation [30] Kong DW, Zhang JL, Gong BX, Lin XH, Guo H. Study on coke reactivity in lump
zone of blast furnace. Iron Steel 2011;46(4):15.
of China (grant number U1960205), China Baowu Low Carbon Metal­ [31] Zocdol JW, Velez MR. Development of surface area and pore structure for
lurgy Innovation Foundation (grant number BWLCF202101 and activation of anthracite coal. Fuel Process Technol 2007;88(4):369–74.
BWLCF202104) and China Minmetals Science and Technology Special [32] Guo WT. The influence of the coke microstructure on the reaction behavior and
performance in the blast furnace. Beijing: University of Science and Technology
Plan Foudation (grant number 2020ZXA01).
Beijing; 2016.
[33] Zhang H, Xu RS, Zhang JL, Daghagheleh O, Schenk J, Li CH, et al.
References A comprehensive review of characterization methods for metallurgical coke
structures. Materials 2022;15:174.
[34] Wu XB, Zhang JL, Kong DW, Bai YN, Gao B. Effects of parameters on the
[1] Rahmatmand B, Tahmasebi A, Lomas H, Honeyands T, Koshy P, Hockings K, et al.
measurement of coke porosity. J Iron Steel Res 2012;24(10):59–62.
A technical review on coke rate and quality in low-carbon blast furnace
[35] Nyathi MS, Mastalerz M, Kruse R. Influence of coke particle size on pore
ironmaking. Fuel 2023;336:127077.
structural determination by optical microscopy. Int J Coal Geol 2013;118:8–14.
[2] Holappa L. A general vision for the reduction of energy consumption and CO2
[36] Yamamoto Y, Kashiwaya Y, Nishimura M, Kubota M. 3D analysis of coke
emissions from the steel industry. Metals 2020;10(9):1117.
microstructure using μ-X-ray CT. Tetsu-to-Hagane 2009;95(2):103–11.
[3] Wang YJ, Zuo HB, Zhao J. Recent progress and development of ironmaking in
[37] Numazawa Y, Igawa D, Matsuo S, Saito Y, Matsushita Y, Aoki H, et al. Numerical
China as of 2019: an overview. Ironmak Steelmak 2020;47(6):1–10.
analysis of reaction degradation for three-dimensional coke pore structure. ISIJ
[4] Zhou DD, Cheng SS, Wang YS, Jiang X. The production of large blast furnaces
Int 2018;58(8):1420–6.
during 2016 and future development of ironmaking in China. Ironmak Steelmak
[38] Nyathi MS, Mastalerz M, Kruse R. Influence of coke particle size on pore
2017;44(10):714–20.
structural determination by optical microscopy. Int J Coal Geol 2013;118(10):
[5] Song JY, Jiang ZY, Bao C, Xu AJ. Comparison of energy consumption and CO2
8–14.
emission for three steel production routes-integrated steel plant equipped with a
[39] Yamamoto Y, Kashiwaya Y, Miura S, Nishimura M, Katou K, Nomura S, et al. In
blast furnace, oxygen blast furnace or COREX. Metals 2019;9(3):364.
situ observation and reaction mechanism of iron oxide catalyst added to coke.
[6] Bataille C, Åhman M, Neuhoff K, Nilsson LJ, Fischedick M, Lechtenböhmer S,
Tetsu to Hagane 2010;96(5):297–304.
et al. A review of technology and policy deep decarbonization pathway options
[40] Garrido J, Linares-Solano A, Martin-Martinez JM. Use of N2 vs CO2 in the
for making energy-intensive industry production consistent with the Paris
characterization of activated carbons. Langmuir 1987;3(1):76.
agreement. J Clean Prod 2018;187:960–73.
[41] Grigore M. Factors influencing coke gasification with carbon dioxide. Sydney:
[7] Åhman M, Nilsson LJ, Johansson B. Global climate policy and deep
University of New South Wales; 2007. p. 110–4.
decarbonization of energy-intensive industries. Clim Policy 2017:17.
[42] Fortin F, Rouzaud JN. Different mechanisms of coke micro–texture formation
[8] Godin J, Liu WZ, Ren S, Xu CC. Advances in recovery and utilization of carbon
during coking coal carbonization. Fuel 1994;73:795–809.
dioxide: a brief review. J Environ Chem Eng 2021;9(4).
[43] Hou CX, Lin JX, Guo R, Cheng H. Effects of microscopic structure on coke solution
[9] Wang XL. Iron and steel metallurgy (ironmaking). 3th ed. Beijing: Metallurgical
loss and deterioration. Fuel & Chemical Processes 2019;50(03):6–10.
Industry Press; 2013.

16
W. Niu et al. Fuel 366 (2024) 131277

[44] Coin C. Coke microtextural description comparison of nomenclature, [79] Ma HX, Zhang JL, Jiao KX, Chang ZY, Wang YJ, Zheng PC. Analysis of erosion
classification and methods. Fuel 1987;66:702–5. characteristics and causes of blast furnace hearth. Iron Steel 2018;53(9):14–9.
[45] Zhang WJ, Shi T, Zhang Q, Cao Y, Qian H, Wu X, et al. Coke texture, reactivity [80] Chung JK, Kim SM, Lee JH. Effect of coke properties on blast furnace condition,
and tumbler strength after reaction under simulated blast furnace conditions. Scanmet III Proceedings Lulea, Sweden 2008;June 8–11.
Fuel 2019;251:218–23. [81] The ISIJ. Blast furnace phenomena and modeling. London: Elsevier Applied
[46] Zhang Z, Xie F, Zheng Y, Cheng H, Wang Q. Coke microcrystalline texture and its Science Publishers; 1987. p. 502.
effect on coke reactivity. J Fuel Chem Technol 2018;46(04):406–12. [82] Vander Velden B, Atkinson CJ, Bakker T, et al. Coke properties and its processes
[47] Kashiwaya Y, Ishii K. Kinetic analysis of coke gasification based on non-crystal/ in the lower zone of the blast furnace. World Iron and Steel 2014;2:1.
crystal ratio of carbon. ISIJ Int 1991;31(5):440–8. [83] Gupta S, Ye ZZ, Kanniala R, Kerkkonen O, Sahajwalla V. Coke graphitization and
[48] Hu DS. Crystallite structure characteristics of coke. Iron Steel 2006;41(11):10. degradation across the tuyere regions in a blast furnace. Fuel 2013;113.
[49] Cuesta A, Dhamelincourt P, Laureyns J, Martínez-Alonso A, Tascón J. Raman [84] Xing X, Rogers H, Zhang GQ, Hockings K, Zulli P, Ostrovski O. Coke degradation
microprobe studies on carbon materials. Carbon 1994;32:1523–32. under simulated blast furnace conditions. ISIJ Int 2016;56(5):786–93.
[50] Li XJ, Hayashi J, Li CZ. FT-raman spectroscopic study of the evolution of char [85] Li KJ, Zhang JL, Liu Z, Barati M, Zhong JB, Wei MF, et al. Interfaces between
structure during the pyrolysis of a victorian brown coal. Fuel 2006;85:1700–7. coke, slag, and metal in the tuyere level of a blast furnace. Metall Mater Trans B
[51] Nomura S, Ayukawa H, Kitaguchi H, Tahara T, Matsuzaki S, Naito M, et al. 2015;46(3):1104–11.
Improvement in blast furnace reaction efficiency through the use of highly [86] Li KJ, Zhang JL, Sun MM. Existence state and structures of extracted coke and
reactive calcium rich coke. ISIJ Int 2005;45(3):316–24. accompanied samples from the tuyere zone of a large-scale blast furnace. Fuel
[52] Grigore M, Sakurovs R, French D, Sahajwalla V. Coke gasification: the influence 2018;225.
and behavior of inherent catalytic mineral matter. Energy Fuel 2009;23(4): [87] Chung JK, Lee SM, Shin MS. Effect of coke size on reducing agent ratio (RAR) in
2075–85. blast furnace. ISIJ Int 2018;58(12):2228–35.
[53] Fan YG, Hu SX, Wang L, Wang YD, He B, Zhu ZZ. Modification mechanism and [88] Meng S, Jiao KX, Zhang JL, Wang C, Zhang L, Guo ZY, et al. Analysis of the coke
production practice of coke sprayed with ZBS modifier. Iron Steel 2021;56(1): distribution characteristics in hearth based on blast furnace dissection. Fuel
127–32. Process Technol 2023;242.
[54] Qiu SX, Zhang SF, Zhu RJ, Wu Y, Qiu GB, Dang J, et al. Influence of TiO2 addition [89] Yasuo O, Yasuhito S. Blast furnace phenomena and modeling. Boston: Kluwer
on the structure and metallurgical properties of coke. Int J Coal Prep Util 2018;41 Academic Pub; 1987.
(7):521–37. [90] Calo J, Perkins M. A heterogeneous surface model for the steady-state kinetics of
[55] Jing XL, Wang ZQ, Zhang Q, Yu ZL, Li CY, Huang JJ, et al. Evaluation of CO2 the boudouard reaction. Carbon 1987;25(3):395–407.
gasification reactivity of different coal rank chars by physicochemical properties. [91] Zhang J. Experimental investigation and modeling of gasification of graphite,
Energy Fuel 2013;27(12):7287–93. coke, and char. School of Materials Science and Engineering, Sydney: The
[56] Huang YQ, Yim XL, Wu CZ, Wang CW, Xie JJ, Zhou ZQ, et al. Effects of metal University of New South Wales; 2013. p. 285.
catalysts on CO2 gasification reactivity of biomass char. Biotechnol Adv 2009;27 [92] Edwards IA, Menendez R, Marsh H. Introduction to carbon science. Stoneham,
(5):568–72. MA (USA): Butterworth-Heinemann; 2013. p. 21–2.
[57] Sakawa M, Sakurai Y, Hara Y. Influence of coal characteristics on CO2 [93] Ramadan W, Zaki MI, Fouad NE, Mekhemer GAH. Particle characteristics and
gasification. Fuel 1982;61(8):717–20. reduction behavior of synthetic magnetite. J Magn Magn Mater 2014;355:
[58] Yang JH, Feng AZ, Du HG. Relationship between mineral catalytic index and coke 246–53.
reactivity. Iron Steel 2001;36(6):5–9. [94] Pineau A, Kanari N, Gaballah I. Kinetics of reduction of iron oxides by hi: part II.
[59] Sakurovs R, French D, Grigore M. Quantification of mineral matter in commercial low temperature reduction of magnetite. Thermochemical Acta 2007;456(2):
cokes and their parent coals. Int J Coal Geol 2007;72(2):81–8. 75–88.
[60] Kimura T. Relationships between inorganic elements and minerals in coals from [95] Chen ZY, Dang J, Hu X, Yan HY. Reduction kinetics of hematite powder in
the Ashibetsu district, Ishikari coal field. Japan Fuel Process Technol 1998;56(1): hydrogen atmosphere at moderate temperatures. Metals 2018;8:751.
1–19. [96] Wagner D, Devisme O, Patisson F, Ablitzer D. A laboratory study of the reduction
[61] Ward CR, Taylor JC, Matulis CE, Dale LS. Quantification of mineral matter in the of iron oxides by hydrogen TMS fall extr. Process Divis Sohn Int Symp 2006;2:
argonne premium coals using interactive rietveld-based X-ray diffraction. Int J 111–20.
Coal Geol 2001;46(2):67–82. [97] Li KJ. Structural evolution behavior and muti-phase reaction mechanism of coke
[62] Lv QQ. Study on coke degradation mechanism and reaction kinetics in a large in blast furnace. Beijing: University of Science and Technology Beijing; 2017.
blast furnace. Wuhan University of Science and Technology; 2021. [98] Zhang SF, Wen LY, Bai CG. A mathematical model for the combustion process of
[63] Wu K, She Y, Liu QH, Zhan WL. Consideration on properties of coke and coke pulverized coal at the blast furnace tuyere and coke in raceway. J Chongqing
deterioration after large-scale blast furnace. Iron Steel 2017;52(10):1. Univ Nat Sci Ed 2005;28(8):42–5.
[64] Yasuo O, Yasuhito S. Blast furnace phenomena and modelling. Boston: Kluwer [99] Klimczyk A, Stachura R, Bernasowski M. Silicon behavior at the blast furnace
Academic Pub; 1987. process. J Achiev Mater Manuf Eng 2012;55(2):712–5.
[65] Liu ZL, Wang JL. Theory and Technology of Iron Making. Beijing: Chemical [100] Gustavsson J. Reactions in the lower part of the blast furnace with focus on
Industry Press; 2009. silicon. Royal Institute of Technology Ph.D.:; 2004.
[66] Piqueras P, Sanchis EJ, Herreros JM, Tsolakis A. Evaluating the oxidation kinetic [101] Gornostayev SS, Heikkinen EP, Heino JJ, Fabritius TMJ. Fe-si particles on the
parameters of gasoline direct injection soot from thermogravimetric analysis surface of blast furnace coke. Int J Miner Metall Mater 2015;22(7):697–703.
experiments. Chem Eng Sci 2021;234:1–9. [102] Gudenau H, Mulanza J, Sharma DGR. Carburization of hot metal by industrial and
[67] Wang ZM. Gas-solid reaction mechanism of coke and iron oxide in blast furnace. special cokes. Steel Res Int 1990;61(3):97–104.
University of Science and Technology Beijing; 2021. [103] Olsson R, Koump V, Perzak T. Rate of dissolution of carbon in molten fe-C alloys.
[68] Qie YN, Lyu Q, Liu XJ, Li JP, Lan CC, Zhang SH, et al. Effect of hydrogen addition AIME Met Soc Trans 1966;236(4):426–9.
on softening and melting reduction behaviors of ferrous burden in gas-injection [104] Wu C, Sahajwalla V. Dissolution rates of coals and graphite in fe-C-S melts in
blast furnace. Min Metals Mater Soc ASM Int 2018;49(5):2622–32. direct ironmaking: ependence of carbon dissolution rate on carbon structure.
[69] Chen ZM, Ye SX, Geng SH, Zhu K, Zhang YW, Zou XL, et al. Softening and melting Metall Mater Trans B 2000;31(1):215–6.
performance of mixed burden under the simulated hydrogen-rich blast furnace [105] Wu C, Sahajwalla V. Dissolution rates of coals and graphite in fe-C-S melts in
atmosphere. Ironmak Steelmak 2022;50(5):538–47. direct ironmaking: influence of melt carbon and sulfur on carbon dissolution.
[70] Lan CC, Zhang SH, Liu XJ, Lyu Q, Jiang MF. Change and mechanism analysis of Metall Mater Trans B 2000;31(2):243–51.
the softening-melting behavior of the iron-bearing burden in a hydrogen-rich [106] Neumann F, Schenck H, Patterson W. Influence of iron companions on carbon
blast furnace. Int J Hydrogen Energy 2020;45(28):14255–65. solubility, carbon activity and saturation point in cast iron. Giesserei 1960;47:
[71] Bai L, Zhang JL, Guo H, Zhang XS, Zhang XD. Circulation and enrichment of alkali 25–32.
metal in blast furnace. Journal of Iron and Steel Research 2008;20(9):5. [107] Nightingale R, Dippenaar R, Lu WK. Developments in blast furnace process
[72] Chen YB, Zuo HB. Review of hydrogen-rich ironmaking technology in blast control at port kembla based on process fundamentals. Metall Mater Trans B
furnace. Ironmak Steelmak 2021;48(6):749–68. 2000;31(5):993–1003.
[73] Zhou DD, Cheng SS. Measurement study of the PCI process on the temperature [108] Sun MM. Analysis of iron-carbon interface behavior and carburizing process in
distribution in raceway zone of blast furnace by using digital imaging techniques. blast furnace. Beijing: University of Science and Technology; 2022.
Energy 2019;174:814–22. [109] Zhou HB, Cheng SS. New cognition on coke degradation by potassium and sodium
[74] Zhang CL, Vladislav L, Xu RS, Sergey G, Jiao KX, Zhang JL, et al. Blast furnace in alkali enriched regions and quantificational control model for BF. J Univ Sci
hydrogen-rich metallurgy-research on efficiency injection of natural gas and Technol Beijing 2012;34(3):333–41.
pulverized coal. Fuel 2022;311. [110] Zhao H, Cheng S. New cognition on coke degradation by potassium and sodium in
[75] Cui P, Zhang L, Yang M, Wang Y. Study on kinetics and model of coke loss alkali enriched regions and quantificational control model for blast furnace.
reaction with CO2 in blast furnace. J Fuel Chem Technol 2006;34(3):280. J Univ Sci Technol B 2012;34(3):333–41.
[76] Li DT, Zhu WC, Chen H. Exploration of coking at the 5500m3 blast furnace tuyere [111] Zheng PC, Zhang JL, Liu ZJ, Chen YB. Effect of alkali metals on thermal properties
of shougang jingtang. Ironmaking 2019;38(4):23–6. of coke. China Metallurgy 2017;27(05):19–26.
[77] Pang KL, Meng XY, Zheng YZ, Liu FJ, Wan CR, Gu ZY, et al. Effect of gasification [112] Mckee DW, Chatterji D. The catalytic behavior of alkali metal carbonates and
reaction on pore structure, microstructure, and macroscopic properties of blast oxides in graphite oxidation reactions. Carbon 1975;13(5):381.
furnace coke. Fuel 2023;350. [113] Wang YG, Xie KC, Ling KC. Action mechanism of alkali metal catalysts in coal
[78] Meng S, Jiao KX, Zhang JL, Zong YB, Zhang L, Ma HB, et al. Dissection study of gasification. J Taiyuan Univ Technol 1988;3:52.
the deadman in a commercial blast furnace hearth. Fuel Process Technol 2021;
221:906–11.

17
W. Niu et al. Fuel 366 (2024) 131277

[114] Malsuo T, Yamazaki I, Masuda S. Development of new hot metal [124] Aladejebi OA, Monaghan BJ, Reid MH, Panhuis MIH, Longbottom RJ. Metallic
dephosphorization process in top and bottom blowing converter. In: 73rd iron effects on coke analog carbon bonding and reactivity. Steel Res Int 2017;88
Steelmaking Conference Proceedings. Detroit: ISS-AIME; 1990. p. 115. (10):1700039.
[115] Villarreal RR, Ramirez CR. Dephosphorization in converter. In: 79th Steelmaking [125] Guo ZY, Jiao KX, Zhang JL, Ma HB, Meng S, Wang ZY, et al. Graphitization and
Conference Proceedings. Pittsburgh: ISS-AIME; 1996. p. 173. performance of deadman coke in a large dissected blast furnace. ACS Omega
[116] Mu L, Liu LT, Zheng HY, Wei G, Shen MF. Effect of zinc on reactivity index of 2021;6(39).
coke. Iron Steel 2011;46(8):22. [126] Charon E, Rouzaud JN, Aléon J. Graphitization at low temperatures
[117] Huang KP, Zhong JB, Zheng PC, Liu ZJ, Chen YB. Effect of zinc on thermal (600–1200℃) in the presence of iron implications in planetology. Carbon 2014;
performance of coke. Hebei Metallurgy 2017;11:13. 66:178.
[118] Li M. Effect of zinc on the properties of coke in blast furnace. The 11th [127] Gornostayev SS, Härkki JJ. Graphite crystals in blast furnace coke. Carbon 2007;
Proceedings of China Iron and Steel Annual Meeting. Beijing: The Chinese Society 45(6):1145–51.
for Metals 2017;1. [128] Zhu H, Zhan W, He ZJ, Yu YC, Pang QH, Zhang JH. Pore structure evolution
[119] He ZJ. Slag composition to coke performance influence experimental study. during the coke graphitization process in a blast furnace. Int J Miner Metall Mater
Anshan: University of Science and Technology Liaoning; 2018. 2020:27(9).
[120] Gupta S, Sahajwalla V, Burgo J, Chaubal P, Youmans T. Carbon structure of coke [129] Sun MM, Zhang JL, Li KJ, Guo K, Wang HY, Wang ZM, et al. Influence of structure
at high temperatures and its influence on coke fines in blast furnace dust. Metall and mineral association of tuyere-level coke on gasification process. Metall Mater
Mater Trans B 2005;36(3):385. Trans B 2018;49(5):2611.
[121] Zhan WL, Sun C, Yu YC, Pang QH, Zhang JH, He ZJ. Evolution of coke [130] Gavel DJ. A review on nut coke utilization in the ironmaking blast furnaces. Mater
microstructure and metallurgical properties during graphitization in a blast Sci Technol 2017;33(4):381–7.
furnace. Chinese J Eng 2018;40(06):690. [131] Hu XP. Influence of coke reactivity improved on high temperature properties of
[122] Kawakami M, Kanba H, Sato K, Takenaka T, Gupta S, Chandratilleke R, et al. iron bearing burden. ISIJ Int 2015;9:1859–65.
Characterization of thermal annealing effects on the evolution of coke carbon [132] Natsui T, Nakano K, Matsukura Y, Sunahara K, Ujisawa Y, Inada T. Evaluation of
structure using raman spectroscopy and X-ray diffraction. ISIJ Int 2006;46(8): sinter reducibility and coke reactivity by experimental blast furnace. J Iron Steel
1165. Inst Jpn 2013;99(4):267–74.
[123] Lundgren M, Khanna R, Ökvist S, Sahajwalla V, Björkman B. The evolution of
structural order as a measure of thermal history of coke in the blast furnace.
Metallurgical and Materials Transactions 2014;45(2):603.

18

You might also like