You are on page 1of 200

Xuewei 

Lv
Zhiming Yan

High Temperature
Physicochemical
Properties of High
Alumina Blast
Furnace Slag
High Temperature Physicochemical Properties
of High Alumina Blast Furnace Slag
Xuewei Lv · Zhiming Yan

High Temperature
Physicochemical Properties
of High Alumina Blast
Furnace Slag
Xuewei Lv Zhiming Yan
College of Materials Science WMG
and Engineering University of Warwick
Chongqing University Coventry, UK
Chongqing, China

ISBN 978-981-19-3287-8 ISBN 978-981-19-3288-5 (eBook)


https://doi.org/10.1007/978-981-19-3288-5

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Singapore Pte Ltd. 2022
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Preface

With the consumption of high-grade iron ore resources, the use of low-grade and
difficult-to-handle iron ore will become an alternative way to alleviate the depletion
of iron ore resources. High alumina iron ore will become an important resource for
future blast furnace ironmaking process, resulting in the high alumina blast furnace
slag. This book shows the high-temperature physicochemical properties and structure
of high alumina slag in the ironmaking process. The book consists of six chapters
demonstrating the effect of Al2 O3 on the properties and structure of slag. Based on
the experimental research and practice requirements, a revolutionary technical route
for blast furnace smelting of high alumina iron ore was proposed. This book could
provide a scientific basis and theoretical guidance for the large-scale utilization of
high alumina iron ore in the ironmaking process. This book is interesting and useful
to the industrial workers of metallurgy engineering, as well as the students and
researchers who work in metallurgical field.

Chongqing, China Xuewei Lv


Coventry, UK Zhiming Yan

v
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Metallurgical Slag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Blast Furnace Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.1 Process Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.2 Formation of Blast Furnace Slag . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 The Main Physicochemical Properties of Blast Furnace Slag . . . . . . 7
1.3.1 Liquids Temperature and Fluidity Temperature . . . . . . . . . . . 7
1.3.2 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3.3 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3.4 Surface Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.5 Sulfide Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.3.6 Electrical Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4 Blast Furnace Ironmaking Requirements for Slag Properties . . . . . . 18
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2 Phase Diagram and Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.1 CaO–SiO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2 CaO–Al2 O3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.3 Al2 O3 –SiO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.4 CaO–SiO2 –Al2 O3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.5 CaO–SiO2 –Al2 O3 –MgO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3 Slag Structure of High Alumina Blast Furnace Slag . . . . . . . . . . . . . . . 43
3.1 Basic Concepts of Slag Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.1.1 Components and Classification of the Blast Furnace
Slag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.1.2 Composite Anions in the Slag . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2 Parameters to Represent the Structure of Slags . . . . . . . . . . . . . . . . . . 46
3.2.1 Basicity (R) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2.2 Optical Basicity () . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

vii
viii Contents

3.2.3 Bridging Oxygen (O0 ), Non-bridging Oxygen (O− )


and Free Oxygen (O2− ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.2.4 NBO/T(Q) and Qn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.3 Characterization Methods of Slag Structure . . . . . . . . . . . . . . . . . . . . 50
3.4 Structure of Silicate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.5 Structure of Aluminosilicate Blast Furnace Slag . . . . . . . . . . . . . . . . 53
3.6 Effect of Alumina on the Structure of Blast Furnace Slag . . . . . . . . . 55
3.6.1 Molecular Dynamics Simulation . . . . . . . . . . . . . . . . . . . . . . . 57
3.6.2 Analysis of Slag Structure by Raman Spectroscopy . . . . . . . 69
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4 High-Temperature Physicochemical Properties of High
Alumina Slag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.1 Liquids Temperature and Fluidity Temperature . . . . . . . . . . . . . . . . . 77
4.2 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.2.1 Effect of Al2 O3 Content . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.2.2 Effect of Al2 O3 ↔ SiO2 Substitution . . . . . . . . . . . . . . . . . . . 83
4.2.3 Relationship Between Slag Viscosity and Its Structure . . . . . 84
4.3 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.4 Surface Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.5 Sulfide Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.5.1 The Effect of Al2 O3 Content and SiO2 ↔ Al2 O3
Substitution on the Slag Sulfide Capacity . . . . . . . . . . . . . . . . 91
4.5.2 Relationship Between Sulfide Capacity and Structure . . . . . 94
4.6 Electrical Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5 Estimation of High Alumina Blast Furnace Slag Properties . . . . . . . . . 103
5.1 Types of Estimation Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.1.1 Numerical Fitting Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.1.2 Thermodynamic Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.1.3 Structural-Based Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.1.4 Computer Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.1.5 Artificial Neural Network Model (ANN) . . . . . . . . . . . . . . . . 105
5.2 Liquidus and Solidus Temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.3 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.3.1 Overview of Viscosity Models . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.3.2 Structure-Based Viscosity Modeling . . . . . . . . . . . . . . . . . . . . 118
5.3.3 Iso-viscosity of High Alumina Blast Furnace Slag . . . . . . . . 135
5.4 Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
5.4.1 Overview of Density Models . . . . . . . . . . . . . . . . . . . . . . . . . . 136
5.4.2 Density of High Alumina Blast Furnace Slag . . . . . . . . . . . . . 137
5.5 Surface Tension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
5.5.1 Overview of Surface Tension Models . . . . . . . . . . . . . . . . . . . 137
Contents ix

5.5.2 Iso-surface Tension of High Alumina Blast Furnace


Slag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
5.6 Sulfide Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
5.6.1 Overview of Sulfide Capacity Models . . . . . . . . . . . . . . . . . . . 146
5.6.2 Structure-Based Sulfide Capacity Modeling . . . . . . . . . . . . . . 146
5.6.3 Iso-sulfide Capacity of High Alumina Blast Furnace
Slag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
5.7 Electrical Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
5.7.1 Overview of Electrical Conductivity Models . . . . . . . . . . . . . 152
5.7.2 Iso-electrical Conductivity of High Alumina Blast
Furnace Slag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
6 The Revolution of High Alumina Slag in Blast Furnace Process . . . . . 183
6.1 The Inevitability of High Alumina Blast Furnace Slag . . . . . . . . . . . 183
6.1.1 Iron and Steel Industry in the World . . . . . . . . . . . . . . . . . . . . 183
6.1.2 Iron Ore Resources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
6.2 Current Technical Routes for High Alumina Blast Furnace
Slag . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
6.2.1 Basicity Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
6.2.2 MgO/Al2 O3 Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
6.3 Feasibility of Revolution High Alumina Blast Furnace Slag . . . . . . . 190
6.3.1 Slag-Metal Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
6.3.2 Hot Metal Quality Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
6.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
Chapter 1
Introduction

1.1 Metallurgical Slag

Metallurgy defines as the aggregate of physic-chemical processes for metal extraction


from ores or other metal-bearing materials to produce different types of metals. The
basic processes of extractive metallurgy are pyrometallurgical, hydrometallurgical,
electrochemical metallurgy, biometallurgy, etc. Steel is considered to be the metal
most closely related to the development and technological progress of humans and
is almost omnipresent in our daily life. It has experienced explosive growth in the
past six decades, and its annual production exceeds by far that of all the other metals
combined (Fig. 1.1).
Nowadays, the production of steel is mainly carried out through one of the
following routes, as shown in Fig. 1.2:
• Integrated ironmaking and steelmaking route: Iron ore is used as raw material
to produce steel products following the sequence granulation-blast furnace (BF)-
basic oxygen furnace (BOF)-refining-continuous casting-rolling. Approximately
72% of the world’s steel is produced using the BF-BOF route (according to the
Worldsteel Association report) [1].
• Electric arc furnace steelmaking route: Recycled scrap or reduced iron ore are
used as raw materials. The electric arc furnace (EAF) and the refining process
are the main stages of its steel production. 28% of the world’s steel production is
obtained by this route, an instrument available to recycle the steel in the iron and
steelmaking industry.
• Smelting/Direct reduction ironmaking route: In order to replace the blast furnaces’
ironmaking process, some alternative ironmaking processes have been devel-
oped in the past few decades, including smelting reduction ironmaking (COREX,
FINEX, HIsmelt, Hisarna) and direct reduction ironmaking (MIDREX, HYL,
Rotary kiln, Rotary hearth furnace). However, the product of these methods is hot
metal or sponge iron, which needs further steelmaking process (BOF or EAF) and
refining to obtain the final steel products.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 1
X. Lv and Z. Yan, High Temperature Physicochemical Properties of High Alumina Blast
Furnace Slag, https://doi.org/10.1007/978-981-19-3288-5_1
2 1 Introduction

Fig. 1.1 World crude steel production from 1900 to 2020 (a), and world production of steel,
aluminum, copper, zinc, and magnesium in 2000 and 2020 (million tons) (b)

In the pyrometallurgical industry, slag is another by-product other than metals in


the extraction and refining processes. It can be divided into different types according
to the different sources and purposes:
• Reduce/smelting slag, which is formed by unreduced oxides with added flux in
reduction smelting processes when using ore or concentrates as raw materials,
such as blast furnace slag, submerged arc furnace slag.
• Oxidation slag, which is formed by the oxides from oxidation of some elements
and added fluxes in the oxidative smelting process. Such as converter slag, copper
smelting converter slag.
• Concentration slag, which enriches the target components to the slag from
raw materials to facilitate the processing of the following process. Such as
high-titanium slag obtained by reduction smelting of titanium concentrates, and
vanadium slag.
• Synthetic slag, which is synthesized in advance based on its metallurgical function
using slag-making materials. Such as secondary refining slag, mold flux, and
electroslag remelting slag.
Therefore, slag is a multicomponent melt composed of oxides unreduced in reduc-
tion smelting, oxides formed by oxidation in oxidative smelting, flux added to meet
the smelting requirements, oxides from the corroded refractory lining, and a small
amount of sulfides, chlorides and metals. Table 1.1 lists the typical components of
some metallurgical slags. The various types of slag listed above play an important
1.1 Metallurgical Slag 3

Fig. 1.2 Ironmaking and steelmaking routes

in industry and well-illustrated by the adage, “Look after the slag and the metal will
look after itself.” Their metallurgical roles in separating metal and gangue, removing
impurities from metal that are harmful to the performance of products, enriching
useful elements and refining metals. It also can protect metals from the environment
and reduce the heat loss of metal. Slag plays a significant role in ensuring smooth
smelting operation, the composition and quality of the metal, the metal yield, and
the various technical and economic indicators of the smelting. Of course, the slag
also has some adverse effects, which corrodes the refractory and reduces the furnace
lining life; the entrainment of metal in the slag reduces the metal yield; too much
slag will increase the energy consumption.
The metallurgical functions of slag are realized by controlling the chemical
composition and temperature of the slag to adjust its physicochemical properties.
The physicochemical properties of the slag are related to its structure, so the study
of slag structure can give a better understanding of the changes in slag properties,
which is beneficial to control the metallurgical processes.
4

Table 1.1 The typical components of metallurgical slags (wt.%)


Slag type CaO SiO2 MgO Al2 O3 FeO S TiO2 Others
Reduce/smelting slag Blast furnace slag (General) 27–48 28–39 3–10 6–18 <1.0 0.4–0.7 <2.0 –
Blast furnace slag (Titanium magnetite) 26–30 24–28 ~8 ~14 <1.0 ~0.5 20–25 –
Electric slag (Reduction period) 50–55 15–18 0–10 6–7 <1.0 – – MnO <0.5
Submerged arc furnace slag (Si–Mn alloy) 20–28 38–42 1–4 13–21 – – – MnO 4–8
Oxidation slag Electric slag (Oxidation period) 40–50 12–20 7–12 3–5 8–15 – – MnO 5–10
Converter slag 36–55 18–25 5–7 1–2 7–11 – – P2 O5 1–2
Copper smelting converter slag – 22–24 1–5 1–5 45–50 1–2 – Cu 1.5–2.5
Concentration slag High-titanium slag 1–6 0.5–5 0.4–8 0.5–3 5–15 – 75–94 –
Vanadium slag – 20–24 – – 28–42 – – V2 O5 9–16
Synthetic slag Secondary refining slag ~55 – – ~45 – – – CaF2 <10
Mold flux 2–20 30–50 – 5–20 – – – C 0–20
Electroslag remelting slag 4–5 0–3 – 25–30 – – – CaF2 ~70
1 Introduction
1.2 Blast Furnace Process 5

1.2 Blast Furnace Process

1.2.1 Process Outline

The modern blast furnace process can be considered the most effective ironmaking
process at present. More than 1.2 billion tons of hot metal were produced by the blast
furnace process in 2020, accounting for 62.8% of steelmaking metallic charge [2].
Figure 1.3 shows a schematic representation of a typical blast furnace with all its
auxiliary facilities.
The blast furnace is operated in countercurrent with a descending solids charge
that is heated and reacts with the ascending flow of reductant gases. At the top of the
furnace, input raw materials include iron-bearing burden, slag formers, and coke are
charged in different layers. The iron-bearing burden consists of sinter, pellet, and/or
lump ore in different ratios. Slag formers are usually limestone and dolomite in small
amounts to control the slag chemistry. As heating and reducing agent, coke is also a
supporting structure for burden in the furnace to ensure good gas permeability. Iron
oxides mainly present as Fe2 O3 and Fe3 O4 , small amount of FeO in iron-bearing
burden. Currently, sinter and pellet are self-fluxing, containing the necessary amount
of slag former. These charges descend due to the gravity. The air or moisture/oxygen-
enriched air is heated in hot stove and brings an important amount of heat to the

Fig. 1.3 Schematic representation of a typical blast furnace plant


6 1 Introduction

furnace. In the lower part of the furnace, hot blast from hot stoves is injected through
tuyeres and reacts with the coke. The CO is formed and ascends in the furnace to
reduce the iron oxides. The other sources of carbon materials as reducing agents are
also used by injecting through the tuyeres, such as pulverized coal, hydrocarbons,
and biomass. At the bottom of the furnace, the products of hot metal and slag are
collected in hearth as molten state, and the separation occurs because of the density
difference. Hot metal is carbon-saturated, with about 4.5 wt% carbon content, and
small quantities of silicon, manganese, sulfur, and phosphorus. The hot metal is
tapped from the bottom of the furnace into torpedo ladles and immediately sent
to the steelmaking plant. Normally, the slag is tapped with the molten iron and
separated outside of the furnace, the solidified slag is used as road aggregate or for
cement manufacture. Other possibilities of the slag are to produce mineral wool and
refractory bricks.
The temperature differences inside the blast furnace are large. In the bottom part,
the temperature of the heart is about 1450–1550 °C to keep the slag and hot metal
in liquidus state. In the front of tuyeres, the highest temperature around 2400 °C is
observed due to the coke combustion. The cohesive zone is located in the belly where
temperatures are between 1000 and 1350 °C. In this region, primary slag with high
content of FeO forms and molten iron drops downwards through the coke bed. In
the shaft, the thermal reserve zone with temperatures from 800 to 1100 °C, the iron
oxides are reduced to wustite by carbon monoxide and can be further reduced to iron.
This zone can take up to 60% of the height of the furnace and has a relatively constant
temperature. In the upper part of the furnace, the charges are heated, evaporation and
decomposition reactions take place. The off-gas with dust out from the top of the
furnace with a temperature 150–200 °C with about 15–20 vol% CO2 , 20–30 vol%
CO, 50–60 vol% N2 . After gas cleaning, the dust is recycled, and the gas can be used
for the stoves as fuel.
Due to ecological and economic reasons, many technologies to improve smelting
intensity and decrease the coke rate are developed, such as a better burden distribu-
tion, high-quality sinter and pellets, high hot blast temperature, high-pressure opera-
tion, enrich oxygen/moisture in hot blast, and injection of auxiliary reducing agents.
Technologies to reduce coke ratio benefits reduce CO2 emissions and promote carbon
neutrality.

1.2.2 Formation of Blast Furnace Slag

The slag starts to form from the cohesive zone where the ferric charges (lump ore,
sinter, and pellets) begin to soften, called the primary slag. The primary slag is mainly
composed of gangue and unreduced iron and manganese oxides (FeO and MnO) from
the raw materials. During the descending process, the primary slag contacts with the
coke and reduces gas to cause the reduction reactions. The iron and manganese
oxides content decrease due to the continuous reduction. At the same time, the ash
from the coke is absorbed. If fluxes, such as lime and dolomite (CaO, MgO), are
1.3 The Main Physicochemical Properties of Blast Furnace Slag 7

added, they will also be absorbed during the descent process. The decrease of FeO
content, which can reduce the melting temperature of slag, and the increase of CaO
and MgO content with high melting temperature makes it easy to cause the primary
slag to resolidify and affect operation. Currently, most of the blast furnaces are using
self-fluxing sinter or pellets, and the slag formation reactions between gangue and
flux in the initial stage have been completed in the sintering/pelleting process.
The primary slag moves downward through the gaps between the coke, and the
gas also rises through these gaps. The amount and physical properties of the slag have
an important influence on the distribution of the gas flow. The composition of the slag
changes when it passes the bosh and then to hearth. The position, extension, shape,
and temperature distribution of cohesive zone are essential for gas permeability,
productivity, and smooth operation of the furnace [3, 4]. If the transformation of
solid to liquid is quick and happens in a high and narrow range of temperature, the
cohesive zone would have a better permeability with uniform gas, driving more stable
and high productivity [5].
The slag volume of primary slag is small. Content of FeO and MnO is high due
to their low melting temperature. The volume becomes larger during flowing down
since CaO, SiO2 , MgO, and Al2 O3 increase. On the contrary, the content of FeO and
MnO decreases due to the reduction.
Finally, the slag accumulates in the hearth and forms a slag layer on the top of hot
metal. When the iron droplets pass through the slag layer, many reactions adjust the
composition of the slag and become the final slag, which accumulates to a certain
amount and is periodically tapped out of the furnace.
In the hearth, coke and sulfur from coke and coal ash pass into the slag, and slag
analysis changes to its final composition. It is mainly composed of CaO, SiO2 , MgO,
and Al2 O3 , with a small amount of FeO, MnO, TiO2 , P2 O5, and S. The composition
of a general blast furnace slag is shown in Table 1.1, in which the sum of the content
of SiO2 , CaO, Al2 O3 , and MgO accounts for more than 90%. In particular, for the
smelting of titanium magnetite, the TiO2 content in the slag may exceed 20% [6].

1.3 The Main Physicochemical Properties of Blast Furnace


Slag

1.3.1 Liquids Temperature and Fluidity Temperature

Knowledge of the liquidus temperatures (T liq ) and fluidity temperature (T flu ) of slag
and metal is important to the metallurgists since it is essential to maintain liquid stages
in most pyrometallurgical processes, and disasters can occur if the slag and metal
freeze in the reactors. The liquidus temperature of the slag is the temperature at which
the solid phase of the slag completely disappears during the heating process. The
fluidity temperature refers to the corresponding temperature when the slag meets a
certain fluidity. Metallurgists are more interested in fluidity temperature while paying
8 1 Introduction

more attention to liquidus temperature. Because some slag needs a certain degree
of superheat after being completely melted to meet the smelting requirements of
fluidity.
Commonly used methods to measure the liquidus temperature of slag are DSC and
microscope test, as shown in Figs. 1.4 and 1.5. The DSC apparatus has two pans, one
for holding the sample and the other for holding the reference material. The pans are
heated with a set heating rate in the furnace, and the temperature difference between
these two pans is monitored continuously (Fig. 1.4a). The departure from the baseline
indicates the exothermic or endothermic event (Fig. 1.4b). For the microscope test,
a pressed cylinder of slag powder is monitored as it is heated at a set heating rate.
The temperatures where the sample takes up a set shape are determined; T softening is
the temperature when the cylinder initiates softening (3/4 initial height), T hemisphere is
the hemisphere temperature which is defined as the liquidus temperature (1/2 initial

Fig. 1.4 a Schematic diagram of DSC apparatus and b typical DSC trace for a slag during heating

Fig. 1.5 Schematic drawings showing typical shapes corresponding to (a) initial size, (b) Tsoftening ,
(c) Themisphere , and (d) Tfluidity .
1.3 The Main Physicochemical Properties of Blast Furnace Slag 9

Fig. 1.6 Dependence of viscosity on temperature for slag with different basicity and the fixed slope
method to determine the fluidity temperature

height), T fluidity is the temperature when the slag can fully move (1/4 initial height).
It should be noted that the liquidus temperature measured by DSC tends to differ
from that measured by the microscope test. According to the viscosity measurement,
Seetharaman et al. [7] proposed an activation energy derivation method to estimate the
liquidus temperature of slags. The second derivative offers a useful way of estimating
the liquidus temperatures of multicomponent silicates, which are often difficult to
determine due to supercooling effects.
Since the fluidity temperature is directly related to the viscosity, there are also some
methods to determine the fluidity temperature based on the viscosity-temperature
(η–T ) curve.
• Fixed slope method: the temperature corresponding to the tangent point of the
45° straight line of the η–T curve [8]. For some acid (“long”) slags, the viscosity
obtained from the fixed slope method at the corresponding temperature is too
large to flow, thus losing the essential meaning of fluidity temperature, as shown
in Fig. 1.6.
• Fixed viscosity value method: the temperature corresponding to a fixed viscosity,
normally 2.0–2.5 Pa s [9]. For some slags, the viscosity suddenly changed before
the target value, which means that the fixed viscosity value method cannot
accurately express the viscosity mutations.

1.3.2 Viscosity

The pyrometallurgical processes require the metal and slag to have an appropriate
viscosity. The slag viscosity is important since it relates to the smooth progress of
the smelting process and affects the heat transfer and mass transfer in the process,
10 1 Introduction

thereby affecting the reactions rate and the discharge of impurities in the molten pool,
the slag-metal separation, and the life of the furnace lining. In addition, according to
the relationship between slag viscosity and composition, it helps to understand the
slag structure better.
Viscosity is a measure of the internal friction force when relative movement
occurs between adjacent layers in the fluid during movement, as shown in Fig. 1.7.
According to the definition, there are a pair of parallel plates with an area of A and
a distance of dy, and the plates are filled with fluid. The upper plate exerts a force
F to form a velocity gradient du/ dy, called the shear rate. F/A is called the shear
stress and is represented by τ. The shear rate is proportional to the shear stress, and
its proportional coefficient η is the fluid viscosity (unit: Pa s, N s/m2 ), namely:

F du
τ= =η (1.1)
A dy

For homogeneous slag, its viscosity does not change with changing the shear
stress, called Newtonian fluid. The main factors that determine the viscosity are
composition and temperature. In the metallurgical slag system, the Arrhenius
equation is usually used to describe the dependence of viscosity with temperature.
 

η = A exp (1.2)
RT

here, A, E η , R, and T are a constant, the apparent activation energy of viscous flow,
the universal gas constant, and absolute temperature, respectively.
For heterogeneous multiphase slag, normally due to the presence of insoluble
particles; or the high melting point component precipitates when the temperature
drops. The viscosity of heterogeneous slag is much higher than that of a homoge-
neous slag, which does not follow the Newton’s law and cannot be described by the

Fig. 1.7 Illustration of the definition of fluid viscosity. Since the shearing flow is opposed by
friction between adjacent layers of fluid (which are in relative motion), a force is required to sustain
the motion of the upper plate. The relative strength of this force is a measure of fluid viscosity
1.3 The Main Physicochemical Properties of Blast Furnace Slag 11

Arrhenius equation anymore. The viscosity of this kind of slag is called “apparent
viscosity”.
Blast furnace operation requires the slag to have good stability. Slag stability refers
to the ability of the fluidity temperature and viscosity of the slag to remain stable
with small changes when the temperature and composition fluctuate. For example,
when the temperature of the blast furnace fluctuates within the normal range, the
viscosity does not suddenly change greatly, or the fluidity temperature does not
change significantly. The stable blast furnace slag has a strong ability to resist changes
in raw material properties and temperature fluctuations. It can reduce or avoid the
suspension, collapse, and nodules in the furnace.
Several methods have been used to measure the viscosity of slags at high temper-
atures, including capillary method, falling ball method, cylinder rotation method,
oscillating method, and inclined plane method. Each of these methods is detailed and
well documented in the literatures [10–12]. Figure 1.8 shows the typical laboratory-
scale measurement methods used for slag viscosity measurements. The capillary
method measures relatively low melting temperature liquid/slag viscosities but is not
well suited to high-temperature measurements due to the friction within the capil-
lary walls and the reaction of the melt with the containers materials. The falling ball
method is based on the Stokes’ law that the viscosity can be calculated if the densities
of the sphere ball and slag and the sphere ball falling velocity are known. However,

Fig. 1.8 Viscosity measuring techniques [11, 13]


12 1 Introduction

this method does not provide absolute viscosity measurements. The cylinder rotation
method is widely used for slag viscosity measurements due to its relative simplicity
and reproducibility. The oscillating method has a high equipment accuracy require-
ment and is suitable for measuring the melts with low viscosity (10–5 –10–1 Pa s,
such as metal or alloy melts). The inclined plane method determines the viscosity
via the length of the slag ribbon obtained on an inclined plane. This method is not
very accurate but simple and fast, suitable for industrial conditions.

1.3.3 Density

Density is the mass per unit volume of any object, which is calculated by dividing
the mass of an object by its volume. As one of the important physical properties of
metallurgical slag, the slag density is closely related to the separation between slag
and metal, slag volume, dynamic phenomena in the smelting process, and structure
of slag.
In the blast furnace process, the iron droplets must be settled from the molten
slag to form a molten metal layer immiscible with the molten slag, to separate the
hot metal from the molten slag and reduce the iron loss in the slag. The main factors
affecting the separation and sedimentation are the density difference between the
slag phase and the molten iron, the size of the iron droplets, and the viscosity of the
slag phase. When the iron droplet suspended or immersed in the slag receives more
gravity than the buoyancy, the drop would sink into the slag. The falling speed of the
droplets can be described by the Stokes equation:

2gr 2
v= (ρm − ρs ) (1.3)
9ηs

here, v is the average speed of droplet falling, m/s; r is the radius of the falling iron
droplet, m; ηs is the viscosity of the slag, Pa s; ρ m and ρ s are the density of the metal
and the slag, kg/m3 ; g is the acceleration due to gravity, m/s2 . When the slag viscosity,
droplet radius, and metal density remain constant, the smaller the slag density results
in a larger descending speed of the metal droplet and better slag-iron separation.
In general, the following methods are widely used for the slag density measure-
ments on laboratory-scale: buoyancy method, maximum bubble pressure method,
and droplet method, as shown in Fig. 1.9. The buoyancy method is based on the
Archimedes principle. A bob with known volume is suspended from a balance. The
weight of the bob before immersion and then again after full immersion in the molten
slag is determined. It is necessary to correct the density value for the effect of the
thermal expansion of the bob and the surface tension force acting on the wire. In the
maximum bubble pressure method, the gas pressure in a capillary tube is increased
continuously until a bubble is formed at the mouth of the tube end and subsequently
detaches. The capillary radius is made sufficiently small to ensure that the bubble
has a spherical contour. The maximum pressure occurs when the bubble formed at
1.3 The Main Physicochemical Properties of Blast Furnace Slag 13

Fig. 1.9 Density measuring techniques

the mouth of the capillary is hemispherical. The maximum pressure is determined


at different heights, and the density can be calculated. The volume of a sessile drop
with known mass is determined through measurements of the droplet dimensions.
The improvements in software to describe the profile of the drop have improved the
accuracy of this method.

1.3.4 Surface Properties

Surface tension and interfacial tension are the surface properties, which differ from
other bulk properties. The surface properties play an important role in the interface
reactions of gas-slag-metal system. For example, the Marangoni effect caused by
the surface tension gradient affects the kinetics of metal/slag reactions and the flow
patterns in the vessel. In addition, the slag-metal separation, the removal of inclusions,
the new phase nucleation, and the slag foaming are also affected by the surface
properties. In addition, the surface tension of slag is needed to calculate the interfacial
tension, as shown in Eq. 1.4.
The surface tension can be understood as the energy consumed to generate a
unified interface between the liquid and gas phases. The slag surface is mainly occu-
pied by oxygen ions. Because the radius of the oxygen ions is larger than the radius of
the cations, the surface tension of the slag mainly depends on the interaction between
the surface oxygen ions and the neighboring cations. The oxides with weaker force
fields are enriched on the surface. Therefore, cations with large electrostatic potential
(Z/r) and high ionic bond fraction have higher surface tension, as shown in Fig. 1.10.
The slag with low surface tension and high viscosity is easy to occur foaming
phenomenon. Because the low slag surface tension means less energy is consumed
when bubbles are generated, that is, bubbles are more likely to be formed. The slag
with large viscosity, on the one hand, the foam film is relatively strong; on the other
hand, the bubbles are difficult to escape from the slag, and the generated small bubbles
are not easy to gather and escape. There are many chemical reactions that generate
gas in the blast furnace, and a considerable amount of gas needs to pass through
the slag layer, the generation of bubbles is inevitable. Once the bubbles are stable
14 1 Introduction

Fig. 1.10 The relationship between oxides surface tension and cation electrostatic potential

in the slag resulting in the slag foaming, which would seriously affect the smelting
operation.
The following equation gives the interfacial tension of metal and slag:

γms = γm + γs − 2∅(γm γs )0.5 (1.4)

here, γ m and γ s are the surface tensions of the metal and slag phases, respectively,
and ∅ is an interaction coefficient [14]. The surface tension of the pure iron γ m is
about 1700 mN/m [15], whereas the surface tension of the normal blast furnace slags
(γ s ) system is with the value of about 450 mN/m [11]; thus the value of γ m is almost
four times than that of γ s . The surface tension of metals and slags are very dependent
upon the concentrations of surfactants present. Surfactants tend to be materials with
low surface tension; the surface layer of a liquid tends to have a high concentration
of surfactants. That is why ppm levels of surfactants can significantly affect surface
tension since surfactants concentrate in the surface layer. The principal surfactants
are soluble sulfur and oxygen in metals, whereas the principal surfactants in slags
are B2 O3 , K2 O and Na2 O, and CaF2 .
The available methods to measure the surface tension of slags are the maximum
bubble pressure method, droplet method, and detachment method, as shown in
Fig. 1.11. In the maximum bubble pressure method, the Laplace equation is used
to derive the surface tension. The shape adopted by sessile drop or pendent drop
presents the balance of surface tension and gravitational forces. These forces involve
the surface tension, which can be derived from the dimensions [11]. In the detach-
ment method, a ring or plate is slowly pulled from the below of the slag surface and
a maximum force is measured via a balance when the meniscus is about to break.
The surface tension is calculated by the maximum force.
1.3 The Main Physicochemical Properties of Blast Furnace Slag 15

Fig. 1.11 Surface tension measuring techniques

1.3.5 Sulfide Capacity

Sulfur is a harmful element in steel products, which can reduce the ductility and
toughness of steel, cause cracks during forging and rolling, and is also not conducive
to welding performance and reduces corrosion resistance. In the solid-state, sulfur has
very little solubility in steel and is distributed on the grain boundaries in a network-
like FeS–Fe eutectic form. Since FeS and Fe form a eutectic compound with a low
melting point of 985 °C, the FeS–Fe eutectic at the grain boundary will melt and
cause cracks during the hot process. This phenomenon is called “hot brittleness”,
which is the main hazard of sulfur. Therefore, from raw material pretreatment to
the continuous casting process, each process needs to reduce the sulfur content of
the product as much as possible. Regarding the blast furnace process, the behavior
of sulfur in the blast furnace and the mechanism of the desulfurization reaction are
important subjects to ensure the quality of steel products.
The sulfur in the blast furnace comes from raw materials. With sinter as the main
iron-containing raw material, the sulfur from coke and pulverized coal injection
accounts for about 80% of the total sulfur content. The amount of sulfur introduced
from materials to produce one ton of hot metal is called “sulfur load”. The sulfur
in the charge is mainly burned at the tuyere and goes to gas in the form of H2 S,
CS, CS2 , COS, and SO2 . The SO2 from combustion and decomposition and the S,
and H2 S from the reduction and generation reactions rise with the gas. The upward
gas collides with the downward materials, and part of the surful in gas is absorbed
again. In comparison, most of the sulfur is transferred to the slag during the slag-iron
reaction and is tapped out of the furnace. This ability of slag to contain or dissolve
sulfur is called the sulfide capacity of the slag.
The concept of sulfide capacity (C S ) was proposed by Fincham and Richardson
[16] in 1954, and it was defined as follows:

1   1  
S2 + O2− = O2 + S2− (1.5)
2 2
  21
PO2 aO2−
CS = (wt% S) × = Kθ × (1.6)
PS2 γS2−
16 1 Introduction

here, (wt% S) is the mass percent of sulfur dissolved in slag; PO2 and PS2 are the
partial pressures of oxygen and sulfur, respectively, in the gas phase; K θ is the
equilibrium constant; ao2− is the activity of oxygen in slag; and γS2− is the sulfur ion
activity coefficient in slag.
The following equations represent the equilibrium reaction of sulfur between
metal and slag:
   
[S] + O2− = [O] + S2− (1.7)

(wt% S) a 2− fS
LS = = Kθ × O × (1.8)
[wt% S] γS2− aO

here (wt% S) and [wt% S] is the mass of sulfur dissolved in slag and metal, respec-
tively; K θ stands for the equilibrium constant of Eq. 1.5; ao is the activity of oxygen
in melt; and f S is the sulfur ion activity coefficients in metal. The L S presents the
sulfur distribution in slag and metal. The greater the value of L S , the higher sulfur
content in the slag, that is, the greater the solubility of the slag to sulfur, which has
the same meaning as the sulfide capacity.
The sulfur partition between slag and metal can be correlated with sulfide capacity
by combining Eqs. 1.6 and 1.7 as follows:

1 1
[S] + O2 = [O] + S2 (1.9)
2 2
The equilibrium constant K S for the above reaction is [17]:

936
log K S = − + 1.375 (1.10)
T
Thus, the K s also can be written as:
  21
aO PO2 (wt% S)aO
KS = = (1.11)
aS PS2 [wt% S] f S CS

According to Eq. 1.11, the relationship between L S and CS is obtained as follows:

936
log CS = log L S + log aO − log f S + − 1.375 (1.12)
T
Generally, there are two main methods for measuring the sulfide capacity of slag:
the slag-gas equilibrium method and the slag-metal equilibrium method, as shown
in Fig. 1.12. According to the definition of sulfide capacity (Eq. 1.8), after the slag
undergoes an equilibrium reaction in an atmosphere with a known oxygen and sulfur
partial pressure, the measured sulfur content in the slag can be used to calculate
the sulfur capacity. Normally, CO–CO2 –SO2 –Ar or H2 –CO2 –SO2 –N2 are used for
1.3 The Main Physicochemical Properties of Blast Furnace Slag 17

Fig. 1.12 Oxygen-sulfur equilibrium of blast furnace slag-iron-gas three phase

In the slag-metal equilibrium method, the sulfur content in the slag and metal after
equilibrium and the oxygen activity and sulfur activity coefficient in the metal is
needed.

1.3.6 Electrical Conductivity

The slag has obvious electrical conductivity and can be electrolyzed, indicating the
ionic nature of molten slag. Although the electrical conductivity of slag is not an
important property for blast furnace operation, the investigation of slag electrical
conductivity helps to understand the slag structure. The slag electrical conductivity
obeys Ohm’s law, from which the formula for measuring conductivity can be derived:

C
κ= (1.13)
R
here, κ is the electrical conductivity, S/m; C is the conductivity cell constant,
measured from molten salt with known conductivity; R is the resistance value of
the slag measured between the two electrodes, The electrical conductivity of the slag
is contributed by both ionic conduction (κion ) and electronic conduction (κelec ).
The ionic conduction involves the movement of ions in slag when an electrical
field is applied. Since anions are polymeric with large sizes and do not move readily,
so the electrical conductivity is usually associated with the movement of the cations
present, e.g., Na+ , Ca2+ , etc. The ionic conductivity of the slag is directly proportional
to the migration rate of the ions participating in the conduction. The temperature
dependence is usually represented by the Arrhenius equation:
 
−E κ
κ = A exp (1.14)
RT
18 1 Introduction

Fig. 1.13 Various types of cells for electrical conductivity measurement

here, A, E η , R, and T are a constant, the apparent activation energy of viscous flow,
the universal gas constant, and absolute temperature, respectively.
Electronic conductivity involves a “charge hopping” mechanism and has been
identified in slags containing transition metal oxides, such as FeO, MnO, TiO2 , etc.
[18–20]. Therefore, the slag containing these compounds is a mixed conductor of
ionic-electronic co-conductivity. The increase of temperature reduces the effect of
electron conduction but enhances the effect of ionic conduction.
Many methods have been used to determine the slag’s electrical conductivity,
which is well described by Mills [11]. These various types of cells are shown in
Fig. 1.13. Polarization is the main problem encountered, which is caused by the
concentration gradient generated in the electrolyte, which manifests as a back elec-
tromagnetic force opposite the normal ion flow. The four-electrodes method is used
to minimize polarization, where two electrodes carry current, and the other two
electrodes monitor potential [21, 22].

1.4 Blast Furnace Ironmaking Requirements for Slag


Properties

In the blast furnace process, to obtain metallic iron from iron oxide materials, it
is necessary to realize the chemical separation of Fe and O via reduction and the
physical separation of hot metal and oxide gangue. The latter is achieved by using the
difference in density between slag and hot metal. Therefore, the smooth of the blast
furnace practice requires blast furnace slag to have suitable physical and chemical
properties, including:
• Suitable Liquids temperature. Generally, the hearth temperature of the blast
furnace is about 1450 °C. When the hot metal is saturated with carbon, the liquidus
temperature is about 1150 °C. Therefore, the melting point of the slag must be
lower than 1450 °C with a certain degree of superheat to ensure that the slag and
iron are in a completely molten state.
References 19

• Suitable fluidity. The blast furnace process requires the slag to have an appropriate
viscosity. Low slag viscosity of primary slag is beneficial to maintain sufficient
gas permeability and drainage through the coke grid. Low viscosity of final slag
is necessary to separate slag and hot metal as well as free running during tapping.
However, if the viscosity is too low, resulting in the corrosion of refractory and
reduces the life of the hearth lining. Therefore, the slag viscosity is required to be
within a certain range in the blast furnace operation, that is, 1–6 dPa s.
• Ability to participate in desired chemical reactions. Such as reducing silicon,
manganese, or other beneficial elements, and the absorption/removal of harmful
elements such as sulfur and alkali metals oxides. Hot metal desulphurization is one
of the most important functions of blast furnace slag. Slag composition, fluidity,
and slag volume affect the desulfurization performance.
• Suitable kinetic conditions. Gas and metal need to pass through the slag layer
smoothly and chemical reactions can be controlled by the slag, which required
the slag to meet the necessary kinetic conditions.
• Stability. The performance of the slag does not deteriorate rapidly due to sudden
changes in smelting conditions. When the composition or temperature suddenly
changes, the physicochemical properties of the slag do not change significantly
to ensure the smooth progress of smelting.
Slag should also play an important role in protecting the furnace lining. Appro-
priate slag physicochemical properties can reduce the erosion of the furnace lining,
and the reasonable composition can install a protective shield for the furnace lining.
Recently, the titania-containing slag (1–3 wt% TiO2) protection technology has been
promoted in the industry. Due to the reduction of TiO2 , the molten iron contains a
certain amount of [Ti]. Ti, C, and N produce high melting point compounds TiC and
Ti(C, N) (melting temperature more than 2000 °C), and its solubility in hot metal
decreases as the temperature decreases. Therefore, in the areas severely eroded in
the hearth, the metal temperature decreases due to the increase in the relative cooling
strength, which leads to the precipitation and deposition of TiC and Ti (C, N), thereby
automatically repairing the refractory of the furnace [23, 24].

References

1. Energy use in the steel industry (2020). World Steel Association. https://www.worldsteel.org/
publications/fact-sheets.html
2. Lüngen, H. B., & Schmöle, P. (2020). Comparison of blast furnace operation modes in the
world. Steel Research International, 91(11), 2000182.
3. Gudenau, H.-W., Kreibich, K., & Nomiya, Y. (1980). Influence of zones of different perme-
ability on the gas flow in the blast furnace with special consideration of the cohesive zones.
Stahl und Eisen, 100(25), 1526–1534.
4. Gudenau, H., Kreibich, K., & Nomiya, Y. (1980). Model investigation into the gas permeability
in the blast furnace with and without a cohesive zone. Stahl und Eisen, 100(9), 488–494.
5. Babich, A., Senk, D., & Gudenau, H. W. (2016). Ironmaking & Steelmaking, 43(1), 11-21
20 1 Introduction

6. Pang, Z., Lv, X., Jiang, Y., Ling, J., & Yan, Z. (2019). Blast furnace ironmaking process with
super-high TiO2 in the slag: Viscosity and melting properties of the slag. Metallurgical and
Materials Transactions B, 51, 722-731.
7. Seetharaman, S., Du, S., Sridhar, S., & Mills, K. C. (2000). Estimation of liquidus tempera-
tures for multicomponent silicates from activation energies for viscous flow. Metallurgical &
Materials Transactions B, 31(1), 111–119.
8. Lv, X., Yan, Z., Liang, D., Zhang, J., & Bai, C. (2016). Transition of blast furnace slag from
silicates-based to aluminates-based: Viscosity. Metallurgical and Materials Transactions B,
48(2), 1092–1099.
9. Wang, X. (2014). Iron and steel metallurgy (Part of ironmaking). Metallurgical Industry Press.
10. Sohn, I., Wang, W., Matsuura, H., Tsukihashi, F., & Min, D. J. (2012). Influence of TiO2 on the
viscous behavior of calcium silicate melts containing 17 mass% Al2 O3 and 10 mass% MgO.
ISIJ International, 52(1), 158–160.
11. M Allibert; Verein Deutscher Eisenhüttenleute.; et al (1995). Slag atlas. edited by Verein
Deutscher Eisenhüttenleute (VDEh)
12. Khanna, R., & Sahajwalla, V. (2013). Treatise on process metallurgy. Volume 1: Process
fundamentals. Elsevier.
13. Brooks, R. F., Dinsdale, A. T., & Quested, P. N. (2005). The measurement of viscosity of
alloys—A review of methods, data and models. Measurement Science and Technology, 16(2),
354–362.
14. Chung, Y., & Cramb, A. (2000). Dynamic and equilibrium interfacial phenomena in liquid
steel-slag systems. Metallurgical and Materials Transactions B, 31(5), 957–971.
15. Kawai, Y., Mori, K., Kishimoto, M., Ishikura, K., & Shimada, T. (1974). Surface tension of
liquid Fe-C-Si alloys. Tetsu-to-Hagané, 60(1), 29–37.
16. Fincham, C., & Richardson, F. D. (1954). The behaviour of sulphur in silicate and aluminate
melts. Proceedings of the Royal Society of London. Series A. Mathematical and Physical
Sciences, 223(1152), 40–62.
17. Hatch, G. G., & Chipman, J. (1949). Sulphur equilibria between iron blast furnace slags and
metal. JOM Journal of the Minerals Metals and Materials Society, 1(4), 274–284.
18. Barati, M., & Coley, K. S. (2006). Electrical and electronic conductivity of CaO-SiO2 -FeOx
slags at various oxygen potentials: Part I. Experimental results. Metallurgical and Materials
Transactions B, 37(1), 41–49.
19. Barati, M., & Coley, K. S. (2006). Electrical and electronic conductivity of CaO-SiO2 -FeOx
slags at various oxygen potentials: Part II. Mechanism and a model of electronic conduction.
Metallurgical and Materials Transactions B, 37(1), 51–60.
20. Hu, K., Lv, X., Yu, W., Yan, Z., Lv, W., & Li, S. (2019). Electric conductivity of TiO2 -
Ti2 O3 -FeO-CaO-SiO2 -MgO-Al2 O3 for high-titania slag smelting process. Metallurgical and
Materials Transactions B, 50(6), 2982-2992.
21. Mitchell, A., Joshi, S., & Cameron, J. (1971). Electrode temperature gradients in the electroslag
process. Metallurgical Transactions, 2(2), 561–567.
22. Mitchell, A., & Cameron, J. (1971). The electrical conductivity of some liquids in the system
CaF2+CaO+Al2 O3 . Metallurgical and Materials Transactions B, 2(12), 3361–3366.
23. Li, Y., & Fruehan, R. J. (2002). Study on the mechanism of blast furnace titanium ore shielding.
Baosteel Technology (1), 12–16.
24. Zhao, Y. (2014). The impact of Titanium on Skull Formation in the Blast Furnace Hearth. In
AISTech-2013, Pittsburgh, Pa, USA, 95–103.
Chapter 2
Phase Diagram and Equilibrium

The blast furnace slag can be generally considered to be mainly composed of four
oxides, namely SiO2 , CaO, Al2 O3, and MgO. According to the combination principle
of acid oxides and basic oxides, they can form a wide variety of complex compounds.
Through the phase diagram of binary system and multivariate system formed by them,
the composition, structure, characteristics of precipitation during solidification, and
conditions of phase equilibrium coexistence of these composite compounds can be
understood, which provides a necessary basis for the study of the structure and
properties of slag [1].

2.1 CaO–SiO2

The phase diagram of CaO–SiO2 system is shown in Fig. 2.1, which is more complex
due to the formation of several calcium silicates with different properties (stable or
unstable) and the occurrence of polycrystalline transformation.
This system has two congruent melting compounds, namely calcium metasilicate
CaO · SiO2 (CS) and calcium orthosilicate 2CaO · SiO2 (C2 S). And there are two
incongruent melting compounds, namely tricalcium silicate 3CaO · SiO2 (C3 S) and
tricalcium disilicate 3CaO · 2SiO2 (C3 S2 ). In the figure, SiO2 , CS, and C2 S all have
polycrystalline transformation phenomena, marked with 575, 867, 1125, 1465, 847,
1437, and 1125 horizontal lines are the isotherm of their crystal transformation,
the numbers marked on the line are their change temperature (°C). There is also
a two-liquid phase area [2], on the rich SiO2 side, its liquid phase has immiscible
phenomenon. There are also a large number of variableless points in the figure, the
nature of which is shown in Table 2.1.
For more complex binary phase diagrams, the system can be divided into several
subsystems by congruent melting compounds. For the CaO–SiO2 system, the phase
diagram can be divided into CaO–C2 S system, C2 S–CS system, and CS–SiO2 system,
with congruent compounds as the phase boundary.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 21
X. Lv and Z. Yan, High Temperature Physicochemical Properties of High Alumina Blast
Furnace Slag, https://doi.org/10.1007/978-981-19-3288-5_2
22 2 Phase Diagram and Equilibrium

Fig. 2.1 The phase diagram of CaO–SiO2 system

CaO–C2 S system is a phase diagram with a eutectic reaction, but there is only a
stable C3 S within the range of 1300–1799 °C in the solid phase. Above or below the
temperature range, C3 S cannot exist and would decompose into CaO and C2 S. The
vertical line in the phase diagram represents C3 S.
The C2 S–CS system is a phase diagram with an unstable compound (C3 S2 ). When
the temperature decreases, C3 S2 is formed by peritectic reaction: L + C2 S → C3 S2 .
When heated, it decomposes at 1475 °C: C3 S2 → L + C2 S.
Because of the limited solubility of the components in the liquid phase of CS–
SiO2 system, the phase diagram of the coexistence of two immiscible liquid phases
is formed. This phenomenon is common in SiO2 -alkaline metal oxide binary phase
diagram. In these two mutually saturated liquid phases, Aslag-liq is the saturated
melt of SiO2 in CS phase, and Aslag-liq#2 is the saturated melt of CS in SiO2 phase.
When the temperature is above 1689 °C, both of them coexist in equilibrium, and
their equilibrium components are shown by two lines, respectively, which are called
dissolution curves. At 1689 °C, the phase equilibrium is Aslag-liq#2 → Aslag-liq
+ SiO2 (monotectic reaction). When the temperature is higher than 1689 °C, SiO2
gradually disappears and only two liquid phases coexist. Their saturation solubility
changes continuously with the increase of temperature and gradually approaches,
finally reaching the same value and becoming homogeneous liquid phase. The point
2.1 CaO–SiO2 23

Table 2.1 Properties of variableless points in phase diagram of CaO–SiO2 system


Punctuation Phase equilibrium Nature Composition/mol Temperature/°C
in the figure relationship CaO SiO2
A Aslag-liq → CaO + Eutectic 0.704 0.296 2017
C2 S(s3) reaction
B CaO + C2 S(s3) → C3 S Peritectoid 0.751 0.249 1799
reaction
C C3 S → CaO + C2 S(s2) Eutectoid 0.751 0.249 1300
reaction
D C2 S(s3) → Aslag-liq Congruent 0.669 0.331 2154
melting
E C2 S(s3) → C2 S(s2) Crystal 0.669 0.331 1437
transformation
F C2 S(s2) → C2 S(s) Crystal 0.669 0.331 847
transformation
G C3 S2 → C2 S(s3) + Incongruent 0.601 0.399 1464
Aslag-liq melting
H Aslag-liq → C3 S2 + Eutectic 0.577 0.423 1463
CS(s2) reaction
I CS(s2) → Aslag-liq Congruent 0.502 0.498 1540
melting
J CS(s2) → CS Crystal 0.502 0.498 1125
transformation
K Aslag-liq → SiO2 (s4) + Eutectic 0.387 0.613 1437
CS(s2) reaction
L Aslag-liq#2 → Aslag-liq Monotectic 0.282 0.718 1689
+ SiO2 reaction
M Aslag-liq#2 → Aslag-liq Monotectic 0.012 0.988 1689
+ SiO2 reaction
N SiO2 (s6) → SiO2 (s4) Crystal 0 1 1465
transformation
O SiO2 (s4) → SiO2 (s2) Crystal 0 1 867
transformation
P SiO2 (s2) → SiO2 (s1) Crystal 0 1 575
transformation

on the curve that reaches the same value is called the critical point. When the tempera-
ture is lower than 1689 °C, Aslag-liq#2 disappears, but Aslag-liq exists, and SiO2 will
be continuously precipitated as the temperature drops. When Aslag-liq was cooled
to 1437 °C, CS–SiO2 eutectic formed.
The pressure–temperature system phase diagram and crystal transformation rela-
tionship [3] of SiO2 are shown in Fig. 2.2. It should be noted that the vapor pressure of
SiO2 is very small (about 10–12 Pa at 1727 °C). The pressure coordinate in the figure
does not represent the actual pressure value, and the pressure curve only represents
24 2 Phase Diagram and Equilibrium

Fig. 2.2 Pressure–temperature system phase diagram and crystal transformation relationship of
SiO2

the trend of pressure change when the temperature changes. In addition, the dotted
line in the figure represents the equilibrium state of metastable crystal form, and the
solid line represents the equilibrium state of stable crystal form. The changes fall into
two kinds: The first type (horizontal) is the transformation of SiO2 (s2) that known
as α-quartz (hexagonal biconical)  SiO2 (s4) that known as α-tridymite (hexag-
onal system plate)  SiO2 (s6) that known as α-cristobalite (cubic octahedron) is
different in the way of bonding between any two SiO tetrahedrons in their lattice, so
the transition between them is slow, occurring under slow heating or cooling condi-
tions. The second type is longitudinal and is the transformation of the subspecies of
the above three crystal forms with the same crystal structure, namely α-tridymite,
β-tridymite, and γ-tridymite (SiO2 -s3,s3.5,s4). When the SiO2 is transformed from
the high-temperature type to the low-temperature type, only the positions of atoms in
the lattice and the connection angles between the tetrahedrons change, which occurs
under the condition of rapid heating or cooling. The phase diagram was measured
under slow heating or cooling conditions, therefore there is no subspecies change in
the crystal types of SiO2 .
In the three types of crystal transformation of SiO2 , the volume change will occur.
Figure 2.3 shows the percentage of volume change on heating for each type of crystal
and its subspecies. The SiO2 content of silicon brick is very high, so before use, slow
heating and baking should be carried out under 800 °C, in order to eliminate the
abrupt change of volume, cluster, and avoid cracking in use.
2.1 CaO–SiO2 25

Fig. 2.3 Percentage of thermal expansion in quartz crystal transformation

There are two crystal forms of CS, namely pseudowollastonite (α-CS) and β-CS.
The latter is converted to α-CS of the isomelting compound at 1125 °C (melting
point is 1540).
The crystal transformation of C2 S is shown as Fig. 2.4.
C2 S has four crystal forms: α, α , β, γ(Ca2 SiO2 -s,s1.5,s2,s3). Among them, α -C2 S
has a subspecies of β-C2 S and can be reversibly and rapidly converted to β-C2 S at
675 °C. When α -C2 S → γ-C2 S occurs, the density decreases from 3.28 × 103 kg/m3
to 2.97 × 103 kg/m3 , so the volume increases by about 10%. This transformation
leads to the powdering phenomenon of poorly calcined cement frit, alkaline silicate
slag, poorly prepared dolomite refractories, and high alkalinity sinter. βC2 S has good
hydrorigidity and is a beneficial component of cement. γ-C2 S is almost anhydrous
hardness, so it is a harmful component of cement. Therefore, when cement clinker is
sintered, drastic cooling measures should be adopted to preserve α -C2 S or change
it into β-C2 S. In addition, adding W (P2 O5 ) ≈ 1% P2 O5 (calcium phosphate, B2 O3 ,
Cr2 O3 , etc.) which forms a solid solution with SiO2 can stabilize the lattice of α -C2 S.

Fig. 2.4 The crystal transformation of C2S


26 2 Phase Diagram and Equilibrium

Thus the α → β transition temperature is reduced by dozens of degrees, and the β


→ γ transition is prevented.
In addition, C2 S can form solid solutions with other silicates, such as CaO–MnO–
SiO2 , CaO–FeO–SiO2, or 2FeO–SiO2 , resulting in easy soluble of C2 S.

2.2 CaO–Al2 O3

In the CaO–Al2 O3 system, the alkaline earth CaO can form a series of complex
compounds with the gender oxide Al2 O3 , with a total of five compounds (see Fig. 2.5):
C3 A (Ca3 Al2 O6 ), C12 A7 (Ca12 Al14 O3 ), CA (CaAl2 O4 ), CA2 (CaAl4 O7 ) and CA6
(CaAl12 O19 ). C3 A and CA6 are incongruent melting compounds, and the rest are
congruent melting compounds. However, according to the information [2, 4, 5],
C12 A7 is a congruent melt compound in the usual humidity air, when C3 A and
CA are found to form a low comet if in a completely dry atmosphere at 1360 °C,
composed of 36% (mol) Al2 O3 , 64% (mol) CaO. Therefore, at this time the congruent
melt compound C12 A7 , in the phase diagram there is no stable phase region of it, as
shown in the blue bright area in the figure. Almost all the compounds in the phase
diagram are important phases in silicate cement and aluminate cement. In addition to

Fig. 2.5 The phase diagram of CaO–Al2 O3 system


2.2 CaO–Al2 O3 27

Fig. 2.6 CA hydration reactions

C12 A7 , these compounds have high melting points and decomposition temperatures.
When synthesizing slag is selected, the composition of the system can appear in the
liquid phase zone at a temperature of 1450–1550 °C when the composition of the
system is mol (CaO) of 60–66.4%.
Tricalcium aluminate (C3 A) is an important mineral in silicate cement and may
also be found in dolomite refractory materials, which are heated to 1539 °C and
decomposes into CaO and liquid phases: C3 A → CaO-Aslag-liq. C3 A reacts strongly
with water, and when C3 A is more in silicate cement, the cement hardens fast.
Calcium aluminate (CA) is the main mineral component of high aluminum
cement, aluminum-60 cement, and alumina cement. Generally, colorless column or
long striped crystals, belonging to a single oblique crystal system, density 2.98 g/cm3 ,
hardness 6.5, melting point 1600 °C. When meeting with water, hydration reactions
occur and harden, but the hydration product varies from hydration temperature to
hydration temperature, as shown in Fig. 2.6. Among hydration products, CAH10
(CA · 10H2 O) is called hydrated calcium aluminate, needle or plate-shaped crys-
tallization, with high strength, C2 AH8 (C2 A · 8H2 O) is called hydrated dicalcium
aluminate, and the crystallization state is similar to hydrated calcium aluminate,
C3 AH6 (C3 A · 6H2 O) is called hydrated tricalcium aluminate, with octahedral or
cubic crystallization, low strength, AH3 (Al2 O3 · 3H2 O) is called aluminum glue,
in granular form.
Calcium dialuminate (CA2 ) is the main mineral composition of aluminum-70
cement but also aluminum-60 cement, alumina cement, and high aluminum cement
mineral composition, it is a column or needle-shaped colorless crystal, is a quadrantal
crystal system, density of 2.91 g/cm3 , melting point 1750 °C. Compared with CA,
CA2 has high melting point, slow hydration speed, low early strength, and high
later strength. The hydration reaction is similar to CA, but in the hydration product,
aluminum glue AH3 is more. In the above phase diagram, the CaO–Al2 O3 system
would undergo eutectic reaction at 52.5 mol% Al2 O3 at 1597 °C: Aslag-liq → CA
+ CA2 .
Dodecalcium heptaaluminate (C12 A7 ) is a secondary mineral component in alumi-
nate cement. Belongs to the cubic crystal system, round grain or octa-surface crystal-
lization, density 2.69 g/cm3 , melting point 1455 °C, it has the characteristics of rapid
coagulation when encountering water, and its content in cement should not be too
much in electromelt alumina cement, often contains a certain amount of C12 A7 , so the
condensation speed is fast, and it needs to be used after the addition of anticoagulant.
28 2 Phase Diagram and Equilibrium

In earlier studies, C12 A7 was mistakenly regarded as the molecular formula of


C5 A3 (Ca5 Al6 O14 ), known as the “stable C5 A3 ”. Later, it is identified from another
phase composition, is called unstable C5 A3 . There can be two paths in the synthesis
of dodecalcium heptaaluminate [6]. For the low-temperature path: the presence of
intermediate phase C5 A3 exists: 7Ca5 Al6 O14 + CaO → 3Ca12 Al14 O33 . For the high-
temperature path: directly synthesized from C3 A and CA. It is worth noting that in
addition to its role in cement, the dodecalcium heptaaluminate materials are applied in
the fields of electricity, optics, and chemistry with their novel conductivity, oxidation,
and storage emissivity. In addition, C12 A7 also has a series of other characteristics,
such as superconductivity, catalysis, antibacterial, etc. Therefore, many experts and
scholars have conducted in-depth research on its preparation methods and existing
characteristics [4–10].
Calcium hexaaluminate (CA6 ), which exists in aluminate cement, is generally
hexagonal plate-shaped crystal. It does not melt uniformly at 1850 °C to form
corundum and liquid phase: CA6 → Al2 O3 + Aslag-liq. The hydraulic property
of CAS is weak and almost has no gelling property. The chemical composition and
melting point of various calcium aluminate minerals are listed in Table 2.2. Gener-
ally, the melting point increases with the increase of Al2 O3 content, but the gelling
property decreases, in calcium aluminate minerals.
In the phase diagram of CaO–Al2 O3 system, in addition to providing with the
composition and properties of various aluminate gels, another important aspect is
its two end elements. For CaO, a little absorption of Al2 O3 forms low melting point
compound C3 A, and the liquid phase temperature of the system decreases from 2570
to 1535 °C and 1035 °C. It can be seen that Al2 O3 is a strong flux for CaO. Therefore,
Al2 O3 is regarded as a harmful impurity in calcareous and dolomitic refractories with
CaO as the main component, which should be especially limited. On the contrary,
Al2 O3 absorbs CaO to form high melting point compound CAS (decomposition at
1850 °C), and the liquid phase temperature of the system decreases from 2050 to
1850 °C. CA6 can form after the slag is absorbed by the high alumina brick on the top

Table 2.2 Chemical compositions and melting point of calcium aluminate minerals
Name Chemical formula Composition/mol Melting point/°C
CaO Al2 O3
Tricalcium C3A 0.751 0.249 1535 (decomposition)
aluminate
Dodecalcium C12A7 0.625 0.375 1455
heptaaluminate
Calcium CA 0.500 0.500 1600
aluminate
Dicalcium CA2 0.334 0.666 1750
aluminate
Calcium CA6 0.143 0.857 1850 (decomposition)
hexaaluminate
2.3 Al2 O3 –SiO2 29

of the electric furnace, indicating that Al2 O3 has good stability with CaO. This is also
one of the important reasons why high alumina bricks have occupied the position of
electric furnace top for a long time.

2.3 Al2 O3 –SiO2

The phase diagram of Al2 O3 –SiO2 system is shown in Fig. 2.7. It is relatively simple.
The melting points of binary Al2 O3 and SiO2 are 2054 °C and 1723 °C, respectively.
There is a crystalline transformation of SiO2 in the system. The relevant properties
of SiO2 have been introduced earlier and do not be repeated here.
The amphoteric oxide Al2 O3 shows basic properties when coexisting with acidic
oxide, so it can form compound 3Al2 O3 · 2SiO2 (A3 S2 ) with strongly acidic oxide
SiO2 , which is called mullite. As shown in the brown part of the figure, mullite is
a uniform melt, a solid solution with a small amount of Al2 O3 dissolved, and has a
certain melting point (1890 °C). Mullite is found in the volcanic eruptive rocks of
Moore Island, a small island in western Scotland. It is named after its “Hometown”.
There are few mullites in nature. Mullite is a chain silicate, acicular or columnar
crystal, melting point of 1850 °C, small coefficient of thermal expansion, average of

Fig. 2.7 The phase diagram of Al2 O3 –SiO2 system


30 2 Phase Diagram and Equilibrium

5.3 × 10–6 °C−1 from room temperature to 1000 °C and has strong thermal shock
resistance. People describe it as the excellent silicate of ceramists, which may be
due to its rare and excellent properties in nature. Theoretical composition of mullite:
71.8 wt% Al2 O3 , 28.2 wt% SiO2 . Equivalent to A/S mass ratio of 2.55. In fact, mullite
can dissolve in excess Al2 O3 in the structure, up to about 77.3 wt% Al2 O3 , composed
of A2 S, so its composition fluctuates in the range of A3 S2 –A2 S. It is mullite solid
solution (simplified as A3S2). Mullite composed in the range of solid solution does
not appear liquid phase below 1890 °C.
Due to the existence of mullite, Al2 O3 –SiO2 binary system is divided into A3 S2 –
Al2 O3 and two subsystems. Mullite becomes an important dividing line for the perfor-
mance difference in the system. With the increase of SiO2 content, the melting point
of the system decreases continuously until the lowest point is 1594 °C, and the lowest
point of the system melting point is reached when mole (SiO2 ) is 0.957. The reverse
eutectic reaction is Aslag-liq → mullite + SiO2 (S6). In the solid–liquid two-phase
region on the left side of the eutectic point, the amount of mullite decreases with the
increase of SiO2 content. Although the phase combination does not change, there
is still a large difference in performance due to a wide range. Thus, it is one of the
problems that should be paid attention to in the application of this phase diagram.
The eutectic temperature of A3 S2 –Al2 O3 subsystem is 1890 °C, the eutectic compo-
sition point is close to A3 S2 side, mole (SiO2 ) = 0.325, and the coexisting solid
phase are two high melting point phases of mullite and corundum, which indicates
that when the Al2 O3 content or A/S ratio of Al2 O3 –SiO2 system material exceeds the
theoretical composition of mullite, the liquid phase temperature of the system begins
to appear, which is increased by nearly 250 °C, and the properties of the material
change significantly.
The melting point of SiO2 in the figure is 1723 °C. When Al2 O3 is added, the
liquid phase temperature drops sharply to 1595 °C. The lowest eutectic composition
point is very close to SiO2 , and the molar content of Al2 O3 is 0.043, which makes
the liquefaction line of SiO2 + Aslag-liq become very steep. If 1% Al2 O3 is mixed
with SiO2 , it can be calculated that about 18.2% liquid phase is produced by using
the lever rule, and the amount of liquid phase increase sharply with the increase
of temperature. On the contrary, if 1% SiO2 is introduced with Al2 O3 , the liquid
phase appears at 1890 °C. Similarly, it can be calculated by the lever principle that
the amount of liquid phase is only 4.8%, indicating that the flux effect of SiO2 on
Al2 O3 is not strong. Based on this idea, introducing a small amount of SiO2 micro
powder into corundum (a-Al2 O3 ) castable can not only improve the workability and
medium temperature strength of the material but also promote the early sintering of
the material by reacting with mullite of Al2 O3 . It is estimated that even if SiO2 enters
the liquid phase at the service temperature (1650–1750 °C), it increases its viscosity
and changes the liquid phase properties of Al2 O3 -rich area.
2.4 CaO–SiO2 –Al2 O3 31

2.4 CaO–SiO2 –Al2 O3

There are 15 compounds in the whole CaO–Al2 O3 –SiO2 system (when the crystal
transformation is not considered), including three pure components and their melting
points: CaO (2572 °C), Al2 O3 (2054 °C), SiO2 (1723 °C); 10 binary compounds,
including six congruent melting compounds. Their melting points are CS (1540 °C),
C2 S (2154 °C), C12 A7 (1455 °C), CA (1604 °C), CA2 (1765 °C), A3 S2 (1890 °C);
Four incongruent melting compounds and their decomposition temperatures are
C3 S2 (1464 °C), C3 S (1799 °C), C3 A (1539 °C) and CA6 (1833 °C). There are two
ternary compounds in the phase diagram: anorthite CAS2 (CaAl2 Si2 O8 , 1553 °C)
and aluminum andalusite C2 AS (Ca2 Al2 SiO7 , 1590 °C).
Figure 2.8 is the CaO–Al2 O3 –SiO2 ternary phase diagram based on the phase
diagram data reported by Osborm and Muan [11]. Considering that there is no stable
C12 A7 in the dry environment, there is no C12 A7 in the diagram. There are 20 nonva-
riable points in the phase diagram. The diagram is divided into 20 triangles. The
specific characteristics of nonvariable points are shown in Table 2.3. The temper-
ature range of this phase diagram is 1184–2572 °C, and the isotherms of different
liquidus are represented by different color lines and the temperature increases from
inside to outside. It can be seen from the phase diagram that the liquid phase region of
the system increases gradually with the increase in temperature. This phase diagram
plays an important role in the phase transformation theory and industry of silicate.
Figure 2.9a shows the approximate composition range of various silicate materials
in this ternary phase diagram. Using this phase diagram, it can be determined that

Fig. 2.8 The phase diagram of CaO–Al2 O3 –SiO2 system


32 2 Phase Diagram and Equilibrium

Table 2.3 Properties of variableless points in phase diagram of CaO–SiO2 –Al2 O3 system
Num Corresponding triangles Composition/mol Temperature/°C
SiO2 CaO Al2 O3
1 Ca2 SiO4 (s3)–Ca3 SiO5 –CaO 0.20 0.70 0.10 1727
2 Al2 O3 (s4)–CaAl2 Si2 O8 (s2)–Mullite 0.48 0.22 0.30 1546
3 Ca2 Al2 SiO7 –CaAl2 O4 –CaAl4 O7 0.08 0.49 0.43 1529
4 Ca2 Al2 SiO7 –Ca2 SiO4 (s3)–CaAl2 O4 0.14 0.59 0.27 1474
5 Mullite–SiO2 (s6)–SiO2 (s4) 0.81 0.07 0.12 1465
6 Ca3 Al2 O6 –Ca3 SiO5 –CaO 0.10 0.70 0.20 1447
7 Al2 O3 (s4)–CaAl12 O19 –CaAl2 Si2 O8 (s2) 0.37 0.34 0.28 1447
8 Ca2 Al2SiO7 –CaAl12 O19 –CaAl4 O7 0.33 0.39 0.28 1443
9 Ca2 SiO4 (s3)–Ca3 Al2 O6 –Ca3 SiO5 0.11 0.70 0.19 1441
10 Ca2 SiO4 (s3)–Ca2 SiO4 (s2)–Ca3 Al2 O6 0.10 0.70 0.20 1437
11 Ca2 SiO4 (s3)–Ca2 SiO4 (s2)–CaAl2 O4 0.08 0.61 0.31 1437
12 Ca2 SiO4 (s3)–Ca2 SiO4 (s2)–Ca3 Si2 O7 0.42 0.56 0.02 1437
13 Ca2 Al2 SiO7 (s)–Ca2 SiO4 (s3)–Ca2 SiO4 (s2) 0.37 0.53 0.10 1437
14 Ca2 Al2 SiO7 –CaAl12 O19 –CaAl2 Si2 O8 (s2) 0.37 0.38 0.25 1392
15 CaAl2 Si2 O8 (s2)–Mullite–SiO2 (s4) 0.75 0.12 0.13 1367
16 Ca2 Al2 SiO7 –Ca2 SiO4 (s2)–Ca3 Si2 O7 0.39 0.52 0.08 1333
17 Ca2 SiO4 (s2)–Ca3 Al2 O6 –CaAl2 O4 0.04 0.65 0.31 1332
18 Ca2 Al2 SiO7 (s)–Ca3 Si2 O7 –CaSiO3 (s2) 0.42 0.48 0.10 1273
19 Ca2 Al2 SiO7 –CaAl2 Si2 O8 (s2)–CaSiO3 (s2) 0.44 0.45 0.11 1257
20 CaAl2 Si2 O8 (s2)–CaSiO3 (2)–SiO2 (s4) 0.64 0.28 0.08 1184

the preparation conditions of various silicate materials, select the firing and melting
temperature, and understand the changes and properties of materials in the cooling
process, so as to obtain materials with required properties. In metallurgy, the main
components of blast furnace slag, some ferroalloy smelting slag, cast steel protective
slag, and out of furnace refining slag can be attributed to this slag system. Figure 2.9b
shows the local phase diagram within the composition range of blast furnace slag,
which is determined according to the smelting conditions of blast furnace (basicity,
viscosity, and temperature). It can be seen from the figure that the melting temper-
ature in this phase diagram is not high, especially near the variable free points 16,
18, and 19. This local phase diagram can be used to adjust the composition of blast
furnace slag and control the smelting conditions so that the slag can obtain the appro-
priate melting temperature under the guarantee of various technical and economic
indexes, so as to facilitate the smooth progress of smelting. Obviously, the melting
point of the actual blast furnace slag is lower than that of the determined in the phase
diagram, because there are a small amount of other oxides (such as MnO, MgO, etc.)
that can reduce the melting point in the blast furnace slag.
2.4 CaO–SiO2 –Al2 O3 33

Fig. 2.9 a Approximate composition range of various silicate materials and b local phase diagram
of composition range of blast furnace slag in CaO–Al2 O3 –SiO2 ternary phase diagram. 1. Glass, 2.
BF slag, 3. Silicate cement, 4. Refractory, 5. Ceramic, 6. High aluminum brick, mullite, corundum
34 2 Phase Diagram and Equilibrium

Table 2.4 and Fig. 2.10 show that the CaO–Al2 O3 –SiO2 system has only one
(14) single-phase region of liquid phase at 1600 °C, accounting for about 1/2 of the
whole area. There are eight two-phase regions of one solid phase and one liquid
phase, which are 1, 3, 5, 6, 7, 9, 11, and 13. The straight lines introduced in these
phase regions are called connecting lines, and the two ends of the line represent the
composition of solid–liquid two phases balanced by any composition on the line at
1600 °C, The number of liquid–solid two phases can be determined by lever rules
depending on the position of composition points of components. The liquid phase
composition of components in the same phase area and different connecting lines

Table 2.4 Phase diagram of CaO–Al2 O3 –SiO2 system equilibrium phase in 1600 °C equilibrium
phase region
Phase Equilibrium phase Phase region Equilibrium phase
region
1 CaO + Aslag-liq 8 Mullite + Al2 O3 + Aslag-liq
2 CaO + Ca3 SiO5 + Aslag-liq 9 Al2 O3 + Aslag-liq
3 Ca3 SiO5 + Aslag-liq 10 Al2 O3 + CaAl12 O19 + Aslag-liq
4 Ca3 SiO5 + Ca2 SiO4 (s3) + 11 CaAl12 O19 + Aslag-liq
Aslag-liq
5 Ca2 SiO4 (S3) + Aslag-liq 12 CaAl12 O19 + CaAl4 O7 + Aslag-liq
6 SiO2 (s6) + Aslag-liq 13 CaAl4 O7 + Aslag-liq
7 Mullite + Aslag-liq 14 Aslag-liq

Fig. 2.10 Phase diagram of CaO–Al2 O3 –SiO2 system isothermal section at 1600 °C
2.5 CaO–SiO2 –Al2 O3 –MgO 35

is different, but the solid phase composition is the same. The remaining 2.4.8.10.12
are two solid and one liquid three-phase areas. The three vertices of the triangle
represent the three equilibrium phases of the composition in the three-phase area
at 1600 °C. The quantitative relationship of the equilibrium phases can be obtained
by the parallel line rule or lever rule depending on the position of the composition
points. The 1600 °C isothermal section of CaO–Al2 O3 –SiO2 system tells us that
Al2 O3 has certain stability to CaO when manufacturing refractories used at 1600 °C,
while Al2 O3 in CaO is dangerous; Al2 O3 –SiO2 refractories have certain corrosion
resistance to CaO–SiO2 slag, which increases with the increase of Al2 O3 content;
Al2 O3 is dangerous to SiO2 , while Cao is better; CaO also has considerable stability
to SiO2 .

2.5 CaO–SiO2 –Al2 O3 –MgO

Compounds and solid solutions formed in the MgO–CaO–Al2 O3 –SiO2 quaternary


system are shown in Fig. 2.11. For example, Mg-containing melilite (i.e., magnesium
andalusite) C2 MS2 and Al-containing melilite (i.e., aluminum andalusite) C2 AS can
form a continuous solid solution, i.e., melilite. Such solid solutions are often part
of blast furnace slag [12]. There are 34 phases in the quaternary system, including

Fig. 2.11 Compounds and solid solutions in MgO–CaO–Al2 O3 –SiO2 quaternary system (the
cylinder in the figure represents solid solution)
36 2 Phase Diagram and Equilibrium

four components, 14 binary mesophases, 15 ternary mesophases, and 1 quaternary


mesophase.
The above phase diagram is quite complex, and the phase equilibrium relationship
cannot be intuitively expressed. Table 2.5 lists the properties, approximate composi-
tion, and temperature of some relevant invariable points in the above phase diagram
[13]. For the multivariate phase diagrams of quaternary systems and above, a fixed
component can be used to plot a ternary phase diagram to explore the relevant research
laws.
For blast furnace slag, the binary basicity is generally 1.2 (The molar ratio of
CaO and SiO2 is 1.286). When the basicity is fixed, the phase diagram is shown in
Fig. 2.12, and the characteristics of specific variableless points are shown in Table 2.6.
The low-temperature liquid phase area in the phase diagram is mainly concentrated
on the side close to (CaO)1.2 · SiO2 component. This is because there is a low melting
point calcium silicon compound phase in the slag system. The closer it is to (CaO)1.2
· SiO2 component, the lower melting point in the slag system, the better the slag
fusibility. Near both ends of MgO and Al2 O3 , the slag melting point continues to
rise. The red line in the phase diagram represents the content of 10% mol, 20%
mol, and 30% mol MgO, respectively. With the increase of MgO content, the red
line gradually shortens, which means that the liquid phase zone of the slag becomes
narrower. The high melting point spinel phase is mainly precipitated, making the slag
more difficult to melt; The brown line in the phase diagram represents the content of
10% mol, 20% mol, and 30% mol Al2 O3, respectively. With the increase of Al2 O3
content, the brown line also gradually shortens and the liquid phase area of slag
narrows. At the same time, high melting point melilite phase is mainly precipitated,
which also reduces the melting property of slag.
In the MgO–CaO–Al2 O3 –SiO2 quaternary system, the influence of MgO content
change on blast furnace slag is noteworthy. The CaO–Al2 O3 –SiO2 phase diagrams
at 5 wt%, 10 wt%, 15 wt%, and 20 wt% MgO content is listed in Fig. 2.13. The red
area is the liquid phase area at 1300 °C, the yellow area is the liquid phase area at
1400 °C, and the blue area is the liquid phase area at 1500 °C. In the MgO–CaO–
Al2 O3 –SiO2 quaternary system, the low-temperature liquid phase region is mainly
concentrated on the side of the calcium silicon line close to silicon. With the increase
of MgO content from 5 to 15 wt%, the low-temperature liquid phase region moves up
slightly, indicating that the increase of basicity oxide MgO can increase the content
of acid oxide SiO2 in the slag and make the slag more selective. At the same time,
the liquid phase region at 1400 °C first becomes larger and then smaller, reaching the
maximum at about 10%, This is also the reason why the MgO content of many blast
furnace slag is about 10 wt%; When the MgO content is 5–15 wt%, the liquid phase
region at 1400 °C and 1300 °C increases, but when the MgO content is 20 wt%, the
liquid phase area at 1400 °C decreases slightly and the liquid phase area at 1300 °C
disappears directly.
2.5 CaO–SiO2 –Al2 O3 –MgO 37

Table 2.5 Properties, approximate composition, and temperature of variableless points of MgO–
CaO–Al2 O3 –SiO2 slag system
Variableless points Reaction Composition/mass% Temperature/°C
CaO MgO Al2 O3 SiO2
1 L + A3 S2  9–11 6–8 20–22 61–63 1300
M2 A2 S5 + CAS2 +
SiO2
2 L  M2 A2 S5 + 6–8 13–15 17–19 60–62 1200
SiO2 + MS + CAS2
3 L  M2 S + MS + 8–10 16–18 19–21 53–55 1240
CAS2 + M2 A2 S5
4 L  SiO2 + MS + 12–14 10–12 18–20 56–58 1170
CAS2 + CMS2
5 L + MA  CAS2 + 9–11 17–19 20–22 50–52 1300
M2 A2 S5 + M2 S
6 L  M2 S + CMS2 14–16 12–14 18–20 52–54 1200
+ MS + CAS2
7 L  SiO2 + β-CS + 22–24 6–8 15–17 53–55 1135
CMS2 + CAS2
8 L + α-CS  β-CS 22–24 5–7 16–18 53–55 1150
+ SiO2 + CAS2
9 L + Al2 O3  MA 12–14 6–8 32–34 46–48 1475
+ CAS2 + A3 S2
10 L  CMS2 + CAS2 29–31 9–11 12–14 46–48 1210–1225
+ Mel
11 L  α-CS + CAS2 32–34 6–8 15–17 43–45 1220–1266
+ Mel
12 L + α-CS  β-CS 30–32 7–9 14–16 45–47 1230
+ CAS2 + Mel
13 L  β-CS + CAS2 30–32 8–10 13–15 45–47 1210–1220
+ CMS2 + Mel
14 L + MgO + M2 S  38–40 11–13 8–10 39–41 1400–1425
MA + CMS
15 L  M2 S + CMS + 37–39 11–13 9–11 39–41 1380
MA + Mel
16 L + MgO  CMS 42–44 10–12 9–11 36–38 1400–1410
+ C3 MS2 + MA
17 L + MgO  C3 MS2 47–49 6–8 13–l5 30–32 1410
+ MA + β-C2 S
18 L  MA + CAS2 + 11–13 24–26 21–23 40–42 1230
M2 S + Mel
19 L + CA2  MA + 35–37 3–5 49–51 9–11 1475
CA + C2 AS
(continued)
38 2 Phase Diagram and Equilibrium

Table 2.5 (continued)


Variableless points Reaction Composition/mass% Temperature/°C
CaO MgO Al2 O3 SiO2
20 L + CA2  CA + 45–47 5–7 39–41 7–9 1350
β-C2 S + MA
21 L + MA  CA + 46–48 5–7 38–40 7–9 1390
β-C2 S + MgO
22 L  MgO + β-C2 S 47–49 4–6 38–40 7–9 1300
+ CA + C12 A7
23 L  MgO + β-C2 S 50–52 5–7 35–37 6–8 1295
+ C3 A + C12 A7
24 L + C3 S  β-C2 S 53–55 5–7 25–27 13–15 1380
+ C3 A + MgO
25 L + CaO  MgO + 56–58 5–7 24–26 11–13 1395
C3 S + C3 A
26 L  CMS2 + M2 S 22–24 11–13 13–15 50–52 1225
+ CAS2 + Mel
27 L + C3 MS2  47–49 5–7 14–16 30–32 1390
β-C2 S + MA + Mel
28 L  C3 MS2 + MA 41–43 10–12 10–12 36–38 1380
+ CMS + Mel

Fig. 2.12 Properties of variableness points in phase diagram of (CaO)1.286 · SiO2 –Al2 O3 –MgO
system
Table 2.6 Properties, approximate composition, and temperature of variableness points of MgO–CaO–Al2 O3 –SiO2 slag system at basicity 1.2
Num Corresponding triangles Composition/mol Temperature (°C)
(CaO)1.286 · SiO2 MgO Al2 O3
1 Al2 O3 (s4)–CaMg2 Al16 O27 –Spinel 0.20 0.12 0.67 1708
2.5 CaO–SiO2 –Al2 O3 –MgO

2 Al2 O3 (s4)–Ca2 Mg2 Al28 O46 –CaMg2 Al16 O27 0.21 0.12 0.67 1701
3 Al2 O3 (s4)–Ca2 Mg2 Al28 O46 –CaAl12 O19 0.27 0.08 0.65 1664
4 Ca2 Mg2 Al28 O46 –CaAl12 O19 –CaAl4 O7 0.42 0.04 0.54 1531
5 Ca3 MgSi2 O8 –Monoxide-a-Ca2 SiO4 0.56 0.37 0.06 1511
6 Ca2 Mg2 Al28 O46 –CaAl4 O7 –CaMg2 Al16 O27 0.46 0.06 0.48 1499
7 CaMg2 Al16 O27 –Melilite–Spinel 0.48 0.12 0.40 1488
8 CaAl4 O7 –CaMg2 Al16 O27 –Melilite 0.48 0.06 0.45 1480
9 Ca3 MgSi2 O8 –Monoxide–Spinel 0.53 0.36 0.11 1441
10 Ca3 Si2 O7 –CaSiO3 (s2)-a-Ca2 SiO4 0.91 0.07 0.02 1416
11 Ca3 Si2 O7 –a Ca 2 SiO4 -a-Ca2 SiO4 0.90 0.06 0.03 1414
12 Ca3 MgSi2 O8 –Melilite–Spinel 0.57 0.29 0.14 1412
13 Ca3 MgSi2 O8 –Melilite-a-Ca2 SiO4 0.65 0.23 0.12 1407
14 Melilite–a Ca 2 SiO4 -a-Ca2 SiO4 0.70 0.17 0.13 1399
15 Ca3 Si2 O7 –Melilite–a Ca2 SiO4 0.81 0.05 0.15 1325
39
40 2 Phase Diagram and Equilibrium

Fig. 2.13 Phase diagrams of CaO–Al2 O3 –SiO2 with fixed MgO content (5 wt%, 10 wt%, 15 wt%,
20 wt%)
References 41

References

1. Huang, X. (2013). The principle of iron and steel metallurgy. Metallurgical Industry Press.
2. Chen, S. (2007). Analysis and application of phase diagram. Metallurgical Industry Press.
3. Fenner, C. N. (1913). The stability relations of the silica minerals. American Journal of Science,
36(214), 440.
4. Guo, J. (1995). C12A7 and its derivative structure. Foreign Building Materials Science and
Technology (1), 14–15.
5. Tong, Z., Xie, S., Zhang, L., et al. (2011). Preparation and application status of dodecacalcium
heptaaluminate materials. Nonferrous Metal Science and Engineering, 002(006), 16–21.
6. Ruszak, M., Witkowski, S., Pietrzyk, P., et al. (2011). The role of intermediate calcium alumi-
nate phases in solid state synthesis of mayenite (Ca12 Al14 O33 ). Functional Materials Letters,
4(02), 183–186.
7. Hayashi, K., Matsuishi, S., Kamiya, T., et al. (2002). Light-induced conversion of an insulating
refractory oxide into a persistent electronic conductor. Nature, 419(6906), 462–465.
8. Hayashi, K., Hirano, M., Matsuishi, S., et al. (2002). Microporous crystal 12CaOx7Al(2)O(3)
encaging abundant O(−) radicals. Journal of the American Chemical Society, 124(5), 738.
9. Tomsen, V., Tomsen, V. F. J., Lucio, M. M., et al. (2015). Study of the evolution of phase
calcium aluminate through the method for polymeric precursors C12A7. Materials Science
Forum, 820, 137–142.
10. Matsuishi, S. (2003). High-density electron anions in a nanoporous single crystal:
[Ca24 Al28 O64 ]4+(4e−). Science, 301(5633), 626–629.
11. Osborm, E. F., & Muan, A. (1964). In E. M. Levin, C. R. Robbins, & H. F. Mcmurdie (Eds.),
Phase diagrams for ceramists (p. 219). American Ceramic Society and Edward Orton Jr.
Ceramic Foundation.
12. Chen, Z. (2014). Phase diagram and refractory materials. Metallurgical Industry Press.
13. Jia, C., & Jia, X. (2012). Using geometric prediction method to formulate phase diagrams of
ternary and quaternary systems. Metallurgical Industry Press.
Chapter 3
Slag Structure of High Alumina Blast
Furnace Slag

The physicochemical properties of slag at high temperature are closely related to its
structure, which are mainly reflected in the following aspects: (1) Melting property
refers to the chemical reaction of each other to form liquid phase under the appropriate
temperature and atmosphere conditions, and it is the process of the transition of
ordered oxide crystal into a disordered liquid phase. (2) The fluidity and conductivity
reflect the migration of ions and ion clusters and the adjustment of slag structure.
(3) The surface properties are also related to the distribution of ions and clusters
in slag structure of the surface layer. (4) The desulfurization performance is related
to the distribution of free oxygen ions in slag. Therefore, it is of great significance
to study the structure of blast furnace slag for mastering the performance of blast
furnace slag, which is also why many metallurgical researchers pay attention to
the relationship between slag composition and structure. To deeply understand and
clarify the scientific law between slag physicochemical properties and its chemical
composition, many metallurgical scientists combine structure simulation methods
such as molecular dynamics with structure characterization methods such as Raman
spectrum to fully determine slag structure information.

3.1 Basic Concepts of Slag Structure

3.1.1 Components and Classification of the Blast Furnace


Slag

Conventional blast furnace slag belongs to silicate-based system. At present, the


research on the structure of blast furnace slag took the related research methods of
silicate glass and geological lava as the reference [1–3]. According to the theory of
network structure, the common components in the blast furnace slag are divided into

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 43
X. Lv and Z. Yan, High Temperature Physicochemical Properties of High Alumina Blast
Furnace Slag, https://doi.org/10.1007/978-981-19-3288-5_3
44 3 Slag Structure of High Alumina Blast Furnace Slag

Table 3.1 The parameters of the ions in blast furnace slag


Classification Ion Radius Electrostatic Electronegativity Oxygen Single
10−10 m potential z/r coordination bond
number energy
(kJ/mol)
Network Si4+ 0.41 9.76 1.90 4 443
formers B3+ 0.26 11.54 2.00 3 498
4 372
P5+ 0.34 14.71 2.15 4 368–464
Intermediates Al3+ 0.50 6.00 1.50 4 330–422
6 222–280
Ti4+ 0.68 5.88 1.54 4 458
6 305
Network Ca2+ 1.06 1.89 1.00 6 134
modifiers Mg2+ 0.65 3.08 1.20 6 155
Ba2+ 1.43 1.40 0.90 6 138
Mn2+ 0.91 2.20 1.50 6 121
Fe2+ 0.75 2.67 1.80 6 108
Li+ 0.60 1.67 1.00 4 151
Na+ 0.95 1.05 0.90 6 84
K+ 1.39 0.72 0.80 6 81

network formers, network modifiers, and intermediates according to the single bond
energy between cation and oxygen as forming oxides.
The main parameters of common oxides in blast furnace slag are listed in Table
3.1 [4, 5]. According to the single bond energy, oxides can be divided into three
categories: those with single bond energy of more than 335 kJ/mol are called network
formers, which can form network structure alone as SiO2 , B2 O3 , P2 O5 , etc. Those
with single bond energy below 225 kJ/mol are network modifiers, such as Na2 O,
Cao, MgO, etc. The cations of these oxides have weak interaction with O2− ions,
which can provide O2− ions to destroy the network structure of modified slag. Those
with single bond energy between them are intermediates, such as Al2 O3 , TiO2 , etc.,
and their functions are between network formers and network modifiers.

3.1.2 Composite Anions in the Slag

The theory of the ionic structure of molten slag suggests that the primary ions that
make up slag are simple ions, such as Ca2+ , Mg2+ , Mn2+ , O2− , F− , etc. And complex
anions are formed by combining more than two atoms or simple ions, such as Six Oy z− ,
3.1 Basic Concepts of Slag Structure 45

Px Oy z− , Alx Oy z− , and so on (x and y are the atomic numbers of Si, P, Al, and O,
respectively. Z indicates the number of ion clusters).
Silicon oxide complex anion (Six Oy z− ) is the primary complex anion in the silicate
slag system. Its primary structural unit is SiO4 4− , which is called tetrahedral structure.
The central Si4+ is in the tetrahedral gap tightly integrated by O2− [4, 6–8]. These
tetrahedrons are connected with other tetrahedrons by oxygen ions in the vertex angle.
Because the electrostatic potential of Si4+ in each tetrahedron is larger enough, the
maximum space and the minimum repulsion force can be achieved only by joining the
tetrahedrons through the vertex angle rather than the edge or surface. The structure
and shape of the polymerized SiO2 complex anion are shown in Fig. 3.1. It can
be seen that the more polymerization of SiO4 4− network, the more bridge oxygens
(O2− ) are shared. the network structure of SiO4 varies from point, line, and surface to
three-dimensional network structure. The slag structure changes with the change of
composition, e.g., adding network modifier oxide (MO) can dissociate the complex
SiO2 structure with higher polymerization into simple network structure of SiO2
with low polymerization, which is called depolymerization of network (Fig. 3.2).
According to the different functions of the oxygen in the network of the slag, oxygen
ions can be divided into three categories in molten slag, the one with the saturated
valence of two Si4+ ions is called bridging oxygen (BO), which is expressed by
O0 ; Non-bridged oxygen (NBO), which is unsaturated in valence with one Si4+ , is
expressed as O− ; It exists in the form of free oxygen ions, which is called free oxygen,
or O2− .
P2 O5 is similar to SiO2 in molten slag due to that PO4 3− unites can polymerize
to form a series of Px Oy z− , but PO4 3− also can exist because of the low phosphorus
content in the slag. Generally, PO4 3− is a tetrahedral structure. Because P5+ is +5

Fig. 3.1 Schematic diagram of the Six Oy z− structure


46 3 Slag Structure of High Alumina Blast Furnace Slag

Fig. 3.2 Schematic of depolymerization

valence, there must be a double bond between O2− and P5+ . B2 O3 is also considered
as network forming unit. Under different conditions, triangle BO3 3− or tetrahedral
BO4 5− structural units can be formed, and they can also be connected with SiO4 4−
by bridge oxygen to participate in network structure formation. However, because B
can form a triangle or tetrahedron, when the chemical composition of B-containing
materials changes, the triangle and tetrahedron would change, resulting in the muta-
tion of their structure and properties. This phenomenon is called the boron anomaly
[5]. In addition, the melting point of B2 O3 is low, and a small amount of B2 O3 can
obviously reduce the melting point of slag, so it can be used as a flux.
Al2 O3 is a typical oxide in blast furnace slag. When Al3+ forms an ionic bond
with other anions, it acts as a network modifier (the oxygen coordination number
of Al is 5 or 6); When it forms AlO4 5− tetrahedron through Al–O covalent bond
and participates in the formation of network structure, it acts as a network former
(coordination number is 4). However, due to the different charge numbers, aluminum
oxygen tetrahedron AlO4 5− requires metal cations to provide charge compensation
to achieve charge balance [6, 9, 10]. Due to the particularity of Al2 O3 , it is essential
to determine the role of Al3+ in slag in order to mater the properties and structure
of the slag system containing Al2 O3 . When smelting vanadium titanomagnetite in
a blast furnace or using titanium bearing burden for furnace protection operation,
there will be a certain amount of TiO2 in the slag, which exists in the form of four
coordinated TiO4 4− or six-coordinated metal ions [11–14].

3.2 Parameters to Represent the Structure of Slags

The properties of slag are closely related to its structure. In order to predict the perfor-
mance of slag according to its structure, it is essential to use appropriate parameters
to characterize the structure of slag and establish the relationship between structure
and the properties. Generally, the commonly used characterization parameters of
V
silicate structure are basicity (R) and optical basicity ( ), NBO/T and Qn , bridging
oxygen (BO, O0 ), non-bridging oxygen (NBO, O− ), and free oxygen (FO, O2− ).
3.2 Parameters to Represent the Structure of Slags 47

3.2.1 Basicity (R)

According to Lewis slag acid-based theory [15], the basic oxide in slag can offer free
oxygen ion O2− , the electron donor as a network modifier; Acidic oxides can absorb
free oxygen O2− and transform themselves into less polymerization units. Network
formers are electron acceptors. In addition, a few oxides can offer O2− in acidic slag,
showing basicity, while in the basicity slag, it can absorb O2− and transform into less
polymerization units, showing acidity. These kinds of oxides are called amphoteric
oxides, i.e., TiO2 and Al2 O3 . As the carrier of these electrons, the activity of O2− in
slag can be used to characterize the basicity of slag, but there is no direct and feasible
experimental method to determine O2− activity. Therefore, in practice, the ratio of
the mass fraction of basic oxides to the mass fraction of acid oxides in slag is defined
as slag basicity [4, 16].

Binary basicity: R2 = CaO (wt%)/SiO2 (wt%) (3.1)

Ternary basicity: R3 = (CaO (wt%) + MgO (wt%))/SiO2 (wt%) (3.2)

Quaternary basicity: R4 = (CaO(wt%) + MgO(wt%))/(SiO2 (wt%)


+Al2 O3 (wt%)) (3.3)

In the above formulas, for the convenience of application in the industrial, the
influence of oxides with the same nature on gain or loss of the electron is regarded
as equivalent which is the main flaw of the basicity. Binary basicity is often used
as an index for operation regulation of the blast furnace. The basicity coefficient on
different oxides was suggested to consider the various contribution to the basicity of
the oxides, however, the calculation became more complex in the industrial, which
was limitation in the application.

3.2.2 Optical Basicity (Λ)

Optical basicity was proposed by J. A. Duffy for studying the silicate from 1971
to 1975 [17–21] and was advocated by Sommerville [22] to be applied to the field
of slag. As mentioned earlier, the basicity of slag is related to the basicity of its
constituent oxides, which in turn is related to its O2− behavior. Since the activity
of O2− cannot be measured directly, an optical method was used to determine the
“electron release” ability of the oxide, in which a display agent is added to the oxides.
From the perspective of application in the industrial, CaO is suggested as the
reference oxide, the ratio of the electron release ability of one oxide to that of CaO is
defined as the optical basicity, which is expressed by the symbol . Table 3.2 shows
the optical basicity of common oxides in slag. The optical basicity of oxides can be
48 3 Slag Structure of High Alumina Blast Furnace Slag

Table 3.2 Optical basicity values of some oxides (values cited in Slag Atlas)
Oxides K2 O Na2 O Li2 O CaO MgO MnO Fe2 O3
UV shift 1.15 1 1 0.78
Pauing electronegativity 1.4 1.15 1 1 0.78 0.59 0.48
Electron densicties 1.16 1.11 1.06 1 0.92 0.95
Oxides FeO BaO TiO2 Al2 O3 SiO2 B2 O3 P2 O5
UV shift 1.15 0.605 0.48 0.42
Pauing electronegativity 0.51 1.15 0.61 0.6 0.42 0.4
Electron densities 0.94 1.08 0.65 0.66 0.92 0.42

measured from the shift in the absorption band of the UV region and also can be
calculated from Pauling electron negativity values [18]. The optical basicity of the
slag is represented by the sum of the electron release capacity of each constituent
compound in the slag (Eq. 3.4). The greater the optical basicity, the more free oxygen
ion O2− in the slag can be offered, the lower the polymerization degree of the slag
and the simpler the slag structure [9].

(X i n i i )
=  (3.4)
X i ni

where  is the optical basicity of slag; i is the optical basicity of oxide i; ni is the
number of oxygen atoms in the oxide; X i is the mole fraction of oxide i, mol%.

3.2.3 Bridging Oxygen (O0 ), Non-bridging Oxygen (O− )


and Free Oxygen (O2− )

The concentrations of bridging oxygen (O0 ), non-bridging oxygen (O− ), and free
oxygen (O2− ) in slag can also be used to characterize the slag structure. It still
needs a thermodynamic model to determine these concentrations. In recent years,
many methods based on thermodynamic models have been developed to calculate
the concentrations of different oxygen ions in slag, which is very useful in predicting
the physicochemical properties of slag. For the melt-in equilibrium states, there is an
equilibrium relationship between bridging oxygen (O0 ), non-bridging oxygen (O− ),
and free oxygen (O2− ), as shown in Eq. 3.6.
     
2 O− = O0 + O2− (3.5)

Its equilibrium constant:


a(O0 ) a(O2− )
Kθ = 2
(3.6)
a(O −
)
3.2 Parameters to Represent the Structure of Slags 49

The greater equilibrium constant means the higher the degree of polymerization
of the melt.

3.2.4 NBO/T(Q) and Qn

In the silicate slag system, the depolymerization degree of silicon and oxygen is
related to the non-bridged oxygen number (NBO) bound by Si4+ in slag. It is
expressed by the ratio between the content of effective network modifier used to
depolymerize the network structure contributed from basic oxides and the content
of network formers contributed by the acid oxides. Its symbol is NBO/T (T is Si4+ ,
Ti4+ , and Al3+ ), as shown in Eq. 3.7. NBO/T can be used to characterize the depoly-
merization degree of the network structure of slag. The larger NBO/T, the higher
non-bridge oxygen number and the higher depolymerization degree of slag.
    
NBO/T = 2 X MO + X M2 O − X Al2 O3 / X SiO2 + X Al2 O3 (3.7)

Q = 4 − NBO/T (3.8)

where X is the mole fraction of oxide, mol%.


Some researchers believe that the degree of polymerization of slag network is
easier to understand than the degree of depolymerization. The symbol Q (Eq. 3.8)
represents the degree of polymerization of the slag network structure. The larger Q
value, the higher degree of slag polymerization and the more bridge oxygen number.
NBO/T and Q are essential parameters for evaluating slag structure information.
NBO/T or Q is the non-bridge oxygen number or bridge oxygen number carried by
Si atom in slag calculated according to the slag composition theory, and the structural
information of slag can also be obtained by deconvolving the spectrum of Raman
and other spectral experimental data, which is usually expressed by Qn (n is the
bridge oxygen number). Compared with the theoretical calculation value, Qn based
on spectral data is closer to the actual situation of slag. The previous studies show that
the higher content of Q3 and Q4 in the slag, the higher the degree of polymerization
of the slag. On the contrary, the higher the content of Q0 , Q1 , and Q2 , the higher
the degree of depolymerization of the slag. Park [23, 24] used the ratio of complex
structure content to simple structure content (Q3 /Q2 ) to evaluate the degree of slag
polymerization and the relationship between the degree of polymerization and the
physicochemical properties of slag.
50 3 Slag Structure of High Alumina Blast Furnace Slag

3.3 Characterization Methods of Slag Structure

In the early stage of studying structure of slags, researchers tried to use XRD to
analyze the water quenched slag, but they could not directly obtain effective slag
structure information. Warren et al. [25] obtained the radial distribution function of
SiO2 glass by Fourier analysis, determined that the oxygen coordination number of
Si was 4, and provided a basis for “random network” theory [26]. High-temperature
XRD, X-ray absorption fine structure (EXAFS), and abnormal X-ray (AXS) analysis
methods developed in recent years have also been able to reveal the local chemical
environment around Si element in multicomponent slag system, but the structure
with polymerization degree higher than dimer (Si2 O7 )6− cannot be fully clarified,
and these techniques are limited by the need for synchrotron radiation to obtain
relatively accurate information about specific elements [27–31].
Many spectral methods have also been used to study the structural information
of slag, including Raman spectroscopy (Raman), infrared spectroscopy (FTIR),
X-ray photoelectron spectroscopy (XPS), Mossbauer spectroscopy, and nuclear
magnetic resonance spectroscopy (NMR). Kujumzelis [32], Mysen et al. [33–38]
and McMillan et al. [39–41] used the Raman spectra to systematically measure and
gave a semi-quantitative analysis of the structure of silicate slag by using the charac-
teristic energy band strength. Infrared spectrum and Raman spectrum are molecular
vibration spectra, which can also be used for slag structure analysis. Unlike the Raman
scattering spectrum, the infrared spectrum is the absorption spectrum. The Raman
spectrum is sensitive to nonpolar groups in molecules, while the infrared spectrum is
very sensitive to polar groups. Therefore, the infrared and Raman spectra are comple-
mentary. However, because the accurate solution of the structural unit from the slag
infrared spectrum has not been determined, it is mainly used for qualitative analysis
[1]. The principle of XPS is to radiate the sample with an X-ray to excite and emit the
inner electrons or valence electrons of atoms or molecules within a depth of 10 nm
of the sample surface. By measuring the energy of photoelectrons and the amount
of photoelectrons escaping from the sample surface, the relevant information of slag
structure can be obtained. Therefore, XPS can be carried out for the surface analysis,
but cannot be used to analyze the internal structure of slags like Raman or infrared
spectroscopy. Although XPS can also be used for quantitative analysis, it can only
offer the contents of O0 , O− and O2− and cannot determine the depolymerization type
of silicate slag. Based on the Mossbauer effect, the chemical bond properties, valence
states, and coordination numbers of slag structure can be obtained. However, only
a limited number of nuclei have the Mossbauer effect, which limits its application.
In fact, so far, only a few Mossbauer nuclei such as 57 Fe and 119 Sn have been fully
applied [42–44]. Nuclear magnetic resonance spectroscopy is also widely used in the
structural analysis of slag, but almost all anisotropic interactions are retained in solid
materials, resulting in the sharp broadening of NMR lines, resulting in the inability to
distinguish any fine structure. Therefore, it is necessary to perform nuclear magnetic
magic angle rotation (MAS) on solid samples, which can greatly improve the solid
nuclear magnetic resonance spectrum [45–47].
3.3 Characterization Methods of Slag Structure 51

The development of spectral analysis and detection technology provides strong


technical support for metallurgical researchers to study the structure of high-
temperature slag, but it cannot be avoided that the high-temperature structure of slag
depends on its high-temperature characterization. Although high-temperature in-situ
melt structure spectral analysis is feasible, its application is limited by the high cost
of measuring instruments and the noise of high-temperature analysis. Therefore,
the slag quenched by water or liquid nitrogen at high temperature is still used to
study its high-temperature melt structure. In addition to experimental detection, the
research method of slag structure can also be calculated and simulated. The research
method of computational simulation mainly carries out relevant simulation by the
computer, which has low cost and can obtain all kinds of structural information of
slag at high temperature, which is also the reason why more and more researchers pay
more attention to computational chemistry. The popular calculation and simulation
method for metallurgical slag is molecular dynamics, which is based on Newton’s
classical mechanics and realizes the analysis of the architecture by calculating the
motion trajectories of atoms and molecules in a specific space. The commonly used
slag structure analysis methods and the available information are shown in Table 3.3.
Researchers often choose appropriate methods to study the slag structure information
according to needs [2].

Table 3.3 Summary of various methods used to provide information on the structure of slags
Analytical method Information obtained
X-ray absorption fine structure analysis Coordination number, bond length, bond angle,
(XAFS) effective charge, local ordering degree, etc
Neutron diffraction spectrum (Electron Atomic number, coordination number, radial
diffraction) distribution function, etc
Ultraviolet spectrum (UV) Bridging oxygen, non-bridging oxygen, the
oxidation state of impurity ions, etc
Fourier transform infrared spectra (FTIR) Chemical bond, coordination number, atomic
agglomeration, infrared transmittance, etc
Raman spectra (Raman) Chemical bond, vibration, and rotation of bond,
coordination number, atomic agglomeration, etc
Auger electron spectroscopy (AES) Surface element analysis, chemical valence
analysis, etc
X-ray photoelectron spectroscopy (XPS) Surface element analysis, chemical valence
analysis, chemical bond, etc
Nuclear magnetic resonance (NMR) Quadrupole coupling constant, coordination
number, bond, etc
Mossbauer spectrum (Mössbauer) Chemical bond properties, valence state,
identification of four and six coordination, and
concentration of Fe3+
Molecular dynamics (MD) Coordination number, bond length, bond angle,
oxygen ion type, concentration, etc
52 3 Slag Structure of High Alumina Blast Furnace Slag

3.4 Structure of Silicate

The structure of silicate melts has been comprehensively reviewed by Mysen [6].
Although Mysen’s work focuses on the geological sciences, it can also be extended to
the metallurgical slags. Mills et al. [9], Sohn and Min [1] and Sajid et al. [48] reviewed
the structure and properties relationship for silicate-based slags more relevant to
ironmaking and steelmaking processes. In this section, the role of SiO2 in the structure
of silicate melts has been focused.
According to knowledge on the structural of glasses and slags, it is assumed that
the structure of blast furnace slag is a three-dimensional interconnected network in
which silica (Si–O) tetrahedral is a basic structure unit where each Si4+ is bonded with
four oxygen (O–) ions. The addition of alkaline or alkaline earth cations (e.g., Na+ ,
Ca2+ ) into slag causes the Si–O–Si bond to depolymerize and create ionic bonds (O–
Ca2+ ). The O2− connected ionically with cation is referred to as non-bridging oxygens
(NBO). Depolymerization of slag network resulted from the increase of free oxygen
ions (FO) represented by O2− which have no bonding with any Si4+ ions. In Si–O
tetrahedral, each oxygen ion is bonded covalently to another Si-tetrahedral (Si–O–Si)
known as bridging oxygen (BO). So, the blast furnace slag contains both ionic and
covalent bonds. The bridging of oxygen between Si-atoms (Si–O–Si) forms a three-
dimensional polymerized structure. Generally, SiO2 is considered as network former
and shows fourfold coordination. The structure of silicate melts may be described in
terms of SiO4 4− (monomer), Si2 O7 6− (dimer), Si2 O6 4− (chain), and Si2 O5 2− (sheet)
and SiO2 (3-D) as shown in Fig. 3.1. With the help of Raman spectroscopy, these
structural units of [SiO4 ] tetrahedra with 0, 1, 2, 3, and 4 of non-bridging oxygen
in silicate melts are determined to be in the range of 800–1300, specifically corre-
sponding to 1200, 1100, 950, 900, 850 cm−1 , respectively (Table 3.4) [33, 34, 49, 50].
The anion balance relationship describing the structural units with NBO/Si between
4 and 0 can be expressed by the following equations:

4 = Si2 O7 + O , (2Q = Q + O )
2SiO4− 6− 2− 0 1 2−
(3.9)

7 = Si2 O6 + 2SiO4 , (2Q = Q + 2Q )


2Si2 O6− 4− 4− 1 2 0
(3.10)

6 = 2Si2 O5 + 2SiO4 , (3Q = 2Q + 2Q )


3Si2 O4− 2− 4− 2 3 0
(3.11)

5 = SiO2 + SiO4 , (2Q = 3Q + Q )


Si2 O2− 4− 3 4 0
(3.12)

These anionic structural units are in dynamic equilibrium in the melt, and their
content ratio is affected by the factors such as composition and temperature. With the
increase of NBO/Si, the number of complex anionic structural units (e.g., Si2 O2− 5 )
decreases and the number of simple anionic structural units (e.g., SiO4− 4 ) increases.
Eventually, this leads to a decrease in the degree of polymerization of the melt. Many
factors cause the increase of NBO/Si. Typically, the number of basic cations in the
3.5 Structure of Aluminosilicate Blast Furnace Slag 53

Table 3.4 Characteristic peaks for various structural units


Structural units Wavenumber (cm−1 ) Connection form Vibrational mode
SiO4 4− (Q0 ) ~850 Monomer Si–O2− stretching vibration [33, 34,
36, 37, 49–52]
Si2 O7 6− (Q1 ) ~900 Dimer Si–O− stretching vibration [33, 34,
36, 37, 49–52]
Si2 O6 4− (Q2 ) ~950 Chain − O–Si–O− stretching vibration [33,
34, 36, 37, 49–52]
Si2 O5 2− (Q3 ) ~1100 Sheet − O–Si–O0 stretching vibration [33,
34, 36, 37, 49–52]
SiO2 (Q4 ) ~1200 3-D Si–O anti-symmetrical stretch [33,
34, 36, 37, 49–52]

melt increases, which can ionize more free oxygen ions so as to depolymerize the
complex Si–O structural units. In addition, the increase in temperature also leads to
the depolymerization of complex anions because the depolymerization reaction is an
endothermic reaction. The increase in temperature provides energy and promotes the
forward progress of the depolymerization reaction. It is helpful to obtain the structural
unit information on the molecule-level of slag to explain the change regularity and
the structural mechanism of macro properties of slag.

3.5 Structure of Aluminosilicate Blast Furnace Slag

The primary structural unit of silicate slag is SiO4 4− tetrahedron. Since the ion radius
(0.041 nm) of Si4+ is slightly smaller than that of Al3+ , when Al2 O3 is added into
silicate melt, Si4+ of SiO4 4− tetrahedron can be partially replaced by Al3+ to form
AlO4 5− tetrahedron participating in network. Because the valence state of Al3+ is
different from that of Si4+ , extra metal cations are required for charge compensation
to maintain the charge balance in AlO4 5− tetrahedron uni. If M2+ (e.g., Ca2+ ) ions
participate in the charge compensation, M2+ should be located between two Al3+ ions,
therefore these M2+ ions for charge compensation are no longer worked as network
modifiers. In this case, Al2 O3 is a network former. If there are many kinds of alkali
metal cations in aluminosilicate slag, there would be the problem of “which cation
is preferred to work for charge compensation”. Mills [9] pointed out that cations
with small electrostatic potential have the priority of charge compensation, that is,
K+ > Na+ > Li+ > Ca2+ > Fe2+ > Mg2+ . The aluminum avoidance principle holds
the idea that Al–O–Si bond is easier to form based on the energy minimization rule,
meaning that there is no Al–O–Al bond, and AlO4 5− can only be connected with
SiO4 4− tetrahedron through bridge oxygen in aluminosilicate slag. The aluminum
avoidance principle is hypothesized by Loewenstein, which is still controversial
issue. You et al. [53] and Stebbins et al. [54, 55] have confirmed the existence of the
54 3 Slag Structure of High Alumina Blast Furnace Slag

Al–O–Al bond in the calcium aluminosilicate slag system through experiments and
calculations.
When Al2 O3 is continuously added to the slag, ions can also act as network
modifiers, showing higher coordination (five coordination or six coordination). The
measurement results of viscosity and electricity resistance of aluminosilicate slag
showed that there is a maximum value when the Al/M ratio (the ratio of Al2 O3 to
the mole fraction of basic oxide MO) is 1.0. For this case, the metal cations of all
basic oxides act as the charge compensator of AlO4 5− tetrahedron. When the Al/M
ratio is greater than 1.0 if all Al3+ forms AlO4 5− tetrahedron, the metal cations for
charge compensation should be not enough. There are two views cases on Al3+ for
the current case:
1. part of Al3+ changes from four coordination (tetrahedron) to six coordination
(octahedron);
2. Al3+ ions still maintain four coordination (tetrahedron) in the whole range,
but the tetrahedral structure may rearrange, resulting in appearance of triva-
lent oxygen clusters. The trivalent oxygen cluster comprises three tetrahedrons
connected by one oxygen ion, the oxygen ion is no longer bivalent but trivalent.
The existence of trivalent oxygen ions was also proved in the experiment. For
this case, AlO4 5− tetrahedron was no longer required charge compensation, as
shown in Fig. 3.3 [56–58].
Since for the common blast furnace slag composition, the content of basic oxides
is greater than that of Al2 O3 . As discussed above, the Al3+ should be four-coordinated
to form AlO4 5− tetrahedron due to sufficient charge compensators. However, Park
[59] and Sun et al. [60] showed the presence of amphoteric transition points in the
CaO–SiO2 –Al2 O3 –10 wt% MgO (R = 1.0) slag system by infrared spectroscopy (FT-
IR) with 10 wt% and 16 wt% Al2 O3 . Zhang et al. [61] also found that the slag system
has five-coordinated and six-coordinated Al–O structures ([AlO5 ] and [AlO6 ]), and

Fig. 3.3 Schematic diagram


depicting a tri-cluster
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 55

its content increased with the increase of Al2 O3 content for CaO–SiO2 –Al2 O3 –
9 wt% MgO (R = 1.05) slag system by MAS-NMR analysis. Though molecular
dynamics simulation (MD) on the 30 mol% CaO–SiO2 –(0.06–0.46 mol%) Al2 O3
system, Zheng et al. [62] indicated that trivalent oxygen clusters and five-coordinated
Al–O structures ([AlO5 ]) coexist and their content increase with the increasing of
the Al2 O3 content.

3.6 Effect of Alumina on the Structure of Blast Furnace


Slag

In materials science field, it is generally believed that there is a direct correlation


between microstructure and macroscopic properties. The change of Al2 O3 content in
slag always changes the structure of slag at high temperatures, leading to the evolution
of physical and chemical properties, and this change may result in the transition of
blast furnace slag from silicate system to aluminosilicate system. Researchers in
metallurgical engineering have carried out a lot of studies on the effect of Al2 O3
content on the properties and structure of slag, but there are still shortcomings,
mainly in:
• The current understanding of the properties of blast furnace slag is based on the
physical and chemical basis of silicate slag. In contrast, the high-temperature
physical and chemical properties of traditional slag show different degrees of
inapplicability when dealing with high alumina slag systems. The knowledge
system of high-temperature physical chemistry of blast furnace slag system may
not be able to meet the requirement of utilization of the high alumina iron ore
resource in blast furnace smelting process.
• The studies on the blast furnace slag mainly focus on the Al2 O3 content in the
slag in the range of 12–20 wt%. These studies mainly referred to the effect of
Al2 O3 , MgO, and basicity on the properties of blast furnace slag. However, how
about the slag properties for the cases in which the Al2 O3 is higher than 20 wt%?
• The studies on the structure of aluminosilicate blast furnace slag are relatively
lack. The performance of the amphoteric behavior of Al2 O3 in the blast furnace
slag system is not clearly understood, and there are still many contradictions in
literature.
The authors studied the effect of high alumina on the structure and properties of the
blast furnace slag system, namely CaO–MgO–SiO2 –Al2 O3 system. The composition
of the studied slag is shown in Table 3.5. Series A focuses on the effect of Al2 O3 on
the structure and properties of slag with the fixed basicity of 1.20, MgO of 9 wt%,
and TiO2 of 1 wt%. The substitution of SiO2 with Al2 O3 was studied in series B, with
a fixed 40 wt% CaO and 1 wt% TiO2 in the slag. According to the phase diagram of
CaO–9.0 wt% MgO–1 wt% TiO2 –SiO2 –Al2 O3 (Fig. 3.4), the liquids fraction of slag
gradually increases with increasing temperature. Generally, this slag system has a
56 3 Slag Structure of High Alumina Blast Furnace Slag

Table 3.5 Chemical compositions of studied slags (wt%)


No SiO2 CaO Al2 O3 MgO TiO2 C/S* A/S**
A1# 33.63 40.36 16.00 9.00 1.00 1.20 0.48
A2# 33.18 39.81 17.00 9.00 1.00 1.20 0.51
A3# 32.72 39.27 18.00 9.00 1.00 1.20 0.55
A4# 32.27 38.72 19.00 9.00 1.00 1.20 0.59
A5# 31.81 38.18 20.00 9.00 1.00 1.20 0.63
A6# 30.9 37.09 22.00 9.00 1.00 1.20 0.71
A7# 30.00 36.00 24.00 9.00 1.00 1.20 0.80
A8# 29.09 34.9 26.00 9.00 1.00 1.20 0.89
A9# 28.18 33.81 28.00 9.00 1.00 1.20 0.99
B1# 34.00 40.00 16.00 9.00 1.00 1.18 0.47
B2# 32.00 40.00 18.00 9.00 1.00 1.25 0.56
B3# 30.00 40.00 20.00 9.00 1.00 1.33 0.67
B4# 28.00 40.00 22.00 9.00 1.00 1.43 0.79
B5# 26.00 40.00 24.00 9.00 1.00 1.54 0.92
B6# 24.00 40.00 26.00 9.00 1.00 1.67 1.08
B7# 22.00 40.00 28.00 9.00 1.00 1.82 1.27
B8# 20.00 40.00 30.00 9.00 1.00 2.00 1.50
C/S* = CaO (wt%)/SiO2 (wt%); A/S** = Al2 O3 (wt%)/SiO2 (wt%)

Fig. 3.4 Liquid phase


diagram of the
CaO–SiO2 –Al2 O3 –9 wt%
MgO–1 wt% TiO2 slag
system
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 57

broad liquid region within the temperature of blast furnace operating, however, it is
not all the region meets the requirements of blast furnace operation. For example, if
the basicity is less than 1.0, the liquid zone locates in the area with high SiO2 content
(zone 1) with good thermal stability. However, the slag fluidity and desulfurization
ability don’t meet the requirements of blast furnace operation. The traditional blast
furnace slag is located in zone 2, the basicity is between 1.0 and 1.3, and the content
of Al2 O3 is less than 18 wt%. In the region where the content of Al2 O3 is 18–45 wt%,
the content of CaO is 35–45 wt% with low SiO2 content, exhibits a relatively narrow
liquid zone (zone 3). When the temperature is higher than 1450 °C, the studied
slag compositions are all in the full liquid area. Theoretically, the experimental slag
composition should meet the basic melting requirement of BF operation.
For the research on the structure of the designed slag system, molecular dynamics
simulation and Raman spectroscopy were used.

3.6.1 Molecular Dynamics Simulation

3.6.1.1 Introduction of MD Simulation

Molecular dynamics simulation (MD) is a kind of method that obtains the phase
trajectory of the molecular system by numerically solving the classical mechanical
motion equations of the system and counting the structure and properties of the
system. Table 3.6 briefly summarizes the history of molecular dynamics simulation.
Due to the development of molecular simulation theory, methods, and computer tech-
nology, molecular dynamics simulation has become the third method to understand
the structure and properties of slag from the molecular level after the experimental
and theoretical methods [63].
For a certain particle system, the mechanical motion equations of the particles
are constructed under the given initial conditions, and the iterative calculation is
performed within a specific time step to make it tend to a certain dynamic equilib-
rium. By counting the trajectory and position of each particle, the structural infor-
mation, thermodynamic and kinetic properties of the system can be obtained. MD
is based on molecular mechanics, and the key is to choose an appropriate potential
energy function. For the convenience of calculation, experiments are often used to
fit the parameters in the potential energy function. Therefore, most potential energy
functions in molecular dynamics research methods are empirical or semi-empirical
in nature, and only specific potential parameters can be selected to simulate a specific
system. Potential functions have been developed for decades, and the accuracy of their
simulations has continued to improve. Commonly used intermolecular interaction
potential functions include Lennard–Jones potential, Morse potential, Buckingham
potential, and Born–Huggins–Mayer potential [63].
Lennard–Jones Potential (LJ)
The form of the Lennard–Jones potential energy function is:
58 3 Slag Structure of High Alumina Blast Furnace Slag

Table 3.6 History of molecular dynamics simulation


Year Author Comments
1957–1959 Alder and Wainwright [64] A classical molecular dynamics simulation method
was proposed and successfully applied to an ideal
“hard sphere” liquid model
1964 Rahman [65] Molecular dynamics simulations of liquids were
carried out using the continuum potential model
1968 Verlet [66] The well-known Verlet algorithm is given, which is a
numerical iterative algorithm for the displacement,
velocity, and acceleration of particle motion during
molecular dynamics simulation
1972 Lees and Edwards [67] Extended to non-equilibrium systems with velocity
gradients
1980 Andersen [68] Molecular dynamics simulation of the system under
constant pressure is discussed
1981 Parrinello and Rahman [69] The molecular dynamics model under constant
pressure conditions is given, and it is extended to the
range where the shape of the cell can be changed with
the motion of the particles in it
1983 Gillan and Dixon [70] Generalized to non-equilibrium systems with
temperature gradients, molecular dynamics of
non-equilibrium states are formed
1984 Nosé [71] Molecular dynamics simulations under constant
temperature conditions are given to solve the
temperature control problem in molecular dynamics
simulations
1985 Car and Parrinello [72] A first-principles molecular dynamics method, which
organically unifies electron theory and molecular
dynamics method is proposed
1991 Çağin and Pettitt [73] A grand canonical ensemble molecular dynamics
method for adsorption problems


12
6
  δi j δi j
u LJ ri j = 4εi j − (3.13)
ri j ri j

where ri j is the distance between particles i and j. εi j and δi j are potential energy
parameters, which vary with the types of atoms; exponents 12 and 6 are related
to the hardness of the repulsive force and the range of action of the attractive force,
respectively. There are two sources of the repulsive force term: the Coulomb repulsion
between the nucleus and the nucleus and the overlap energy generated by the Pauli
exclusion principle between electrons and electrons.
Morse Potential
The form of the Morse potential energy function is:
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 59

   2
u Morse ri j = De 1 − e(−β (ri j −re )) (3.14)

where ri j is the distance between particles i and j; re is the equilibrium bond length;
De and β are potential energy parameters. The curve analytical expressions of Morse
potential and Lennard–Jones potential are similar to a large extent, and Morse poten-
tial is often used to describe the potential pair part of various multi-body potential
functions.
Buckingham Potential
The form of the Buckingham potential energy function is:
 
u BK ri j = Ai j e−Bi j ri j − Ci j /ri6j (3.15)

where ri j is the distance between particles i and j; Ai j and Bi j are potential energy
parameters; Ci j is the van der Waals force constant. Adjusting Buckingham’s expo-
nential term coefficient can adjust the repulsive force hardness. However, the compu-
tational complexity of the exponential function is many times larger than that of the
power function, and the Buckingham potential function is rarely used in practical
molecular simulations.
Born-Huggins-Mayer Potential (BHM)
The form of the Buckingham potential energy function is:

  Z i Z j e2 σi +σ j −ri j
C D
u BHM ri j = + Ai j e ρi j − 6 − 8 (3.16)
4π ε0 ri j ri j ri j

where ri j is the distance between particles i and j. Z i and Z j are the charges of the
particles; σ i and σ j are the distance parameters characterizing the particle radius. Aij
and ρ ij are adjustable parameters. C and D are van der Waals forces constant. The first
term in the BHM potential function is the Coulomb potential, the second term is the
short-range repulsion term, the third term describes the interaction between dipoles
and dipoles, and the fourth term describes the dipole-quadrupole re interaction. This
potential function considers the Coulomb potential between ions so that it can be well
applied to ionic compounds such as alkali metal oxides. Among them, the BHM-
type potential function is mainly aimed at ionic compounds, which is in line with the
characteristics of metallurgical slag and has good applicability.

3.6.1.2 Methodology of MD Simulation

The simulation was performed using the Born–Huggins–Mayer (BMH) potential,


which has been extensively used to study glass or slag structures. According to Table
3.5, the number of different atoms was calculated, given that the total number is
approximately 6200, with the density setting at 3.0 g/cm3 . All atoms were assumed
60 3 Slag Structure of High Alumina Blast Furnace Slag

to be placed in a cubic box. The 3D periodic boundary conditions were applied to the
model box, where all the atoms were randomly inserted and the Gear integration of
motion equations was used. The atoms were equilibrated at 5000 K for 15,000 steps
with a time step of
t = 1 fs (10–15 s). Subsequently, the system was cooled down
to 1773 K within 2000 steps and equilibrated at 1773 K for 20,000 steps to acquire
the structural information by statistics analysis. The simulation was performed with
more steps, but minimal variation was observed.

3.6.1.3 Radial Distribution Function (RDF)

The slag has long-range disorder and short-range order at high temperatures. The
radial distribution function (gij (r)) can be used to describe the features of short-range
orders of the melt. The analytical expression is as follows:

1 n(r ) V  n(r )
gi j (r ) = = (3.17)
ρ V Ni N j j 4πr 2
r

where N i and N j are the number of i atoms and j atoms in the system, respectively; V
is the volume of the system; ρ is the system density; n(r) is characterized by i atom
around
r. The average number of j atoms in the range of r/2 spherical diameter.
The radial distribution function is a density functional, which represents the average
number of j atoms per unit volume away from i atoms. For the same atom, N j = N i −
1. Radial distribution function is one of the most important analysis parameters in
molecular dynamics simulation. The abscissa corresponding to the first peak can be
regarded as the bond length between i and j atoms.
The radial distribution function of each bond of A1 slag sample is shown in
Fig. 3.5. The abscissa distance corresponding to the first peak of the function repre-

Fig. 3.5 The RDFs of the


slag system. Reprinted from
Ref. [75]
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 61

Table 3.7 The atom bonding lengths by molecular dynamics simulation (Å)
Si–O Al–O Ti–O Mg–O Ca–O O–O
1.61 1.75 1.9 1.95 2.3 2.65

sents the bond length of the corresponding bond. The bond length reflects the binding
capacity and structural stability between the two ions. Generally speaking, the shorter
bond length, the stronger corresponding binding capacity. The formed bond table has
a more stable structure and is not easy to be destroyed. Table 3.7 shows the bond
length of each ion in the slag combined with oxygen ion, which is consistent with
the results of neutron diffraction experiment by Tandia and Mauro [74] and MD
simulation by Zhang et al. [11]. It is not difficult to find from Table 3.7 that in the
slag system, the Si–O bond length is the shortest and the most stable, followed by
Al–O bond, the Ti–O, and Mg–O bond lengths are closed, the Ca–O bond energy is
the weakest, and the corresponding alkalinity is stronger. In addition, the peak width
and intensity of the first peak in the radial distribution function can also explain the
binding strength between ion pairs. The sharper means more intense of the binding
force, the stronger binding force is. It can also be seen from Fig. 3.5 that the sharpness
of Si–O is significantly higher than that of other ion pairs, which is consistent with
the above conclusion. The other components in the experimental scheme are also
simulated, and the radial distribution function has no obvious change.

3.6.1.4 Coordination Number (CN)

The bond length between ion pairs can be obtained from the radial distribution
function, and the information of the first coordination number can be obtained by
integrating it into the trough of the first peak. The coordination number reflects the
ion coordination of a polymer structure in a short range, and the analysis is shown
as follows:
r
4π N j
CNi j (r ) = r 2 gi j (r )dr (3.18)
V
0

where it refers to the number of coordination ions j nearest to the sphere centered
on ion i. The coordination number thus obtained has roughly the same meaning as
that in the crystal structure, which can well describe the short-range ordered bonding
state in the amorphous structure of slag.
For the slag system, more attention is paid to the binding state between various
ions and oxygen ions. Therefore, the coordination number of ions and oxygen ions
is mainly discussed here. Figure 3.6 shows the change of oxygen atom coordination
number of each atom of A1 slag samples, and Table 3.8 shows the oxygen atom
62 3 Slag Structure of High Alumina Blast Furnace Slag

Fig. 3.6 The CNs of the


slag CASMT [75]. Reprinted
from Ref. [75]

Table 3.8 The CNs of the slag A1


Bond type Si–O Al–O Mg–O Ti–O Ca–O
Coordination number 4.00 4.09 4.98 5.80 6.00
Radius (Å) 2.00 2.70 3.00 3.00 3.20

coordination number of each atom obtained by setting the first peak valley as the
truncation radius. It can be seen from the figure and table that the CNSi–O curve
has a platform when the coordination number is 4.00, indicating that the oxygen
coordination number of Si ion is 4, and SiO4 4− is the basic structural unit as the
network formant; The coordination number of CNAl–O is 4.09, indicating that Al3+
ions mostly exist as network formers in the form of AlO4 5− tetrahedron, but they
are not as stable as SiO4 4− tetrahedron and some Al3+ may have higher coordination
structural units. Figure 3.7 shows the change of oxygen coordination number of Al3+
ions with Al2 O3 content. It can be seen that more than 90% of Al3+ four coordination
structures take up structural unit, about 5% of Al3+ ions exist in five coordination
structures, and there are a small number of three coordination and six coordination
structural units. It can be seen that the slag system studied has enough metal cations
(Ca2+ , Mg2+ ) as charge compensation ions to maintain the charge balance of AlO4 5−
tetrahedron. CNTi–O is 5.80, and the platform is obvious, indicating that TiO2 forms a
six-coordinated octahedral structure as a network modifier; While CNMg–O is 4.98 and
CNCa–O is 6.0, Ca2+ and Mg2+ tend to form hexacoordinated octahedral structure.
It is not difficult to find that there is no stable platform in the coordination curve
of Ca–O, indicating that the hexacoordinated octahedral structure of Ca–O is very
unstable. This instability just leads to that CaO can better provide free oxygen and
significantly promote the depolymerization of slag. In contrast, Mg–O and Ca–O
have similar coordination structures, indicating that both of them play a role in
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 63

Fig. 3.7 Concentration of


Al3+ coordination number of
O as a function of mole
fraction of Al2 O3 . Reprinted
from Ref. [75]

promoting slag to depolymerize. The difference is that Mg–O has a relatively stable
coordination platform and short Mg–O bond length, which shows that the binding
force between Mg–O is stronger than Ca–O, and the depolymerization of MgO is
weaker than CaO.

3.6.1.5 Connection of AlO4 5− to SiO4 4− Tetrahedron

In this blast furnace slag system, Si4+ and Al3+ ions are surrounded by four oxygen
ions to form a stable tetrahedral structure. The connection between tetrahedrons
directly affects the polymerization degree of blast furnace slag. AlO4 5− and SiO4 4−
tetrahedron can form a complex three-dimensional network structure alone, or they
can be connected with each other to form a more complex network polymer. There-
fore, specifying the connection between AlO4 5− and SiO4 4− and revealing the two
tetrahedral structures are helpful to understand the changes in three-dimensional
network structure in the high aluminum slag system. The connection between tetra-
hedron and tetrahedron in three-dimensional structure can be analyzed in detail by
further analyzing the coordination number of Si–Si, Si–Al and Al–Al, Al–Si and the
distribution of Si–O–Si, Si–O–Al, and Al–O–Al.
Figure 3.8 shows the relationship between the Si–Al coordination number of Si4+
and Al3+ ions and Al2 O3 content and Al2 O3 /SiO2 . With the increase of Al2 O3 content
and Al2 O3 /SiO2 , the coordination number of Al–Al and Si–Al increases, while the
coordination number of Si–Si and Al–Si decreases, and the coordination number
of Al–Al and Si–Al increases apparently than that of Si–Si and Al–Si. This shows
that the SiO4 4− tetrahedron connected around the two tetrahedrons becomes less and
less, while AlO4 5− becomes more and more. It should be noted that once the number
of silicon atoms in the slag sample is almost the same as that of aluminum atoms,
the coordination numbers of Si–Si and Al–Al become quite different. For example,
the number of silicon atoms in A1 sample is similar to that of aluminum atoms in
64 3 Slag Structure of High Alumina Blast Furnace Slag

Fig. 3.8 Coordination number of Si–Si, Si–Al, and Al–Al, Al–Si as a function of Al2 O3 content
(a) and Al2 O3 /SiO2 (b) (1500 °C). Reprinted from Ref. [75]

A8 sample, but the Si–Si coordination number of A1 sample is 1.48 and the Al–Al
coordination number of A8 sample is 2.32, which is nearly twice the difference,
indicating that Al3+ ions do not simply replace Si4+ ions in the structure, and the
structural environment around Al3+ ions is more complex than Si4+ ions. On the one
hand, because the electrostatic potential of Al3+ ions is less than that of Si4+ ions, the
repulsive force between silicon ions is greater than that between aluminum ions. On
the other hand, in order to maintain the charge balance of tetrahedral structure after
aluminum ion replaces silicon ion, metal cations need to be near AlO4 5− tetrahedron.
Figure 3.9 shows the relationship between the content of Si–O–Si, Si–O–Al, and
Al–O–Al bonds in slag and Al2 O3 content (Series A) and Al2 O3 /SiO2 mass ratio
(Series B) at 1500 °C, which can also clearly describe the bridging state of AlO4 5−
and SiO4 4− tetrahedron in slag system. It is undeniable that there is an Al–O–Al
bond, and the Al avoidance principle does not apply. With the increase of Al2 O3
content, the content of Si–O–Si bond decreases, and the content of Si–O–Al and
Al–O–Al bond increases. The content of Si–O–Al bond is higher than that of the
other two bonds, indicating that AlO4 5− and SiO4 4− tend to be connected with each

Fig. 3.9 Concentration of Si–O–Si, Si–O–Al, and Al–O–Al as a function of Al2 O3 content (a) and
Al2 O3 /SiO2 (b) (1500 °C). Reprinted from Ref. [75]
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 65

other. When Al2 O3 replaces SiO2 , the calculated structure shows that the content
of Si–O–Si is lower, the content of Al–O–Al is higher, and the content of Si–O–
Al is basically unchanged; In terms of composition, every 1 g of SiO2 replaced by
1 g of Al2 O3 in series B is equivalent to the reduction of 0.0166 mol of SiO4 4−
for every increase of 0.0196 mol of AlO4 5− tetrahedron. The fluctuation of AlO4 5−
tetrahedron is always higher than that of SiO4 4− tetrahedron, so the number of SiO4 4−
is the limiting condition of Si–O–Al content. With the decrease of Si–O–Si and the
significant increase of Al–O–Al, the network structure content of AlO4 5− and SiO4 4−
in the system remains unchanged. According to series tests, with every 1 g of Al2 O3
addition, about 0.5 g of SiO2 reduces, that is, the increase of 0.0196 mol of AlO4 5−
tetrahedron would only reduce 0.083 mol of SiO4 4− . Therefore, compared with B
series, there will be more silicon and aluminum bridged into a network, which is
bound to increase the complexity of the network.

3.6.1.6 Distribution of Different Types of Oxygen

According to the combination mode and function of oxygen ions in slag, there are
three existing states of oxygen ions in slag: Bridge oxygen (BO), that is, oxygen
ions in Si(Al)–O–Si(Al) bond, which connect AlO4 5− or SiO4 4− tetrahedron as a
bridge and form a network. The higher BO concentration is, the more complex
the structure is. Non-bridged oxygen (NBO), that is, the oxygen ion in the Si(Al)–
O–Ca(Mg, Ti) bond, one end is connected to the tetrahedron and the other end
is connected to the alkali metal cation, which destroys the bonding between the
tetrahedrons, thus reducing the complexity of the slag structure. Free oxygen (FO),
that is, oxygen ions in Ca(Mg, Ti)–O–Ca(Mg, Ti) bonds, which are not connected to
any tetrahedral structural units. In the slag system studied in this book, the change
in slag composition inevitably leads to the change in the distribution state of three
oxygen types. The results are shown in Fig. 3.10. It is similar to most silicate slag,
free oxygen content is low and stability is poor. With the increase of Al2 O3 content,

Fig. 3.10 Concentration of types of oxygen as a function of Al2 O3 content (a) and Al2 O3 /SiO2
(b) (1500 °C). Reprinted from Ref. [75]
66 3 Slag Structure of High Alumina Blast Furnace Slag

Fig. 3.11 Concentration of Si–O–(Ca, Mg), Al–O–(Ca, Mg) as a function of Al2 O3 content (a) and
Al2 O3 /SiO2 (b) (1500 °C). Reprinted from Ref. [75]

the non-bridging oxygen content decreases, the bridging oxygen content increases,
and gradually becomes the main oxygen ion. On the one hand, the decrease of
CaO content reduces the content of network modifiers. On the other hand, AlO4 5−
needs metal cations as charge compensators to further reduce the content of network
modifiers. When Al2 O3 replaces SiO2 , the bridging oxygen content increases slightly,
and the non-bridging oxygen content decreases, but it always occupies the position
of main oxygen ions. With the increase of Al2 O3 content, SiO2 content decreases
accordingly, and the total content of network tetrahedron in slag does not change
much. Although the content of CaO and MgO remains unchanged, the increase of
AlO4 5− content needs to consume metal oxides to maintain charge balance, which
is the main reason for the increase of bridge oxygen.
Figure 3.11 shows the distribution of different types of non-bridged oxygen. It is
not difficult to find that the content of Si–O–(Ca, Mg) bond decreases significantly
with the increase of Al2 O3 content and Al2 O3 /SiO2 ratio; The content of Si–O–
(Ca, Mg) bond is higher than that of Al–O–(Ca, Mg) bond, indicating that non-
bridged oxygen tends to connect with SiO4 4− tetrahedron. One possible reason is
that AlO4 5− tetrahedron needs charge compensation. When metal cations are used as
charge compensators, they repel the surrounding metal cations as network modifiers.
AlO4 5− tetrahedron is more surrounded by AlO4 5− tetrahedron, and Al3+ is more
inclined to combine with bridge oxygen to form more complex structures, which is
consistent with the results of Allwardt et al. [47] and Petkov et al. [76]. In addition,
although the difference between the number of Ca2+ and Mg2+ in the slag composition
is nearly three times, the concentrations of Al–O–Ca and Al–O–Mg fluctuate at the
same level, indicating that the charge compensation ability of cationic Ca2+ to AlO4 5−
tetrahedron is stronger than that of cationic Mg2+ .
In addition to the characterization of different oxygen ion types (bridging oxygen,
non-bridging oxygen, and free oxygen), the structural information of slag can also be
characterized by Qn , where n is the number of bridging oxygen in a single tetrahedron.
The larger n value, the more bridge oxygen in a tetrahedral structure, that is, the higher
the degree of slag polymerization. Figure 3.12a, b shows the variation relationship
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 67

Fig. 3.12 Concentration of


Qn and (Q4 + Q3 )/(Q2 +
Q1 ) ratio as a function of
Al2 O3 content. Reprinted
from Ref. [75]

between various polymerization structure contents of Al and Si atoms and Al2 O3


content. It can be seen from the figure that with the increase of Al2 O3 content, whether
for silicon polymer Q(Si) or aluminum polymer Q(Al), Q4 content increases, Q1 and
Q2 contents tend to decrease, and Q0 content is very little and has no obvious change;
The change of Q3 (Al) is relatively stable, while Q3 (Si) increases significantly; Q3 (Al)
and Q4 (Al) account for the majority in the connection form of AlO4 5− tetrahedron,
while Q2 (Si) and Q3 (Si) in the connection form of SiO4 4− tetrahedron. This means
that with the increase of Al2 O3 content, the network complexity composed of AlO4 5−
or SiO4 4− increases, and three oxygen ions in most SiO4 4− tetrahedrons are connected
with other tetrahedrons, which further shows that Al3+ tends to form more complex
structures.
The ratio of complex structure content to simple structure content can be used to
evaluate the degree of polymerization of slag system, and the relationship between it
and the physicochemical properties of slag can be constructed. For this simulation,
the degree of polymerization of slag system is characterized by the ratio of (Q4
+ Q3 )/(Q2 + Q1 ). Figure 3.12c shows the relationship between (Q4 + Q3 )/(Q2
+ Q1 ) ratio and Al2 O3 content. With the increase of Al2 O3 content, CaO content
decreases. Although the binary alkalinity is maintained, the network polymerization
degree composed of Al3+ increases significantly, and the Si4+ polymerization degree
also shows an obvious increasing trend, resulting in a significant increase in the
68 3 Slag Structure of High Alumina Blast Furnace Slag

Fig. 3.13 Concentration of


Qn and (Q4 + Q3 )/(Q2 +
Q1 ) ratio as a function of
Al2 O3 /SiO2 ratio. Reprinted
from Ref. [75]

polymerization degree of slag network structure and more polymerization states


between AlO4 5− or SiO4 4− tetrahedron.
The relationship between various polymerization structure contents of Al and
Si atoms and Al2 O3 /SiO2 ratio is shown in Fig. 3.13a, b. Although Al2 O3 /SiO2
gradually increases, Qn does not fluctuate significantly for silicon polymer Q(Si)
or aluminum polymer Q(Al), and the contents of Q2 and Q3 always occupy the
dominant position. It should be noted that the content of complex structures Q3 (Al)
and Q4 (Al) is much higher than that of Q3 (Si) and Q4 (Si), indicating that AlO4 5−
tetrahedron is preferentially located in the middle of the network structure, while
SiO4 4− tetrahedron tends to act as a terminal at the network boundary. Zheng et al.
[62] reported this phenomenon by molecular dynamics simulation of aluminosilicate
slag. Petkov et al. [76] also found that the depolymerization of network structure
preferentially occurs around SiO4 4− tetrahedron through high-energy X-ray study of
aluminosilicate, that is, the depolymerization of network structure is mainly through
the formation of Si–O–M bond rather than Al–O–M bond. Losq et al. [77] also found
by high-temperature 27 Al NMR that Al3+ preferentially connects SiO4 4− tetrahedrons
in the form of Q4 (Al), and these SiO4 4− tetrahedrons are connected with non-bridging
oxygen (Q3 , Q2 , and Q1 ).
Figure 3.13c shows the relationship between (Q4 + Q3 )/(Q2 + Q1 ) ratio and
Al2 O3 /SiO2 ratio. The network polymerization degree composed of Al3+ has a
minimum value when the Al2 O3 /SiO2 ratio is 0.79, but the polymerization degree of
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 69

Si4+ has no obvious change. Combined with the composition analysis, when the ratio
of Al2 O3 /SiO2 exceeds 0.79, the number of Al3+ in the slag exceeds the number of
Si4+ , and the slag changes from silicate to aluminate. It shows that in the slag system
studied in this paper, Al2 O3 instead of SiO2 will reduce the degree of polymerization
of slag in silicate system, while Al2 O3 instead of SiO2 will increase the degree of
polymerization of slag structure in aluminate system.

3.6.2 Analysis of Slag Structure by Raman Spectroscopy

The slag samples were heated to 1500 °C and kept for three hours to homoge-
neous. Then quickly quenched on a water-cooled plate, but the cooling rate of water
quenching is not precisely known. X-ray diffraction (XRD) analysis was carried
out to ensure the phase of samples. The absence of any characteristic peaks in the
XRD pattern indicates that water quenching offers sufficient cooling rate to obtain
an amorphous structure. Raman spectroscopy (Lab. RAM HR Evolution; HORIBA
Scientific, France) with a 532-nm laser was used for further analysis of the slag struc-
ture. The analysis of the Raman spectra was constrained from 1200 to 400 cm−1 , and
the area fraction and quantitative amounts of the individual silicate and aluminate
bonds were ascertained by deconvoluting the Raman spectra with the Peak Fit V4
software.
The influence of Al2 O3 content and Al2 O3 /SiO2 ratio on slag structure detected by
Raman spectroscopy is shown in Fig. 3.14. With the increase of Al2 O3 content, the
phenomenon of “red shift” can be obviously observed, that is, the spectral line moves
toward high wave number, indicating that the slag structure tends to be complex,
which is consistent with the results of molecular dynamics simulation. The peak
intensity of Al–O structure in low-frequency region increases with the increase of

Fig. 3.14 The Raman spectra of the slag and typical deconvolution. Reprinted from Ref. [75]
70 3 Slag Structure of High Alumina Blast Furnace Slag

Al2 O3 content. The Raman spectrum of each slag sample is further analyzed by
Gaussian spectrum to obtain the relative content of each structural unit.
In the silicate slag system, the content of Qn (Si) structural unit can be obtained by
Raman spectroscopy. Q2 and Q3 represent a more complex polymerization structure.
When their content in the slag samples is high, it indicates that the anion group
of the slag can form a large three-dimensional or chain network structure, and the
polymerization degree of the slag is high; Q0 and Q1 represent relatively simple
polymerization structure. When the content of Q0 and Q1 in the slag sample are
high, the anion group of the slag is mainly tetrahedral monomer or island structure,
and the polymerization degree of the slag is low. When Al2 O3 is added to the silicate,
the slag changes from silicate to aluminosilicate, and the proportion of characteristic
peaks related to Al–O bond increases (400–800 cm−1 low-frequency region), which
indicates that Al3+ ions in the slag participate in the formation of three-dimensional
network structure of the slag. In the Al–O unites, the oxygen ions in Al–O0 structure
and Si–O–Al structure are bridge oxygen, representing a complex polymerization
structure. The Al–O− structure indicates that there is non-bridge oxygen in AlO4 5−
tetrahedron, showing a relatively simple polymerization structure.
Figure 3.15a shows the effect of Al2 O3 content on the content of each structural
unit in slag. Q4 (Si) is not obtained by spectrum resolution, which may exist, but the
content can be basically ignored. It can be seen from the figure that with the increase
of Al2 O3 content, the content of Al–O− , Al–O–Si, and Si(Q0 + Q1 ) structural units

Fig. 3.15 Effect of Al2 O3


on the structure units content.
Reprinted from Ref. [75]
3.6 Effect of Alumina on the Structure of Blast Furnace Slag 71

gradually decreases, the content of Al–O0 structural units gradually increases, and the
content of Si(Q2 + Q3 ) first increases and then decreases slightly. The turning point is
when the mass fraction of Al2 O3 is 24%. The polymerization degree of slag increases,
which is due to the decrease of CaO content in slag and the charge compensation effect
of AlO4 5− tetrahedron. The experimental results of Raman spectroscopy and the
simulation results of molecular dynamics are contradictory in the change trend of Al–
O–Si structural units, which may be due to the fact that molecular dynamics does not
consider the dynamic balance between each structural unit. In the slag polymerization
structure, there is a balance between complex polymerization structure and simple
polymerization structure. Therefore, the ratio of complex polymerization structure
to simple polymerization structure is used to reflect the degree of polymerization
(DOP) of slag.

X ( Q 2 +Q 3 ) + X (Al−O0 ) + X (Al−O−Si)
DOP = (3.19)
X ( Q 0 +Q 1 ) + X (Al−O− )

where X i is the percentage content of structural unit i. The greater the DOP value, the
more complex the slag structure is. Figure 3.15b shows the effect of Al2 O3 content
on slag polymerization degree DOP. With the increase of Al2 O3 content, the slag
polymerization degree increases.

Fig. 3.16 Effect of


Al2 O3 /SiO2 on the structure
units content. Reprinted
from Ref. [75]
72 3 Slag Structure of High Alumina Blast Furnace Slag

Figure 3.16 shows the effect of Al2 O3 /SiO2 ratio on the content and polymerization
degree of each structural unit in the slag. When Al2 O3 replaces SiO2 , the contents
of Al–O0 and Si (Q2 + Q3 ) structural units first decrease and then increase; On the
contrary, the content of Al–O–Si and Si(Q0 + Q1 ) structural units first increase and
then decrease, and the turning point is that the Al2 O3 /SiO2 ratio equals 0.92; The
content of Al–O structural units increases gradually. The degree of polymerization of
slag first decreases and then increases, and there is a minimum when the Al2 O3 /SiO2
ratio is 0.92, indicating that the degree of polymerization of slag structure is the
minimum. It should be noted that when the Al2 O3 /SiO2 ratio is about 0.92, the
number of Al3+ ions in the slag exceeds the number of Si4+ ions, and the slag changes
from silicate to aluminate. The results of Raman spectra show that in silicate system,
Al2 O3 replacing SiO2 would reduce the degree of polymerization of slag, while in
aluminate system, Al2 O3 replacing SiO2 would increase the degree of polymerization
of slag, which is consistent with the results of molecular dynamics simulation.
According to the Raman spectra of the two series slag, it can be seen that although
the Al2 O3 in the slag increases gradually, the slag structure is obviously different,
which can be distinguished by the peak fitting of Raman spectrum. However, there are
still some deficiencies in the study of the structure of aluminosilicate slag by Raman
spectroscopy. Stebbins [78] pointed out that there are three coordination forms of Al3+
in slag: AlO4 5− , AlO5 7− , and AlO6 9− . Only AlO4 5− acts as a network framework
and plays the role of network former. Since the formation of AlO4 5− tetrahedron
requires network modifiers to provide metal cations as charge compensators, it is
agreed by many studies that once there are enough metal cations in the slag (that
is, the content of network modifiers that can provide charge compensation is greater
than that of Al2 O3 ), all Al3+ exist in the form of four coordination. However, Poe
et al. [41] pointed out that the oxygen coordination number of Al3+ is related to the
composition of slag system. Even in 10 mol% Al2 O3 –90 mol% CaO system, there is
still a small amount of five coordinated Al3+ . This shows that even though there are
sufficient network modifiers providing charge compensation in the slag, Al3+ ions
are not all network formers. This is consistent with the results of molecular dynamics
simulation. Some Al3+ ions in slag exist in the form of five coordination, but their
proportion cannot easily be obtained by Raman spectroscopy.

References

1. Sohn, I., & Min, D. J. (2012). A review of the relationship between viscosity and the structure
of calcium-silicate-based slags in ironmaking. Steel Research International, 83(7), 611–630.
2. Mills, K. (2011). The estimation of slag properties. Short course presented as part of Southern
African Pyrometallurgy.
3. Mills, K. C. (1993). The influence of structure on the physico-chemical properties of slags.
ISIJ International, 33(1), 148–155.
4. Huang, X. (2014). Principles of iron and steel metallurgy. Metallurgical Industry Press.
5. Tian, Y. (2009). Glass technology (new edition), China Light Industry Press, Beijing.
6. Mysen, B., & Richet, P. (2005). Silicate glasses and melts: Properties and structure. Elsevier.
7. Mysen, B. (2003). Physics and chemistry of silicate glasses and melts. Elsevier.
References 73

8. Mysen, B. O., & Virgo, D. (1994). Structure and properties of silicate glasses and melts; theories
and experiment. Advanced Mineralogy. Springer, Berlin, Heidelberg. 238-254
9. Mills, K. C., Hayashi, M., Wang, L., & Watanabe, T. (2014). Chapter 2.2—The structure
and properties of silicate slags. Treatise on Process Metallurgy Process Fundamentals, 14(8),
149–286.
10. Stebbins, J. F., Farnan, I., & Xue, X. (1992). The structure and dynamics of alkali silicate
liquids: A view from NMR spectroscopy. Chemical Geology, 96(3–4), 371–385.
11. Zhang, S., Zhang, X., Bai, C., Wen, L., & Lv, X. (2013). Effect of TiO2 content on the structure
of CaO–SiO2 –TiO2 system by molecular dynamics simulation. ISIJ International, 53(2013),
1131–1137.
12. Sandstrom, D. R., Lytle, F. W., Wei, P., Greegor, R. B., Wong, J., & Schultz, P. (1980). Coordi-
nation of Ti in TiO2 -SiO2 glass by X-ray absorption spectroscopy. Journal of Non-crystalline
Solids, 41(2), 201–207.
13. Park, H., Park, J.-Y., Kim, G. H., & Sohn, I. (2012). Effect of TiO2 on the viscosity and slag
structure in blast furnace type slags. Steel Research International, 83(2), 150–156.
14. Roskosz, M., Toplis, M. J., & Richet, P. (2004). The structural role of Ti in aluminosili-
cate liquids in the glass transition range: Insights from heat capacity and shear viscosity
measurements. Associate editor: K. J. Russell. Geochimica et Cosmochimica Acta, 68(3),
591–606.
15. Lewis, G. N. (1938). Acids and bases. Journal of the Franklin Institute, 226(3), 293–313.
16. Wang, X. (2000). Iron and steel metallurgy (ironmaking section). Metallurgical Industry Press.
17. Duffy, J., & Ingram, M. (1976). Optical basicity—V: A correlation between the Lewis (optical)
basicity of oxyanions and the strengths of brønsted acids in aqueous solution. Journal of
Inorganic and Nuclear Chemistry, 38(10), 1831–1833.
18. Duffy, J., & Ingram, M. (1975). Optical basicity—IV: Influence of electronegativity on the
Lewis basicity and solvent properties of molten oxyanion salts and glasses. Journal of Inorganic
and Nuclear Chemistry, 37(5), 1203–1206.
19. Duffy, J., & Ingram, M. (1974). Lewis acid-base interactions in inorganic oxyacids, molten salts
and glasses—II [1]: Ligand competition reactions of lead (II) and bismuth (III) with various
anions in molten dimethylsulfone. Journal of Inorganic and Nuclear Chemistry, 36(1), 39–42.
20. Duffy, J., & Ingram, M. (1974). Lewis acid-base interactions in inorganic oxyacids,
molten salts and glasses—III: Co-ordination studies of thallium, lead and bismuth in
sulfate/chloride/bromide glass systems. Journal of Inorganic and Nuclear Chemistry, 36(1),
43–47.
21. Duffy, J., & Ingram, M. (1971). Establishment of an optical scale for Lewis basicity in inorganic
oxyacids, molten salts, and glasses. Journal of the American Chemical Society, 93(24), 6448–
6454.
22. Sommerville, I., & Sosinsky, D. (1984). 2nd International Symposium on Metallurgical Slags
and Fluxes (pp. 1015–1026). Metallurgical Society of the AIME.
23. Park, J. H. (2013). Structure-property relationship of CaO-MgO-SiO2 slag: Quantitative
analysis of Raman spectra. Metallurgical & Materials Transactions B, 44(4), 938–947.
24. Hyun, P. J. (2012). Structure-property correlations of CaO-SiO2 -MnO slag derived from Raman
spectroscopy. Isij International, 52(9), 1627–1636.
25. Warren, B. E., Krutter, H., & Morningstar, O. (1936). Fourier analysis of X-ray patterns of
vitreous SiO2 and B2 O3 . Journal of the American Ceramic Society, 75(1), 11–15.
26. Zachariasen, W. H. (1932). The atomic arrangement in glass. Journal of the American Chemical
Society, 54(10), 3841–3851.
27. Waseda, Y., Sugiyama, K., & Toguri, J. M. (1995). Direct determination of the local structure
in molten alumina by high temperature X-ray diffraction. Zeitschrift Für Naturforschung A,
50(8), 770–774.
28. Waseda, Y., & Toguri, J. M. (2015). The structure and properties of oxide melts. Revista
Brasileira De Saúde Materno Infantil, 6(1), 69–74.
29. Waseda, Y., & Toguri, J. M. (1998). The structure and properties of oxide melts: Application
of basic science to metallurgical processing. World Scientific Publishing.
74 3 Slag Structure of High Alumina Blast Furnace Slag

30. Waseda, Y. J. C. M. Q. (1981). Current structural information of molten slags by means of a


high temperature X-ray diffraction. Canadian Metallurgical Quarterly, 20(1), 57–67.
31. Waseda, Y., & Shiraishi, Y. (1980). The structure of the molten FeO–Fe2 O3 –SiO2 system by
X-ray diffraction. Transactions of the Japan Institute of Metals, 21(1), 51–62.
32. Kujumzelis, T. G. (1936). Raman-Effekt und Struktur der Gläser. Zeitschrift für Physik, 100(3–
4), 221–236.
33. Mysen, B. R. O., Virgo, D., & Scarfe, C. M. (1980). Relations between the anionic structure
and viscosity of silicate melts—A Raman spectroscopic study. American Mineralogist, 65,
690–710.
34. Mysen, B. O., Finger, L. W., Virgo, D., & Seifert, F. A. (1982). Curve-fitting of Raman spectra
of silicate glasses. American Mineralogist, 67(7–8), 686–695.
35. Mysen, B. O. (1990). Role of Al in depolymerized, peralkaline aluminosilicate melts in the
systems Li2 O-Al2 O3 -SiO2 , Na2 O-Al2 O3 -SiO2 , and K2 O-Al2 O3 -SiO2 . American Mineralo-
gist, 75(1), 120–134.
36. Mysen, B. O., & Virgo, D. (1980). Trace element partitioning and melt structure: An
experimental study at 1 atm pressure. Geochimica et Cosmochimica Acta, 44(12), 1917–1930.
37. Mysen, B. O., Virgo, D., & Kushiro, I. (1981). The structural role of aluminum in silicate melts;
a Raman spectroscopic study at 1 atmosphere. American Mineralogist, 66(7–8), 678–701.
38. Mysen, B. O., Virgo, D., Scarfe, C. M., & Cronin, D. (1985). Viscosity and structure of iron-and
aluminum-bearing calcium silicate melts at 1 atm. American Mineralogist, 70(5–6), 487–498.
39. Daniel, I., Gillet, P., Poe, B. T., & McMillan, P. F. (1995). In-situ high-temperature Raman
spectroscopic studies of aluminosilicate liquids. Physics and Chemistry of Minerals, 22(2),
74–86.
40. McMillan, P. F., Poe, B. T., Gillet, P., & Reynard, B. (1994). A study of SiO2 glass and
supercooled liquid to 1950 K via high-temperature Raman spectroscopy. Geochimica et
Cosmochimica Acta, 58(17), 3653–3664.
41. Poe, B. T., Mcmillan, P. F., Coté, B., Massiot, D., & Coutures, J. P. (1994). Structure and
dynamics in calcium aluminate liquids: High-temperature 27Al NMR and Raman spectroscopy.
Journal of the American Ceramic Society, 77(7), 1832–1838.
42. Bowker, J., Lupis, C., & Flinn, P. (1981). Structural studies of slags by Mössbauer spectroscopy.
Canadian Metallurgical Quarterly, 20(1), 69–78.
43. Pargamin, L., Lupis, C., & Flinn, P. (1972). Mössbauer analysis of the distribution of iron
cations in silicate slags. Metallurgical Transactions, 3(8), 2093–2105.
44. Mysen, B., & Dubinsky, E. (2004). Mineral/melt element partitioning and melt structure.
Geochimica et Cosmochimica Acta, 68, 1617–1634.
45. Duer, M. J. (2005). Introduction to solid-state NMR spectroscopy. Wiley-Blackwell.
46. Neuville, D. R., Cormier, L., Flank, A.-M., Briois, V., & Massiot, D. (2004). Al speciation and
Ca environment in calcium aluminosilicate glasses and crystals by Al and Ca K-edge X-ray
absorption spectroscopy. Chemical Geology, 213(1), 153–163.
47. Allwardt, J. R., Lee, S. K., & Stebbins, J. F. (2003). Bonding preferences of non-bridging
O atoms: Evidence from 17O MAS and 3QMAS NMR on calcium aluminate and low-silica
Ca-aluminosilicate glasses. American Mineralogist, 88(7), 949–954.
48. Sajid, M., Bai, C., Aamir, M., You, Z., Yan, Z., & Lv, X. (2019). Understanding the structure
and structural effects on the properties of blast furnace slag (BFS). ISIJ International, 59(7),
1153–1166.
49. Mysen, B. O., Virgo, D., & Seifert, F. A. (1982). The structure of silicate melts: Implications for
chemical and physical properties of natural magma. Reviews of Geophysics, 20(3), 353–383.
50. Mysen, B. O. (1990). Relationships between silicate melt structure and petrologic processes.
Earth-Science Reviews, 27(4), 281–365.
51. Mysen, B. O., Ryerson, F. J., & Virgo, D. (1980). The influence of TiO2 on the structure and
derivative properties of silicate melts. American Mineralogist, 65(11–12), 1150–1165.
52. Seifert, F. A., Mysen, B. O., & Virgo, D. (1982). Three-dimensional network structure in
the systems SiO2 -NaA1 O2 , SiO2 -CaA12 O4 and SiO2 -MgA12 O4 . American Mineralogist, 67,
696–711.
References 75

53. Jiang, G., Wu, Y., You, J., Hou, H., & Chen, H. (2004). A discussion on the microstructural
units of aluminosilicate melts. Acta Petrologica Sinica, 20(3), 753–758.
54. Lee, S. K., & Stebbins, J. F. (1999). The degree of aluminum avoidance in aluminosilicate
glasses. American Mineralogist, 84(5–6), 937–945.
55. Stebbins, J. F., & Zhi, X. (1997). NMR evidence for excess non-bridging oxygen in an
aluminosilicate glass. Nature, 390(6655), 60–62.
56. Lacy, E. D. (1963). Aluminum in glasses and melts. Physics and Chemistry of Glasses, 4, 234.
57. Stebbins, J. F., Oglesby, J. V., & Kroeker, S. (2001). Oxygen triclusters in crystalline CaAl4O7
(grossite) and in calcium aluminosilicate glasses: 17O NMR. American Mineralogist, 86(10),
1307–1311.
58. Iuga, D., Morais, C., Gan, Z., Neuville, D. R., Cormier, L., & Massiot, D. (2005). NMR
heteronuclear correlation between quadrupolar nuclei in solids. Journal of the American
Chemical Society, 127(33), 11540–11541.
59. Park, J. H., Min, D. J., & Song, H. S. (2004). Amphoteric behavior of alumina in viscous flow
and structure of CaO-SiO2 (-MgO)-Al2 O3 slags. Metallurgical and Materials Transactions B,
35(2), 269–275.
60. Sun, C. Y., Liu, X. H., Li, J., Yin, X. T., Song, S., & Wang, Q. (2017). Influence of Al2 O3 and
MgO on the viscosity and stability of CaO-MgO-SiO2 -Al2 O3 slags with CaO/SiO2 =1.0. ISIJ
International, 57(6).
61. Sun, Y., Wang, H., & Zhang, Z. (2018). Understanding the relationship between structure
and thermophysical properties of CaO-SiO2 -MgO-Al2 O3 molten slags. Metallurgical and
Materials Transactions B, 49(2), 677–687.
62. Zheng, K., Zhang, Z., Yang, F., & Sridhar, S. (2012). Molecular dynamics study of the structural
properties of calcium aluminosilicate slags with varying Al2 O3 /SiO2 ratios. ISIJ International,
52(3), 342–349.
63. Yan, L. (2013). Theory and practice of molecular dynamics simulation. Science Press.
64. Alder, B. J., & Wainwright, T. E. (1957). Phase transition for a hard sphere system. Journal of
Chemical Physics, 27(5), 1208–1209.
65. Rahman, A. (1964). Correlations in the motion of atoms in liquid argon. Physical Review,
136(2A), A405–A411.
66. Verlet, L. (1968). Computer “experiments” on classical fluids. II. Equilibrium correlation
functions. Physical Review, 165(1), 201.
67. Lees, A. W., & Edwards, S. F. (1972). The computer study of transport processes under extreme
conditions. Journal of Physics C Solid State Physics, 5(15), 1921–1928.
68. Andersen, H. C. (1980). Molecular dynamics simulations at constant pressure and/or temper-
ature. The Journal of Chemical Physics, 72(4), 2384–2393.
69. Parrinello, M., & Rahman, A. (1981). Polymorphic transitions in single crystals: A new
molecular dynamics method. Journal of Applied Physics, 52(12), 7182–7190.
70. Gillan, M., & Dixon, M. (1983). The calculation of thermal conductivities by perturbed
molecular dynamics simulation. Journal of Physics C: Solid State Physics, 16(5), 869.
71. Nosé, S. (1984). A unified formulation of the constant temperature molecular dynamics
methods. The Journal of Chemical Physics, 81(1), 511–519.
72. Car, R., & Parrinello, M. (1985). Unified approach for molecular dynamics and density-
functional theory. Physical Review, 55(22), 2471.
73. Çağin, T., & Pettitt, B. M. (1991). Molecular dynamics with a variable number of molecules.
Molecular Physics, 72(1), 169–175.
74. Tandia, A., Timofeev, N. T., & Mauro, J. C. (2011). Defect-mediated self-diffusion in calcium
aluminosilicate glasses: A molecular modeling study. Journal of Non-crystalline Solids, 357(7),
1780–1786.
75. Liang, D., Yan, Z., Lv, X., Zhang, J., & Bai, C. (2017). Transition of blast furnace slag from
silicate-based to aluminate-based: Structure evolution by molecular dynamics simulation and
Raman spectroscopy. Metallurgical and Materials Transactions B, 48(1), 573–581.
76. Petkov, V. V., Billinge, S. J., Shastri, S. D., & Himmel, B. (2000). Polyhedral units and network
connectivity in calcium aluminosilicate glasses from high-energy X-ray diffraction. Physical
Review Letters, 85(16), 3436–3439.
76 3 Slag Structure of High Alumina Blast Furnace Slag

77. Losq, C. L., Neuville, D. R., Florian, P., Henderson, G. S., & Massiot, D. (2014). The role of
Al3+ on rheology and structural changes in sodium silicate and aluminosilicate glasses and
melts. Geochimica et Cosmochimica Acta, 126(2), 495–517.
78. Stebbins, J. F., Kroeker, S., Lee, S. K., & Kiczenski, T. J. (2000). Quantification of five- and six-
coordinated aluminum ions in aluminosilicate and fluoride-containing glasses by high-field,
high-resolution 27 Al NMR. Journal of Non-crystalline Solids, 275(1–2), 1–6.
Chapter 4
High-Temperature Physicochemical
Properties of High Alumina Slag

Modern blast furnace ironmaking technology has a history of more than 100 years,
and the impact of the high-temperature physicochemical properties of blast furnace
slag on the process has been well-known by the metallurgists. Slag is a complex
oxides mixing system, and the influence of each component on its properties is not
monotonous and linear. The use of high alumina iron ore may cause the type of blast
furnace slag to gradually transition from the present silicate system to the alumi-
nosilicate system. However, the current understanding of traditional blast furnace
slag is based on the physical chemistry of silicate melt, and the aluminosilicate
system would show varying degrees of inapplicability. In this chapter, the important
high-temperature physicochemical properties of the aluminosilicate slag system and
its relationship between them and the structure are systematically discussed.

4.1 Liquids Temperature and Fluidity Temperature

According to the CaO–SiO2 –MgO–Al2 O3 phase diagram with different alumina


content (10–25 wt%), as shown in Fig. 4.1. With the increase of Al2 O3 content, the
high melting point phase spinel (MgO · Al2 O3 ) area expands. When the basicity is
about 1.0–1.2 (this is the common basicity range of blast furnace slag) with suitable
content of MgO and Al2 O3 , the primary precipitated phase is melilite. With the
increase of MgO content, the slag composition also approaches the spinel phase
region. Peng et al. [1] and Shen [2] also confirmed that the liquids temperature of
the slag increased with the increase of Al2 O3 content at fixed basicity and MgO
content, as well as the fluidity temperature. Generally, in order to make the liquids
temperature of the slag meet the smelting requirements, the possible route is either
reduce the content of alumina in the slag or add flux, normally with the suitable
MgO.

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 77
X. Lv and Z. Yan, High Temperature Physicochemical Properties of High Alumina Blast
Furnace Slag, https://doi.org/10.1007/978-981-19-3288-5_4
78 4 High-Temperature Physicochemical Properties …

Fig. 4.1 Phase diagrams with liquidus of CaO–SiO2 –MgO–Al2 O3 system with 10–25 wt% Al2 O3

4.2 Viscosity

The viscous behavior of BF slag plays an essential role in determining the perfor-
mance and productivity of the ironmaking process. Many researchers have studied
the effect of alumina on the viscosity of the BF slag system, which is summarized
in Table 4.1 and Fig. 4.2. As early as 1987, Kou et al. [3] pointed out that in the
CaO–Al2 O3 –SiO2 system with fixed basicity, the increase of alumina content would
increase the viscosity of the slag. For the CaO–MgO–Al2 O3 –SiO2 system, Peng
et al. [1], Tang et al. [4], Yao et al. [5], and Kim et al. [6] reported that higher
alumina content increased the slag viscosity due to the higher polymerization degree
of the structures. For the initially formed blast furnace slag with FeO, conducted
by Lee et al. [7] it was also found that the slag viscosity increased with increasing
alumina content with the fixed basicity. Most of these studies have pointed out that
the increase of alumina has an adverse effect on fluidity. However, in the CaO–MgO–
Al2 O3 –SiO2 system studied by Park et al. [8] and Sun et al. [9], the viscosity first
increases and then decreases when the basicity and MgO content are fixed. This
result is inconsistent with others and was interpreted as that the alumina may behave
4.2 Viscosity 79

Table 4.1 Effect of Al2 O3 on the viscosity of BF slags


References Composition Conclusions
Kou et al. [3] CSA: R2 = 0.8, 1.0, 10–30 wt% With fixed basicity and other
Al2 O3 components content, the slag
Peng et al. [1] CSAM: R2 = 0.95, 1.05, 1.19, viscosity increases with the
17–20 wt% Al2 O3 , 12 wt% MgO increase of Al2 O3 content
Tang et al. [4] CSAM: R2 = 0.5–0.9, 5–20 wt%
Al2 O3
Yao et al. [5] CSAM: R2 = 1.1–1.2, 15–20 wt%
Al2 O3 , 9–11 wt% MgO
Kim et al. [6] CSAM: R2 = 0.8–1.3; 10 wt%
MgO, 5–20 wt% Al2 O3
Lee et al. [7] CSAMF: R2 = 1.15–1.6,
10–13 wt% Al2 O3 , 5–10 wt% MgO
and 5–20 wt% FeO
Park et al. [8] CSAM: R2 = 1.0, 1.3; 0–15 wt% With fixed basicity and other
Al2 O3 components’ content, the slag
Sun et al. [9] CSAM: R2 = 1.0, 14–17 wt% viscosity first increases and then
Al2 O3 , 5–15 wt% MgO decreases with increasing Al2 O3
content
The turning points are:
10 wt% Al2 O3 (Park et al.);
15 wt% Al2 O3 (Sun et al.);
Liao et al. [10] CSAM: 29.5 wt% CaO, 15 wt% Al2 O3 replaces SiO2 , resulting in
MgO 5–23 wt% Al2 O3 a decrease in the viscosity of the
slag
Machin et al. [11] CSAM: 35 wt% SiO2 , 15 wt% MgO Al2 O3 replaces MgO, resulting in
an increase in the viscosity of the
slag
Nakamoto et al. [12] CSAM: 5–35 wt% The slag viscosity (35 wt%
Al2 O3 –CaO–MgO–SiO2 Al2 O3 –43 wt% CaO–7.5 wt%
MgO–14.4 wt% SiO2 ) at 1673 K
(1400 °C) was lower than 0.6 Pa s
that satisfies the fluidity in a blast
furnace operation

as an amphoteric oxide in range of the slag composition studied. According to Lewis


acids and bases, the acidity of alumina is greater than that of MgO and less than
that of SiO2 . The acidic oxides can consume the free oxygen ions in the slag and
increase the degree of polymerization of the slag, resulting in the increase of slag
viscosity. Therefore, replacing alumina with silica would increase the slag viscosity,
and replacing alumina with MgO will reduce the viscosity of the slag, as reported
by Liao et al. [10] and Machin et al. [11]. According to the phase diagrams for the
CaO–MgO–SiO2 –Al2 O3 system, Nakamoto et al. [12] proposed a slag system with
35 wt% Al2 O3 that the viscosity was lower than 0.6 Pa s at 1400 °C, which meets the
fluidity requirement in blast furnace operation. Although this slag system has not yet
80 4 High-Temperature Physicochemical Properties …

Fig. 4.2 Effect of Al2 O3 on the viscosity of the BF slag systems

been implemented in industry, it provides a new idea to the metallurgical technicians


for utilization of high alumina iron ore on large scale. Despite the numerous studies
carried out in the past, the mechanisms for the viscosity of high alumina slags have
yet to be fully understood.

4.2.1 Effect of Al2 O3 Content

The effect of Al2 O3 on the studied slag (CaO–9 wt% MgO–SiO2 –Al2 O3 –1 wt%
TiO2 ) at fixed basicity is shown in Fig. 4.3. The increase in temperature can make
the melt obtain more energy to overcome the viscous activation energy, and the
strengthening of the thermal vibration relaxes the network structure of the slag. At
the fixed temperature, the viscosity increases first and then decreases slightly with
increasing Al2 O3 content with a turning point near 24 wt%. This result with a turning
point is similar to that of Park et al. [8] and Sun et al. [9] work, which might be
understood from the amphoteric characteristics of Al2 O3 . When the Al2 O3 content
is greater than 24 wt%, the viscosity cannot be accurately measured at a temperature
lower than 1460 °C due to the solid phase precipitation. The thermal stability of the
slag also decreases since the viscosity difference value for every 20 °C increases
from 0.7 to 1.2 dPa s.
The fluidity of slag is closely related to its structure. According to the theory of
polymerization structure, the slag has a similar structure to the crystals in a short-
range, showing a state of short-range order and long-range disorder. The distance and
4.2 Viscosity 81

Fig. 4.3 Effect of Al2 O3


content on the viscosity [13].
Reprinted from Ref. [13]

force of the structural units in the slag are similar to their corresponding crystals, and
each structural unit is under the interaction force of adjacent structural units, that is,
in a certain potential barrier. Once the molten slag flows, its corresponding structural
unit has to overcome the potential barrier. Then the activation energy should be the
minimum energy necessary for the movement of the liquid structural unit, and the
Arrhenius equation (Eq. 1.2) is usually used to calculate the slag activation energy.
In the aluminosilicate slag, the size of the Six Oy z− and the Alx Oy z− is much larger
than that of the cations, so the energy required for its movement is large as well.
Therefore, these composite anions become the main structural units in the slag that
restricts the movement of the slag. The change in the slag composition leads to the
change of the composite anions concentration, polymerizes, or depolymerizes of the
structure, resulting in the change of the slag viscosity.
According to Eq. 1.2, the linear relationship between ln η and 1/T can be obtained,
and the viscous activation energy of the slag can be calculated through the slope.

ln η = ln A + E η /RT (4.1)

here, A, E η , R, and T are a constant, viscous activation energy, the universal gas
constant, and absolute temperature, respectively. Figure 4.4 shows the relation-
ship between viscous activation energy and Al2 O3 content of CaO–9 wt% MgO–
SiO2 –Al2 O3 –1 wt% TiO2 slag system. The viscous activation energy of slag shows
an increasing trend with the increase of Al2 O3 content. Since the viscosities are
measured at a completely molten state, the change in the viscous activation energy
of the slag is mainly caused by the slag structure.
According to the viscosity-temperature (η–T ) curve, as shown in Fig. 4.5a, the
viscosity of slag suddenly increases when the temperature drops to a certain temper-
ature, showing the characteristics of short slag. The possible reason is that the precip-
itation of the solid phase in the slag during the cooling [14, 15]. According to the
fixed slope method, as mentioned in Chap. 1, the fluidity temperature (T flu ) can
82 4 High-Temperature Physicochemical Properties …

Fig. 4.4 Effect of Al2 O3 on


the activation energy.
Reprinted from Ref. [13]

Fig. 4.5 Effect of Al2 O3 on the η–T curves (a), and fluidity temperature (b). Reprinted from Ref.
[13]

be obtained and plotted in Fig. 4.5b. With the increase of Al2 O3 content, the slag
T flu gradually increases, and the stability of slag structure decreases. Ilyushechkin
et al. [14] pointed out that only when the amount of solid phase exceeds a certain
level the viscosity would increase sharply, the size and morphology of the precipi-
tated solid particles also affect the fluidity temperature. The viscosity-temperature
curve essentially reflects the adjustment process of the slag structure with tempera-
ture: when the temperature is higher than T flu , the slag structure slightly adjusts with
decreasing temperature, and the viscosity of the slag decreases relatively small; when
the temperature is lower than T flu , the disordered liquid slag gradually transforms
into an ordered structure, and the network structure of molten slag undergone with
major changes. According to the phase diagram (Fig. 4.1), with the increase of Al2 O3
content, the primary precipitation temperature increases, and the composition point
moves toward the liquids.
4.2 Viscosity 83

4.2.2 Effect of Al2 O3 ↔ SiO2 Substitution

The effect of substitution SiO2 with Al2 O3 on the viscosity is shown in Fig. 4.6.
Increasing (wt% Al2 O3 )/(wt% SiO2 ) from 0.47 to 0.92 causes a slight decrease in
the viscosity of the slags, and an opposite effect was found when [(wt% Al2 O3 )/(wt%
SiO2 )] is more than 0.92. The distance between each isothermal viscosity line in the
figure does not change much, indicating that (wt% Al2 O3 )/(wt% SiO2 ) has little
effect on the thermal stability of slag viscosity. The relationship between viscous
activation energy and the (wt% Al2 O3 )/(wt% SiO2 ) ratio is shown in Fig. 4.7. With
the increase of (wt% Al2 O3 )/(wt% SiO2 ) ratio, the viscous activation energy of slag
first decreases and then increases with a valley value at 0.92, which is similar to that
of viscosity.

Fig. 4.6 Effect of (wt%


Al2 O3 )/(wt% SiO2 ) on the
viscosity. Reprinted from
Ref. [13]

Fig. 4.7 Effect of (wt%


Al2 O3 )/(wt% SiO2 ) on the
activation energy. Reprinted
from Ref. [13]
84 4 High-Temperature Physicochemical Properties …

Fig. 4.8 Effect of (wt% Al2 O3 )/(wt% SiO2 ) on the η–T curves (a), and free-running temperature
(b). Reprinted from Ref. [13]

The effect of (wt% Al2 O3 )/(wt% SiO2 ) on the viscosity-temperature (η–T ) curve
and fluidity temperature are shown in Fig. 4.8. The viscosity of the slag gradually
increases with the decrease in temperature. The fluidity temperature of slag increases
first and then decreases with increasing (wt% Al2 O3 )/(wt% SiO2 ) ratio, which is
opposite to the change of viscosity and viscous activation energy. According to
liquids in Fig. 4.3, when the (wt% Al2 O3 )/(wt% SiO2 ) ratio is near 0.92, the slag
composition is close to the liquidus line (1450 °C), causing the increase of fluidity
temperature.

4.2.3 Relationship Between Slag Viscosity and Its Structure

Temperature and the chemical composition of the slag are the main factors that affect
the viscosity. The increase in temperature causes the viscosity to decrease mainly
due to two aspects. On the one hand, the high temperature strengthens the thermal
vibration of the ions in the slag, increases the distance and weakens the force between
the ions, and reduces the viscous resistance of ion migration, thereby reducing the
viscosity. On the other hand, the network modifier MO reaches the ionization equi-
librium (MO → M2+ + O2− ) in the molten state. The increase in temperature makes
the equilibrium forward to the right, which can provide more free oxygen for the
depolymerization of the network and decrease the viscosity. The influence of slag
chemical composition on viscosity is mainly due to the change in slag structure. As
shown in Fig. 4.9, the Si4+ and Al3+ ions in the slag form SiO4 4− and AlO4 5− tetra-
hedral structural units, and these tetrahedral structural units are connected to each
other by TOT (T = Si, Al) bonds to form a composite anion (Tx Oy z− ). The metal
cations such as Ca2+ and Mg2+ are located between the composite anions as network
modifiers. The relative movement of anions and cations (or cations clusters) in the
slag needs to overcome the interforce, the internal anions and cations (or cations
clusters) of the molten slag need to overcome the force during movement, and the
4.2 Viscosity 85

Fig. 4.9 The sketch of the complex anion group in the molten slag

macroscopic manifestation of internal friction generated in the relative movement is


the viscosity of the molten slag. Therefore, the viscosity of the slag is related to the
relative movement resistance between the anions and cations (or cations clusters) in
the slag.
The connection between two composite anions A and B determines the magnitude
of the relative movement interforce, as illustrated in Fig. 4.10. The C which is a
network former (such as Si, Al), connects A and B through bridging oxygens to form
a larger composite anion, so the resistance that needs to be overcome to make the
larger composite anion migrate is obviously greater than that of the smaller ones.
Of course, relative movement may also occur between A and B, which is related

Fig. 4.10 The rheological mechanism of molten slag


86 4 High-Temperature Physicochemical Properties …

to the strength of the bridging oxygen bond. According to the molecular dynamics
results (Table 3.7), the Al–O bond length (1.75 Å) is greater than the Si–O bond
length (1.61 Å), so the relative movement resistance of Al connecting AB is smaller
than that of Si connecting AB. The C which acts as a network modifier (such as
Ca2+ , Mg2+ ), connects A and B through non-bridging oxygens, due to the longer
non-bridging oxygen bond length and relatively weaker force, the shear force for the
relative movement of the A and B composite anions is small. This is the main reason
for the decrease of viscosity after the network modifier is added to the slag.
The effect of Al2 O3 content on slag viscosity (Fig. 4.3) shows a turning point
at 24 wt%, which may be the manifestation of the amphoteric effect of Al2 O3 . As
mentioned above. It can be expressed as an acidic oxide acting as a network former,
polymerizing the structure and increasing the slag viscosity. Also, it can be expressed
as a basic oxide acting as a network modifier, depolymerizing slag structure to reduce
the viscosity. However, combining the molecular dynamics calculation with Raman
spectroscopy results, the degree of polymerization of the slag did not decrease. Since
there are sufficient charge compensation ions in the slag, the four-coordinated A13+
is dominant, according to the research of Merzbacher [16], Neuville et al. [17–19].
The remaining cations should act as network modifiers in the slag, indicating that
the decrease in viscosity is not a manifestation of the amphoteric effect of Al2 O3 .
The increase in viscosity is due to the increase of Al2 O3 content at a fixed basicity,
the CaO content in the slag gradually decreased. In addition, due to the increase of
AlO4 5− tetrahedron, more cations are needed to act as charge compensators, resulting
in a decrease in network modifiers and an increase in the degree of polymerization
of the slag. When the Al2 O3 content is more than 24 wt%, the number of Al3+ ions
in the slag exceeds the number of Si4+ ions, and the AlO4 5− tetrahedron becomes the
dominant units in the slag. Since the relative movement resistance of Al–O is weaker
than that of Si–O, although the degree of polymerization of the slag increases, the
viscosity is slightly decreased.
The effect of substituting SiO2 with Al2 O3 on the viscosity (Fig. 4.7) shows a
minimum when the (wt% Al2 O3 )/(wt% SiO2 ) ratio is 0.92. According to Fig. 3.16.
Replacing SiO2 with Al2 O3 reduces the degree of polymerization of the slag first
and then increases, which is consistent with the changing trend of viscosity. In the
case of fixed network modifier content with sufficient charge compensator, SiO2 ↔
Al2 O3 substitution would cause the SiO4 4− tetrahedron in the slag to be replaced
by AlO4 5− . Because the strength of the Al–O bond is weaker than that of the Si–O
bond, the viscosity should be lower. However, the formation of AlO4 5− tetrahedral
requires the consumption of network modifiers for charge compensation, and the slag
viscosity increases. The competition of these two factors results in the slag viscosity
first decreases and then increases. It should be noted that when (wt% Al2 O3 )/(wt%
SiO2 ) exceeds 0.79, the number of Al3+ in the slag exceeds the number of Si4+
and the slag changes from silicate to aluminate. In the silicate melt system, the
substitution of Al2 O3 for SiO2 causes the decrease in slag polymerization, while in
the aluminate melt system, the substitution of Al2 O3 for SiO2 increases the degree
of polymerization and increase the viscosity. The changing trend is consistent with
the results of molecular dynamics simulation.
4.3 Density 87

4.3 Density

Density is one of the important physical properties of metallurgical slag, but there
are few studies on the influence of alumina on the density of blast furnace slag. As
shown in Fig. 4.11, Sokolov et al. [20] and Rajavaram et al. [21, 22] reported that
the Al2 O3 decreases the density of CaO–SiO2 –Al2 O3 system with fixed basicity at
low alumina content and increases the density at high alumina content. Winterhager
and Kamnel [23] and Barrett and Thomas [24] found that the density increases with
increasing Al2 O3 content at fixed CaO content in CaO–SiO2 –Al2 O3 system.
The authors of the book investigated the effect of Al2 O3 content and SiO2 ↔
Al2 O3 substitution on the slag density via buoyancy method (Archimedes principle),
as shown in Fig. 4.12. Due to thermal expansion, the higher temperature, the lower
density. With the increase of Al2 O3 content, the slag density has a minimum when
the Al2 O3 content is about 24 wt%, which is similar to the results of Sokolov et al.
[20] and Rajavaram et al. [21, 22]. Substituting Al2 O3 for SiO2 , the density increases
significantly when (wt% Al2 O3 )/(wt% SiO2 ) ratio increases from 0.47 to 0.92, and
then the density gradually decreases. In the work of Winterhager and Kamnel [23]
and Barrett and Thomas [24], the turning point was not found. At the blast furnace
operating temperature, the density ranges from 2.6 to 2.7 g/cm3 .
The relationship between density, slag degree of polymerization (DOP), (Q2 +
Q3 )/(Q0 + Q1 ) ratio, and Al/Si atom ratio are shown in Fig. 4.13, where the degree
of polymerization and (Q2 + Q3 )/(Q0 + Q1 ) ratio is calculated from the Raman
spectroscopy. For the studied slag system with a fixed basicity, the density decreases
with the increase of Al/Si atom ratio in the silicates range, while the DOP and (Q2

Fig. 4.11 Effect of Al2 O3 on the density of the CaO–SiO2 –Al2 O3 slag systems
88 4 High-Temperature Physicochemical Properties …

Fig. 4.12 Effect of Al2 O3 (a) and (wt% Al2 O3 )/(wt% SiO2 ) (b) on the density

+ Q3 )/(Q0 + Q1 ) ratio gradually increase. In the aluminates range (Al/Si atom ratio
> 1), the density increases, and (Q2 + Q3 )/(Q0 + Q1 ) ratio decreases. In the case of
SiO2 ↔ Al2 O3 substitution for the slag with fixed CaO content (40 wt%), increasing
the Al/Si atomic ratio from 0.66 to 0.92 caused an increase in the density and a
decrease in the (Q2 + Q3 )/(Q0 + Q1 ) ratio and the DOP, but it has the opposite effect
in the aluminates range. An increase in the degree of polymerization means that the
composite anions are more closely connected, which leads to an increase in density.
While on the other hand, since the Al–O bond is longer than the Si–O bond, the
increase in alumina increases the molar volume of the slag and reduces the density.
The competition between these two aspects is the main reason for the change in slag
density.

4.4 Surface Tension

The surface tension of slag plays an important role in the gas-slag-metal reactions
and affects the separation between slag and metal, the formation of new phase, and
the slag foaming. The effect of Al2 O3 on the surface tension of blast furnace slag is
relatively clear. The effect of Al2 O3 content and SiO2 ↔ Al2 O3 substitution on the
surface tension of studied slag via the maximum bubble pressure method (MBP),
are shown in Fig. 4.14. The surface tension increases with the increase of Al2 O3
content and (wt% Al2 O3 )/(wt% SiO2 ). Mukai and Ishikawa [26] studied the surface
tension of the CaO–SiO2 –Al2 O3 system with fixed basicity by using the sessile drop
method and reported the similar conclusion. Tian et al. [27] and Parmelee et al.
[28] found that alkaline oxides (Na2 O, Li2 O, etc.) can reduce the surface tension of
aluminosilicate slag. Fenzke [29] and Kim et al. [30] found that replacing SiO2 with
Al2 O3 can effectively increase the surface tension of the slag. In addition, the surface
tension increases as the temperature decreases.
From the point of force, the surface tension is that the ions or clusters on the
slag surface are attracted by the bulk rather than the force from the gas, and the
4.4 Surface Tension 89

Fig. 4.13 Density, (Q2 + Q3 )/(Q0 + Q1 ), and DOP against the Al/Si atomic ratio [25]. Reprinted
from Ref. [25]
90 4 High-Temperature Physicochemical Properties …

Fig. 4.14 Effect of Al2 O3 (a) and (wt% Al2 O3 )/(wt% SiO2 ) (b) on the density. Reprinted from
Ref. [25]

resultant force is along the tangent direction of the slag surface. Therefore, the slag
has a tendency to reduce its surface area. The interaction force between ions is
mainly electrostatic force. From the point of energy, in the bulk of the melts, each
atom is pulled equally in every direction by neighboring atoms, which can stabilize
the chemical environment around an atom. The atoms at the surface have fewer
neighbors, resulting in many unsatisfied bonds with high energies. This forces the
melt surfaces to contract to the minimal area to decrease the high energy of the
melt surface. Therefore, the increase in the number of unsatisfied bonds leads to an
increase in the surface tension of the melts.
For silicate slag, the Six Oy z− is the main composite anions in the slag. With
the increase of SiO2 content, the increased degree of polymerization increases the
ionic radius of the composite anions, resulting in the decrease of the electrostatic
force between the composite anions and the cations. When the electrostatic force
becomes weak, the surface tension decreases. Once Al2 O3 is added to the slag,
AlO4 5− tetrahedron replaces SiO4 4− tetrahedron to form new composite anions, due
to the charge compensation effect of AlO4 5− tetrahedron, the cations act as charge
compensators in the composite anions, leading to an increase in electrostatic force,
as well as the surface tension. Kingery [31] suggested that the liquid surface of oxide
melts is covered with oxygen atoms. Figure 4.15 presents schematic illustrations
of the different types of oxygen atoms in the slag. The increase in Al2 O3 content
leads to an increase of AlO4 5− tetrahedron near the slag surface. Then the charge
compensators cations and unsatisfied bonds on the surface increased, resulting in an
increase in the slag surface tension.
4.5 Sulfide Capacity 91

Fig. 4.15 Schematic illustrations of the different types of oxygen in the bulk and slag surface.
Reprinted from Ref. [25]

4.5 Sulfide Capacity

Sulfur has an adverse effect on the mechanical properties of steel products, including
strength, ductility, and toughness. Therefore, reducing the sulfur content in steel prod-
ucts is an important topic in the smelting process. Slag plays an important role in
desulfurization, especially in the blast furnace process. Nzotta et al. [32–36] system-
atically measured the sulfide capacity slags from binary, ternary, to multicomponent
systems, and found that the sulfide capacity increases with increasing temperature,
basicity, and basic oxide content. Moreover, the replacement of SiO2 with Al2 O3 has
no considerable effect on the desulfurization ability of the CaO–MgO–SiO2 –Al2 O3
slag system [32]. Park found that the SiO2 ↔ Al2 O3 substitution does not affect
the sulfide capacity in the CaO–MnO–MgO–SiO2 –Al2 O slag with high content of
CaO and MnO. Drakaliysky et al. investigated the sulfide capacities of a high Al2 O3
slag system and showed that higher alumina content and slag basicity lead to higher
sulfide capacities at a given Cr2 O3 content in the slag. Karsrud [37] reported that
replacement CaO with MgO decreased the sulfide capacity at fixed Al2 O3 content.
Sommerville [38] concluded the desulfurization capabilities of Al2 O3 as that it is
weaker than that of MnO, CaO, MgO, and FeO, but stronger than that of SiO2 .
Zhang [39], Shankar [40, 41], and Taniguchi et al. [42] investigated the slag with
high alumina, but the influence of alumina was not definitely clear.

4.5.1 The Effect of Al2 O3 Content and SiO2 ↔ Al2 O3


Substitution on the Slag Sulfide Capacity

The effect of Al2 O3 and substituting SiO2 with Al2 O3 on the sulfide capacity of
the studied slag system was measured at 1500 °C using a metal-slag equilibration
method.
The influence of Al2 O3 content and (wt% Al2 O3 )/(wt% SiO2 ) ratio on the slag
sulfur capacity and sulfur distribution are shown in Fig. 4.16. with the increase of
Al2 O3 content, the sulfur distribution ratio and sulfur capacity of the slag gradually
92 4 High-Temperature Physicochemical Properties …

Fig. 4.16 Effect of Al2 O3 (a) and (wt% Al2 O3 )/(wt% SiO2 ) (b) on the sulfide capacity

decrease. Increasing the (wt% Al2 O3 )/(wt% SiO2 ) ratio from 0.47 to 0.79 caused
a significant increase in the sulfide capacity, while the sulfide capacity increased
slightly when the (wt% Al2 O3 )/(wt% SiO2 ) ratio is more than 0.79.
Thermodynamically, sulfide ions in slags react with metal cations to form stable
compounds, such as CaS or MgS. The sulfide affinity for CaO and MgO is expressed
by Eqs. 4.2 and 4.3, respectively. The equilibrium constant of the MO ↔ MS (M =
Ca, Mg) reaction can be used to compare the contribution of different cations to the
sulfur capacity. The equilibrium constant of MgO ↔ MgS equilibrium at 1500 °C
is much smaller than that of CaO ↔ CaS equilibrium, indicating that the sulfide is
thermodynamically more stable in the form of CaS due to the higher affinity for Ca
than Mg [43, 44]. Therefore, the slag sulfur capacity has a great relationship with
the activity of CaO (aCaO ).

(CaO) + [S] = (CaS) + [O];


r G θm,CaS = 105,784.6 − 28.723T (J/mol); K CaS = 2.42 × 10−2 (4.2)

(MgO) + [S] = (MgS) + [O];


 
J
r G θm,MgS = 203,604.6 − 35.023T ; K MgS = 6.77 × 10−5 (4.3)
mol

where K MS is the equilibrium constant of reactions and the mole fraction of MS; the
equilibrium constant has been calculated by the standard free energy at 1500 °C.
The relationship between the sulfide capacity of the CaO–SiO2 –MgO–Al2 O3 –
(TiO2 ) slag systems and the activity of CaO in the present work and other researchers’
work [40, 45–47], as shown in Fig. 4.17a. The sulfide capacities of the slag systems
increased with increasing activity of CaO. The CaO behaves as a basic oxide in the
slag, providing the free oxygen by the dissolution reaction expressed by Eq. 4.4.

(CaO) = Ca2+ + O2− (4.4)

According to the Eq. 1.6,


4.5 Sulfide Capacity 93

Fig. 4.17 Dependence of the log(C s ) on the log(aCaO ) (a), and log(C s ) on the log(aCaO /γ CaS )
(b) [48]. Reprinted from Ref. [48]

   
log(Cs ) = log K θ + log aO2− /γS2− (4.5)

where K θ is the equilibrium constant of the reaction. aO2− is the activity of free
oxygen ions. γS2− is the activity coefficient of sulfide ions.
Assuming that the activity of CaO (aCaO ) represents the activity of free oxygen
ions in the slag (aO2− ), and the activity coefficient of sulfide ions in the slag (γS2− ) is a
constant, the slope in Fig. 4.17a is a theoretical slope of unity. X CaS can be calculated
from sulfur in slag, most probably to exist in the form of CaS. However, the slope
of the logarithm of the sulfide capacity can be revised as a function of aCaO /γ CaS ,
because γ CaS depends on the slag composition and is not constant. As shown in
Fig. 4.17b. A good linear agreement presents between log(C s ) and log(aCaO /γ CaS ),
and the slope is approximately unity. According to the intercept, the equilibrium
constant log(K θ ) can be calculated, which is −2.285.
The aCaO and the γ CaS as a function of the Al/Si atomic ratio are shown in Fig. 4.18.
The aCaO and the γ CaS both increase by the substitution of Al2 O3 for SiO2 , while
the aCaO decreases and the γ CaS increases with increasing Al2 O3 content at a fixed
basicity. It is unambiguous that the effect of SiO2 on decreasing the activity of CaO
is more significant than that of Al2 O3 . In addition, Woo and Lee [49] reported that
the γ CaS increases with increasing Al2 O3 content, which is in good agreement with
the present work. According to Eq. 4.5, the decrease in oxygen activity (aO2− ) and
increase in the activity coefficient of sulfide (γS2− ) result in the decrease of sulfide
capacity. This result might be a response to the decrease in sulfide capacity when the
Al/Si atomic ratio increases at a fixed basicity. In the aluminate-based slag range,
the increase in both aO2− and γS2− in a similar degree by increasing the Al/Si atomic
ratio at a fixed CaO content, while the increased level of aO2− is greater than γS2− in
the silicate-based slag system.
94 4 High-Temperature Physicochemical Properties …

Fig. 4.18 The activity of CaO and activity coefficient of CaS against the Al/Si atomic ratio.
Reprinted from Ref. [48]

4.5.2 Relationship Between Sulfide Capacity and Structure

Fincham and Richardson [50] investigated oxygen atoms participating in the desul-
furization in aluminosilicate slags and figured out that the O2− replaced with S2−
should not be connected with the tetrahedron. In other words, the sulfide capacity
in melts increases with increasing free oxygen (FO) concentration. Figure 4.19
show the sulfide capacity, structure complexity, and FO concentration against the
Al/Si atomic ratio, in which the structure complexity ratio was determined from the
deconvolution of the Raman spectra, and the FO concentration was calculated by
molecular dynamics simulation. The sulfide capacity and FO concentration decrease
with increasing Al/Si atomic ratio, while the structure of the melts becomes more
complicated. It seems likely that the effect of melt structure and FO concentration on
the sulfide capacity shows a similar tendency in silicate-based and aluminate-based
melts. For the slag with fixed CaO content, in the case of Al2 O3 ↔ SiO2 substitution,
the sulfide capacity increases with decreasing structure complexity in silicate-based
melt. With the preceding arguments, it can be found that the sulfide capacity has
a negative correlation with structural complexity in silicate-based melts. However,
in aluminate-based melts with fixed CaO content, the Ca2+ can be consumed to
maintain charge neutrality within the [AlO4 ]5– tetrahedral with further additions of
Al2 O3 , which causes the FO to decrease and the structure complexity to increase.
The slight increase in sulfide capacity in aluminate-based melts might be due to the
length of Al–O and Si–O bonds. The length of the Al–O bond is greater than the
length of the Si–O bond, and the shorter bond length represents the stronger binding
electronegativity with oxygen atoms, which indicates that it is easier to generate O2–
in aluminate-based melts according to the polymerization reactions.
4.6 Electrical Conductivity 95

Fig. 4.19 Sulfide capacity, DOP, and FO concentration against the Al/Si atomic ratio. Reprinted
from Ref. [48]

4.6 Electrical Conductivity

Metallurgists do not pay special attention to the electrical conductivity of blast


furnace slag, but as an important high-temperature physical property of slag, elec-
trical conductivity also plays an important role in the study of slag structure. In
addition, it is possible to utilize the blast furnace slag to produce slag wool via elec-
tric arc furnace process, in this case, the electrical conductivity directly relates to the
power supply and heat distribution in the furnace.
96 4 High-Temperature Physicochemical Properties …

Nesterenko and Khomenko [51], and Winterhager et al. [52] and Zhu et al. [53]
reported that the electrical conductivity of CaO–SiO2 –Al2 O3 –MgO slag system
increases with increasing temperature and basicity, and decreases with the increase of
Al2 O3 content. Adachi et al. [54] demonstrated the effect of Al2 O3 /SiO2 ratio in the
SiO2 –5 wt% CaO–Al2 O3 –MgO and found that substituting SiO2 with Al2 O3 causes
an increase in electrical conductivity. While in the work of Segers et al. [55, 56], the
substitution of SiO2 with Al2 O3 leads to a decrease of the electrical conductivity in
the SiO2 –A12 O3 –CaO–MnO(–MgO) systems. Kawahara et al. [57] suggested that
MgO and CaO have the similar effect on the electrical conductivity of the slag.
Figure 4.20 show the effect of Al2 O3 content and substitution SiO2 with Al2 O3 on
the electrical conductivity of the studied slags. The electrical conductivity increases
with increasing temperature. The electrical conductivity decreases firstly and then
increases at fixed basicity, which obtains the minimum at approximately 24 wt%
Al2 O3 . A similar trend is found in the work of Sarkar et al. [58] and Hoster and
Pötschke [59]. While in the case of SiO2 ↔ Al2 O3 substitution, increasing the (wt%
Al2 O3 )/(wt% SiO2 ) ratio from 0.47 to 0.79 causes a slight increase in electrical
conductivity, and an opposite effect was found when the (wt% Al2 O3 )/(wt% SiO2 )
ratio is more than 0.79.
The measured electrical conductivity in this slag system is less than 0.5 S cm−1 ,
which obviously belongs to ionic conductivity. The cations and anions with simple
structures are mainly involved in the conduction, while the composite anions do
not participate in the conduction due to their large size. The conductivity of ions
is directly proportional to the migration rate of participating ions, and the latter is
proportional to the potential gradient formed by the external electron source, namely:
   
κ ∝ v0+ + v0− = B0+ + B0− ∂ E/∂ x (4.6)

where v0+ , and v0− are the migration rate of cations and anions, respectively, and B0−
are the coefficient, that is the cations and anions migration rate (m/s) when the
potential gradient (∂ E/∂ x) is 1, which is called mobility.

Fig. 4.20 Effect of Al2 O3 and (wt% Al2 O3 )/(wt% SiO2 ) on the electrical conductivity [60].
Reprinted from Ref. [60]
4.7 Summary 97

The greater mobility of the ions, the more electron charge they transmit. The cation
with small electrostatic potential has a greater mobility and a larger migration rate.
The resistance to ions migration comes from the energy barrier of their movement and
slag viscosity. Therefore, for ion-conducting slag, there is the following relationship
between conductivity and viscosity:

κ n η = K ; n = E η /E k (4.7)

where E η is the viscous activation energy, J/mol; E k is the conductance activation


energy, J/mol; K is a constant. It can be seen that conductivity is negatively related
to viscosity.
The relationship between electrical conductivity, viscosity, and degree of poly-
merization (DOP) and (atom Al)/(atom Si) are shown in Fig. 4.21. From the figure,
it can be seen that the viscosity and conductivity have opposite trends. In the case of
fixed basicity, the increase of Al2 O3 content leads to the content of CaO in the slag
gradually decrease, as well as the content of Ca2+ with greater mobility. Due to the
charge compensation effect, parts of the network modifiers are consumed resulting
in an increase in the degree of polymerization of the slag and the viscosity of the
slag, as well as a decrease in the conductivity of the slag. When Al/Si atomic ratio
is more than 1, the slag changes from silicate to aluminate, and AlO4 5− tetrahedron
becomes the dominant in the slag. Since the longer bond length of Al–O compared
with Si–O bond, the shear force of the slag flow becomes smaller, resulting in the
decrease of viscosity, so the conductivity increases. When the SiO2 is substituted by
Al2 O3 , the degree of polymerization of slag first decreases and then increases, with
a turning point of 0.92. The change law of viscosity and degree of polymerization
is consistent but opposite to that of electrical conductivity. The relationship between
structure and viscosity has been discussed in detail in Sect. 4.2.3.

4.7 Summary

The effect of Al2 O3 content and substitution SiO2 with Al2 O3 on the high-
temperature physicochemical properties of aluminosilicate slag were investigated.
The relationship between slag structure and properties was discussed.
1. The viscosity increases with increasing the Al2 O3 content at the range of 16–
24 wt% and the (wt% Al2 O3 )/(wt% SiO2 ) up 0.92, but it decreases when Al2 O3
content is more than 24 wt% and the (wt% Al2 O3 )/(wt% SiO2 ) below 0.92. The
free-running temperature increases with increasing the Al2 O3 content and shows
a peak at a (wt% Al2 O3 )/(wt% SiO2 ) of 0.92. Temperature dependence of the
viscosity is checked, from which the apparent activation energy is calculated.
The change of the apparent activation energy is in accordance with the change
of viscosity.
98 4 High-Temperature Physicochemical Properties …

Fig. 4.21 Electrical conductivity, viscosity, and DOP against the Al/Si atomic ratio. Reprinted from
Ref. [60]

2. The density decreases with increasing the Al2 O3 content up to nearly 24 wt% at a
fixed basicity of 1.2, subsequently increases. Increasing the (wt% Al2 O3 )/(wt%
SiO2 ) from 0.47 to 0.92 caused an increase in the density at a fixed CaO content,
and the density decreases slightly when the (wt% Al2 O3 )/(wt% SiO2 ) is greater
than 0.92.
References 99

3. The surface tension increases with increasing the Al2 O3 content and (wt%
Al2 O3 )/(wt% SiO2 ). The change in the surface tension can be explained by
the decrease in the surface-active component (SiO2 ) and the increase in the
number of charge compensators at the melt surface.
4. Electrical conductivity and viscosity of the slag have a negative correlation.
5. The sulfide capacity decreases with increasing the Al2 O3 content at a fixed
basicity of 1.20. Meanwhile, increasing the (wt% Al2 O3 )/(wt% SiO2 ) from
0.47 to 0.79 with a fixed CaO content causes a significant increase in the sulfide
capacity of the slag and a slight increase when the (wt% Al2 O3 )/(wt% SiO2 ) is
more than 0.79.
6. According to the slag structure, it can be found that the critical point of viscosity,
density, and conductivity is near the composition of the slag translation from
silicate to aluminate. That is, Al/Si atomic ratio is about 1.0, indicating that the
effect of Al2 O3 content and (wt% Al2 O3 )/(wt% SiO2 ) has different effects on
the physicochemical properties of the silicate-based and aluminate-based slag.

References

1. Peng, Q., Yang, C., & Li, G. (2005). Study on the viscosity of high Al2 O3 blast furnace slag
of Xianggang. Journal of Wuhan University of Science and Technology. 28(1), 11-13.
2. Shen, F. (2005). Discussion on blast furnace smelting process of high Al2 O3 content slag.
Ansteel Technology, 6.
3. Kou, T., Mizoguchi, K., & Suginohara, Y. (1978). The effect of Al2 O3 on the viscosity of
silicate melts. Journal of the Japan Institute of Metals, 42(8), 775–781.
4. Tang, X., Zhang, Z., Zhang, M., & Wang, X. (2011). Viscosities behavior of CaO-SiO2 -MgO-
Al2 O3 slag with low mass ratio of CaO to SiO2 and wide range of Al2 O3 content. Journal of
Iron and Steel Research, International, 18(2), 1–17.
5. Yao, L., Ren, S., Wang, X., Liu, Q., Dong, L., Yang, J., & Liu, J. (2016). Effect of Al2 O3 , MgO,
and CaO/SiO2 on viscosity of high alumina blast furnace slag. Steel Research International,
87(2), 241–249.
6. Kim, H., Matsuura, H., Tsukihashi, F., Wang, W., Min, D. J., & Sohn, I. (2013). Effect of Al2 O3
and CaO/SiO2 on the viscosity of calcium-silicate-based; slags containing 10 mass Pct MgO.
Metallurgical & Materials Transactions B, 44(1), 5–12.
7. Lee, Y. S., Min, D. J., Jung, S. M., & Yi, S. H. (2004). Influence of basicity and FeO content
on viscosity of blast furnace type slags containing FeO. Isij International, 44(8), 1291–1297.
8. Park, J. H., Min, D. J., & Song, H. S. (2004). Amphoteric behavior of alumina in viscous flow
and structure of CaO-SiO2 (-MgO)-Al2 O3 slags. Metallurgical & Materials Transactions B.
35(2), 269-275.
9. Sun, C. Y., Liu, X. H., Li, J., Yin, X. T., Song, S., & Wang, Q. (2017). Influence of Al2 O3 and
MgO on the viscosity and stability of CaO-MgO-SiO2 -Al2 O3 slags with CaO/SiO2 =1.0. Isij
International, 57(6), 978-982.
10. Liao, J., Zhang, Y., Sridhar, S., Wang, X., & Zhang, Z. (2012). Effect of Al2 O3 /SiO2 ratio on
the viscosity and structure of slags. Isij International, 52(5), 753–758.
11. Machin, J. S., Yee, T. B., & Hanna, D. L. (1952). Viscosity studies of system CaO–MgO–
Al2 O3 –SiO2 : III, 35, 45, and 50% SiO2 . Journal of the American Ceramic Society, 35(12),
322–325.
12. Nakamoto, M., Tanaka, T., Lee, J., & Usui, T. (2004). Evaluation of viscosity of molten SiO2 -
CaO-MgO-Al2 O3 slags in blast furnace operation. ISIJ International, 44(12), 2115–2119.
100 4 High-Temperature Physicochemical Properties …

13. Lv, X., Yan, Z., Liang, D., Zhang, J., & Bai, C. (2016). Transition of blast furnace slag from
silicates-based to aluminates-based: Viscosity. Metallurgical and Materials Transactions B,
48(2), 1092–1099.
14. Ilyushechkin, A. Y., Hla, S. S., Roberts, D. G., & Kinaev, N. N. (2011). The effect of solids
and phase compositions on viscosity behaviour and T CV of slags from Australian bituminous
coals. Journal of Non-crystalline Solids, 357(3), 893–902.
15. Kim, Y., & Oh, M. S. (2010). Effect of cooling rate and alumina dissolution on the determination
of temperature of critical viscosity of molten slag. Fuel Processing Technology, 91(8), 853–858.
16. Merzbacher, C. I., Sherriff, B. L., Hartman, J. S., & White, W. B. (1990). A high-resolution
29Si and 27Al NMR study of alkaline earth aluminosilicate glasses. Journal of Non-crystalline
Solids, 124(2), 194–206.
17. Neuville, D. R., Henderson, G. S., Cormier, L., & Massiot, D. (2010). The structure of crystals,
glasses, and melts along the CaO-Al2 O3 join: Results from Raman, Al L- and K-edge X-ray
absorption, and 27Al NMR spectroscopy. American Mineralogist, 95(10), 1580–1589.
18. Neuville, D. R., Cormier, L., Ligny, D. D., Roux, J., Flank, A. M., & Lagarde, P. (2008).
Environments around Al, Si, and Ca in aluminate and aluminosilicate melts by X-ray absorption
spectroscopy at high temperature. American Mineralogist, 93(1), 228–234.
19. Cormier, L., Ghaleb, D., Neuville, D. R., Delaye, J. M., & Calas, G. (2006). Chemical depen-
dence of network topology of calcium aluminosilicate glasses: A computer simulation study.
Journal of Non-crystalline Solids, 332(1), 255–270.
20. Sokolov, L., Baidov, V., Kunin, L., & Dymov, V. (1971). Surface and volume characteristics of
melted slags of the calcium oxide-alumina-silica systems. Tsentr Nauchno Issled Inst Chern
Metall, 75, 53–61.
21. Rajavaram, R., Kim, H., Park, J., Lee, C.-H., Cho, W.-S., & Lee, J. (2019). Bridging between
the physical properties; structure and density of CaO–SiO2 –Al2 O3 melts at CaO/SiO2 =1.3 and
different mole% of Al2 O3 . Ceramics International. 45(15), 19409-19414.
22. Rajavaram, R., Kim, H., Lee, C. H., Cho, W. S., Lee, C. H., & Lee, J. (2017). Effect of Al2 O3
concentration on density and structure of (CaO-SiO2 )-xAl2 O3 slag. Metallurgical & Materials
Transactions B, 48(3), 1595–1601.
23. Winterhager, L. G. H., & Kamnel, R. (1966). Forschungsberichte des Landes Nordrhein-
Westfalen. Westdeutscher Verlag.
24. Barrett, L., & Thomas, A. (1959). Surface tension and density measurements on molten glasses
in the CaO-Al2 O3 -SiO2 system. Journal of the Society of Glass Technology, 43(211), 179–190.
25. Yan, Z., Lv, X., Pang, Z., Lv, X., & Bai, C. (2018). Transition of blast furnace slag from
silicate based to aluminate based: Density and surface tension. Metallurgical and Materials
Transactions B, 49(3), 1322-1330.
26. Mukai, K., & Ishikawa, T. (1981). Surface tension measurements on liquid slags in CaO-SiO2 ,
CaO-Al2 O3 and CaO-Al2 O3 -SiO2 systems by a pendant drop method. Journal of the Japan
Institute of Metals, 45(2), 147–154.
27. Tian, Y. L., Guo, S. G., & Sun, S. B. (2015). Effect of Al2 O3 on surface tension of the SiO2 -
Al2 O3 -RO-R2O glass system. Key Engineering Materials, 633, 322–325.
28. Parmelee, B. C. W., & Harman, C. G. (2010). The effect of alumina on the surface tension of
molten glass. Journal of the American Ceramic Society, 20(1–12), 224–230.
29. Fenzke, H.-W. (1969). Surface tension of homogeneous slags of the CaO-SiO2 -Al2 O3 system.
Neue Hutte, 14(9), 6.
30. Kim, A., Akberdin, A., & Kulikov, I. (1984). Surface-tension of melts in the system CaO-SiO2 -
15-percent Al2 O3 -MgO-CaF2 . Russian Metallurgy (5), 32–35.
31. Kingery, W. D. (2006). Surface tension of some liquid oxides and their temperature coefficients.
Journal of the American Ceramic Society, 42(1), 6–10.
32. Nzotta, M. M., Sichen, D., & Seetharaman, S. (1998). Sulphide capacities in some multi
component slag systems. ISIJ International, 38(11), 1170–1179.
33. Nzotta, M. M., Sichen, D., & Seetharaman, S. (1999). A study of the sulfide capacities of
iron-oxide containing slags. Metallurgical and Materials Transactions B, 30(5), 909–920.
References 101

34. Nzotta, M. M., Sichen, D., & Seetharaman, S. (1999). Sulphide capacities of FeO-SiO2 , CaO-
FeO, and FeO-MnO slags. ISIJ International, 39(7), 657–663.
35. Nzotta, M. M. (1997). Experimental determination of sulphide capacities in the Al2 O3 -MgO-
SiO2 , Al2 O3 -MnO-SiO2 and Al2 O3 -CaO-MgO slags in the temperature range 1773–1923 K.
Scandinavian Journal of Metallurgy, 26(4), 169–177.
36. Nzotta, M. M., Nilsson, R., Sichen, D., & Seetharaman, S. (1997). Sulphide capacities in
MgO-SiO2 and CaO-MgO-SiO2 slags. Ironmaking and Steelmaking, 24(4), 300–305.
37. Karsrud, K. (1984). Sulfide capacities of synthetic blast furnace slags at 1500 deg C.
Scandinavian Journal of Metallurgy, 13(3), 144–150.
38. Sommerville, I. D., & Bell, H. B. (1982). The behaviour of titania in metallurgical slags.
Canadian Metallurgical Quarterly, 21(2), 145–155.
39. Zhang, X. (2012). Study on metallurgical properties of ultra-high Al2 O3 slag from Shigang blast
furnace. Dissertation pressed in Chinese, North China University of Science and Technology,
Hebei.
40. Shankar, A., Gornerup, M., Lahiri, A. K., & Seetharaman, S. (2006). Sulfide capacity of high
alumina blast furnace slags. Metallurgical and Materials Transactions B: Process Metallurgy
and Materials Processing Science, 37(6), 941–947.
41. Shankar, A. (2006). Sulphur partition between hot metal and high alumina blast furnace slag.
Ironmaking & Steelmaking, 33(5), 413–418.
42. Taniguchi, Y., Wang, L., Sano, N., & Seetharaman, S. (2012). Sulfide capacities of CaO-Al2 O3 -
SiO2 slags in the temperature range 1673 K to 1773 K (1400 °C to 1500 °C). Metallurgical
and Materials Transactions B, 43(3), 477–484.
43. Turkdogan, E. T. (1980). Physical chemistry of high temperature technology. Published by
Academic Press Inc, Massachusetts, USA.
44. Kim, K., Huh, W., & Min, D. (2014). Effect of FeO and CaO on the sulfide capacity of the
ferronickel smelting slag. Metallurgical and Materials Transactions B, 45(3), 889-896.
45. Park, J. H., & Park, G.-H. (2012). Sulfide capacity of CaO–SiO2 –MnO–Al2 O3 –MgO slags at
1873 K. ISIJ International, 52(5), 764–769.
46. Tang, X., & Xu, C. (1995). Sulphur distribution between CaO-SiO2 -TiO2 -Al2 O3 -MgO slag
and carbon-saturated iron at 1773 K. ISIJ International, 35(4), 367–371.
47. Shi, C.-B., Yang, X.-M., Jiao, J.-S., Li, C., & Guo, H.-J. (2010). A sulphide capacity prediction
model of CaO-SiO2 -MgO-Al2 O3 ironmaking slags based on the ion and molecule coexistence
theory. ISIJ International, 50(10), 1362–1372.
48. Yan, Z., Lv, X., Pang, Z., He, W., Liang, D., & Bai, C. (2017). Transition of blast furnace
slag from silicate based to aluminate based: Sulfide capacity. Metallurgical & Materials
Transactions B, 48(5), 1607-2614.
49. Woo, D. H., & Lee, H. G. (2010). Phase equilibria of the MnO–SiO2 –Al2 O3 –MnS system.
Journal of the American Ceramic Society, 93(7), 2098–2106.
50. Fincham, C. J. B., & Richardson, F. D. (1954). The behaviour of sulphur in silicate and aluminate
melts. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences,
223(1152), 40–62.
51. Nesterenko, S., & Khomenko, V. (1985). Study of the effects of alkalis on the surface-tension
and the electrical-conductivity of slags of the CaO-MgO-SiO2 system containing 5-percent
Al2 O3 . Russian Metallurgy (2), 42–45.
52. Winterhager, H., Greiner, L., & Kammel, R. (1966). Investigations of the density and elec-
trical conductivity of melts in the system CaO-Al 2 O3 -SiO2 and CaO-MgO-Al 2 O3 -SiO2 .
Westdeutscher Verlag.
53. Zhu, J.-H., Hou, Y., Zheng, W.-W., Zhang, G.-H., & Chou, K.-C. (2019). Electrical
conductivities of high aluminum blast furnace slags. ISIJ International, 59(3), 427–431.
54. Adachi, A., Ogino, K., & Hara, S. (1969). Rate and mechanism of silica reduction. Transactions
of the Iron and Steel Institute of Japan, 9(2), 153–161.
55. Segers, L., Fontana, A., & Winand, R. (1983). Electrical conductivity of molten slags of the
system SiO2 –Al2 O3 –MnO–CaO–MgO. Canadian Metallurgical Quarterly, 22(4), 429–435.
102 4 High-Temperature Physicochemical Properties …

56. Segers, L., Fontana, A., & Winand, R. (1979). Electrical conductivity, viscosity and density of
molten slags of the system CaO-SiO2 -MnO. Transactions-Institution of Mining and Metallurgy.
Section C. Mineral Processing & Extractive Metallurgy, 88, C53–C57.
57. Kawahara, M., Morinaga, K.-J., & Yanagase, T. (1983). Behavior of MgO and NiO in molten
slags. Canadian Metallurgical Quarterly, 22(2), 143–147.
58. Sarkar, S. B. (1989). Electrical conductivity of molten high-alumina blast furnace slags. ISIJ
International, 29(4), 348–351.
59. Hoster, T., & Pötschke, J. (1983). Die elektrische Leitfähigkeit FeO-haltiger CaO-Al2 O3 -
SiO2 -Schlacken mit Basizitäten ≤ 1, 5 bei 1450 bis 1650 °C. Arch Eisenhuttenwes, 54(10),
389–394.
60. Pang, Z., Lv, X., Yan, Z., Liang, D., & Dang, J. (2019). Transition of blast furnace slag
from silicate based to aluminate based: Electrical conductivity. Metallurgical and Materials
Transactions B, 50(1), 385–394.
Chapter 5
Estimation of High Alumina Blast
Furnace Slag Properties

The physicochemical properties of slag at high temperatures play a significant role


in the optimization and improvement of metallurgical processes. While the high-
temperature experimental measurements are always with high cost, time-consuming,
and difficult. Therefore, the estimation of slag properties is particularly important.

5.1 Types of Estimation Models

Many models have been developed for slag properties estimation and can be classified
into five groups, i.e., Numerical fitting models, Thermodynamic models, Structural-
based models, Computer simulation, and Machine learning.

5.1.1 Numerical Fitting Models

Numerical fitting is the process to construct a mathematic function by best fitting the
experimental database. This is the easiest way to build a model. For a property, e.g.,
liquidus temperature or viscosity, after obtaining the experimental data, a numerical
analysis is then carried out to get the best fit in which the property is expressed as
follows,

Property = (ai × X i %) (5.1)
i

where ai are the constant fitting parameters, X i % is the content of composition either
wt% or mol%. There are some problems with this approach. The unrealistic constants
can be recorded for components in low concentrations, it is frequently better in this
case to add the component with a low concentration to a component that presents in

© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 103
X. Lv and Z. Yan, High Temperature Physicochemical Properties of High Alumina Blast
Furnace Slag, https://doi.org/10.1007/978-981-19-3288-5_5
104 5 Estimation of High Alumina Blast Furnace Slag Properties

greater amounts and exhibits similar behavior. This approach is successfully applied
to slag systems within a limited compositional.

5.1.2 Thermodynamic Models

The thermodynamic models are based on the database of phase equilibrium and ther-
modynamic data for various systems. Several commercial thermodynamic software
is available, such as FactSage [1], Thermo-Calc [2], and MTDATA [3]. In the case of
FactSage, the solution model is used to describe the Gibbs free energy for each phase.
Simultaneously, the experimental data of thermodynamics and phase equilibrium in
the evaluated literatures are used to optimize the model parameters. According to the
Gibbs energy minimum theory, the optimized model is used to predict the equilib-
rium phase diagrams and thermodynamic properties such as chemical activities, free
energies, enthalpies, and heat capacity for multi-systems. With further improvement
in recent years, these commercial models have been extended to the calculation of
physical properties such as liquidus temperature, precipitation temperature, densities,
viscosities, and surface tensions of slag systems.

5.1.3 Structural-Based Models

The majority of reported models take the structure into account since the properties
always have a close relationship with the slag structure. It is particularly important
to use appropriate parameters to represent the structure of slag and establish the
relationship between structure and properties. Earlier models tended to use basicity
to represent the structure. There are also models that classify oxides according to their
acidities, such as Urbain model [4] and Ribound model [5]. Recent models tended to
use other parameters to represent the structure, for instance, optical basicity, NBO/T
or Q, the concentration of bridging or non-bridging, and free-Oxygen (NO0 , NO− , and
NO2− ). It should be noted that many later models were developed based on the earlier
models. For instance in model due to Reddy et al. [6, 7] the values of NO0 calculated
using Yokokawa model [8]. In contrast, the effect of the structure is determined
through a thermodynamic function in the Seetharaman and Du’s model [9].

5.1.4 Computer Simulations

With the development of computer technology and the deep understanding of the
slag structure at a molecular level, computer simulation has also become a powerful
method to calculate and predict the structure and properties of slag. The popular simu-
lation method for metallurgical slag is molecular dynamics (MD), which is based on
5.2 Liquidus and Solidus Temperatures 105

the statistical mechanics and potential energy functions to simulate the bonding char-
acteristics and motion of the molecules. Researchers have made effective attempts
to calculate the viscosity and diffusion coefficients via molecular dynamics simula-
tions [10, 11]. Since most properties of slag would be explained by understanding the
behavior of electrons, which can hold atoms together by forming bonds. The elec-
tronic structure theory and ab initio electronic structure calculations also have been
used to predict the properties and structure of slag [12, 13]. The MD simulation is
inexpensive in computer time and can simulate large systems with atomic numbers,
but the quality of the results depends on the accuracy of the potential energy func-
tions. The ab initio simulation using the electron density to calculate the properties
gives more accurate results. However, it is expensive in computer time and is limited
to small systems.

5.1.5 Artificial Neural Network Model (ANN)

Artificial intelligence (AI) has flourished in recent decades, and machine learning
is regarded as part of artificial intelligence. Machine learning is a type of algorithm
that automatically analyzes and obtains the rules from data, and the rules are used to
predict the unknown data. The commonly used machine learning algorithm for metal-
lurgical slag properties prediction is artificial neural network (ANN). The concept
of the artificial neural network comes from the observation of biological neuron
systems. As shown in Fig. 5.1, when units are inserted into the “input layer”, they
are transmitted to the “middle layer”, and when the intensity exceeds a critical value,
they are transmitted onto the “output layer”. The neural network model was created
by adjusting the weights of each layer through training data (this process is called
learning), which has the following characteristics:
1. A set of weights that can be adjusted (numerical parameters adjusted by the
learning algorithm). These adjustable weights can be regarded as the strength
of connections between neurons;
2. A Nonlinear function relationship for estimating the input data.
The machine learning approaches are used to estimate the viscosity, surface
tension, and sulfide capacity of slags [14–18].

5.2 Liquidus and Solidus Temperatures

The available commercial thermodynamic software offers the function to calculate


the relationship between phase equilibrium and temperature of multicomponent slag
in a specific range of compositions, which can calculate reliability of liquidus and
solidus temperatures. However, the commercial thermodynamic software has access
requirements, and not all the systems have been optimized. For a wide range of
106 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.1 Schematic diagram illustrating the concept of biological neuron and artificial neural
network model

components, it is difficult to calculate the value of liquidus and solidus tempera-


ture. Therefore, it may be necessary to create a database of liquidus and solidus
temperatures after measurements and correlate these temperatures with their chem-
ical composition to obtain the “best fit” to estimate the liquidus and solidus temper-
atures for the specific system. Example of this approach is determination of liquidus
of mold fluxes from experimental data and chemical composition [19].

Tliq (K) = 1464 + 11.4 × SiO2 % − 11 × CaO% + 4.2 × Al2 O3 % + 5.7


× MgO% − 10.1 × Na2 O% − 15.8 × K2 O% + 1.9
× F% + 8.3 × Fe2 O3 % + 11.6 × MnO% (5.2)

It is should be noted that the numerical fitting model only suitable for the family
slag systems in a specific range of compositions. For the T liq calculation of mold
fluxes, some typical applications allow values to be calculated to be better than ±30 K
[19].
5.3 Viscosity 107

5.3 Viscosity

5.3.1 Overview of Viscosity Models

As one of the most important properties of slag, viscosity models draw more attention
from considerable scholars. Early researchers proposed the dependence equations
of viscosity and temperature since the obvious effect of temperature on viscosity.
Later research work introduced the slag composition, structural parameters, and
thermodynamic parameters into the relationship based on the viscosity-temperature
relationship and developed many viscosity models. Since the slags in metallurgical
processes are not all uniform, the appearance of solid phase makes a large deviation of
the experimental viscosity compared with the viscosity from the calculation models
applied in fully melted slags. Therefore, some models suitable for the viscosity
calculation of solid–liquid coexisting systems have also emerged.

5.3.1.1 Viscosity-Temperature Dependence Models

Since the nineteenth century, extensive studies have been conducted on the tempera-
ture dependence of viscosity. What is best known at present are the Arrhenius model
[20], the Vogel-Fulcher-Tammann model [21–23], the Adam-Gibbs model [24], the
Eyring model [25], the Weymann-Frenkel model [26], and the Avramov-Milchev
model [27, 28].
Arrhenius Model
This model is a well-known approach and is widely used to describe the temperature
dependence of viscosity for silicate slags. At the end of nineteenth century, Arrhe-
nius investigated the viscosity of aqueous solutions and reported the temperature-
dependent behavior:

ln η = A + B/T (5.3)

where η is the viscosity; A and B are composition-dependent constants; T is the


absolute temperature in K. The relationship between ln η and 1/T is not always
linear, so that B is a parameter related to composition and temperature [29]. The
move melt is described as the relative movement of structural units at the molecular
level, and the premise of structural unit displacement is to have certain viscous
activation energy to overcome the energy barrier Eη (Eq. 1.2).
Vogel-Fulcher-Tammann (VFT) Model
Compared to the Arrhenius model, the VFT model introduces another constant
parameter C, which is expressed as:

ln η = A + B/(T − C) (5.4)
108 5 Estimation of High Alumina Blast Furnace Slag Properties

where A, B, and C are composition-dependent constants, respectively. Urbain [30]


found that this model is suitable for describing the relationship between the viscosity
and temperature of silicate melts, supercooled liquids, and glass. However, Mauro
et al. [31] pointed out that the model has a large error when extrapolated to low
temperature.
Adam-Gibbs Model
Based on the configurational entropy theory, Adam and Gibbs proposed a model
to explain the relaxational properties of glass-forming liquids. They stated that the
viscous flow of a liquid system occurs by the cooperative rearrangements of the
varying species (atoms, ions, and molecules) in the liquid. In the 1980s, Richet et al.
[24, 32] investigated the silicate melts viscosity, and proposed that the structural
relaxation time could be employed to indicate the shear strain relaxation time. Thus,
a general Adam-Gibbs viscosity model was achieved and is given by:

ln η = A + B/(T · Sc (T )) (5.5)

where A and B are composition-dependent constants. S C (T ) is the configurational


entropy at T, which can be calculated by:

T
CP
SC (T ) = SC (Tref ) + dt (5.6)
T
Tref

where T ref is a reference temperature and C p is the configurational heat capacity.


The Adam-Gibbs model provides a basis semi-theoretical model for calculating
the viscosity of molten silicates, and well describes the temperature dependence
of viscosity in the range from 1 to 1012 Pa s [33].
Eyring Model
Eyring proposed a model based on the absolute reaction rate theory by coupling
it with the viscosity in the molecular scale [25]. A chemical reaction occurs after
overcoming the energy barriers, and the Maxwell–Boltzmann equation was employed
to describe the distribution of energy. The chemical reaction also involves the change
in the equilibrium distance between atoms or molecules. The viscous flow can also
be described as relative movements of two layers of molecules in the liquids. From
this point, the Eyring model is given by:
   
hN G ∗ h Nρ G ∗
η= · exp = · exp (5.7)
Vm RT M RT

where h is the Planck’s constant; N is the Avogadro’s number; V m is the molar


volume; G∗ is the Gibbs energy of activation of viscous flow; R is the gas constant;
5.3 Viscosity 109

T is the absolute temperature; ρ is the density and M is the molecular weight. The
Eyring model is similar in form to the Arrhenius model but different in theory.
Weymann-Frenkel Model
The hole theory assumes that the liquid has a quasi-crystalline structure, and the
properties of a liquid are determined by the motion of the holes [34, 35]. Once the
molecule in liquid has enough energy to overcome the barrier, it can move from one
equilibrium position to the next equilibrium empty position. On the basis of hole
theory, Weymann [26] deduced the mechanism of viscosity from mechanical and
statistical concepts:
  21  
RT (2mkT )1/2 EW
η= · · exp (5.8)
EW v 2/3 PV KT

where R is the gas constant; T is the absolute temperature; E W is the energy; m and
v are the mass and volume of the structural unit, respectively; k is the Boltzmann
constant; PV is the “hole” probability connected with the structural model of the
liquid. The Weymann and Frenkel can be simplified as following:

B
ln η = ln A + ln T + (5.9)
T
An extra temperature term is introduced compared with the Arrhenius model and
shows the best fit for the temperature dependence of most glasses and slags [36].
Avramov-Milchev Model
Avramov and Milchev [27, 28] link the viscosity to the thermally activated jumping
mechanism of slag, assuming that the viscosity is inversely proportional to the average
jump frequency of the building unit of the system. The specific jump occurs only
after the corresponding activation energy is overcome. The probability of the energy
barrier is used to calculate the average jump frequency:


v = Vi · Pi (E i ) (5.10)
i=0

where ν is the average jump frequency; V i is the molecule jump frequency; Pi (E i ) is
the probability distribution function. Employing the dispersion of activation energy
and the activation energy value at the maximum probability, the solution of Eq. (5.10)
is obtained, and then the dispersion and entropy are linked to derive the Avramov-
Milchev viscosity model:
 α
θ
η = η∞ · exp (5.11)
T
110 5 Estimation of High Alumina Blast Furnace Slag Properties

where: η∞ is the constant; T is the absolute temperature; θ is a parameter in Kelvin;


α is dimensionless parameter.
The above models give the relationship between viscosity and temperature and
discuss the viscosity mechanism on the molecular scale. These models are mainly
applied in simple liquids and cannot be directly applied to complex metallurgical slag
systems such as aluminosilicate melts. Nevertheless, these models provide a reason-
able framework for the further development of viscosity models of multicomponent
systems.

5.3.1.2 Compositions Fitting Models

The composition fitting models are numerical fitting models based on the relationship
between the slag compositions and the viscosity-temperature dependence. Table 5.1
shows some classic viscosity-composition fitting models. The simplest way is to
directly fit the slag compositions to obtain the parameters, such as Lakatos model
[37] and Kim model [38]. This type of model is simple and has good prediction
results within the range of its fitting all data. The Urbain model [39, 40] is based on
the Weymann-Frenkel equation and divides the slag composition into three types:
network formers, network modifiers, and amphoteric oxides. Riboud et al. combine
CaF2 and alkali metal oxides based on the Urbain model and improve the prediction
accuracy of the model. Kondratiev and Jak [41] modified the Urbain model and
applied it to the Al2 O3 –CaO–FeO–SiO2 system. The model uses a set of model
parameters to accurately predict the entire composition (pure, binary, ternary, and
quaternary systems) in a certain temperature range. CaO–SiO2 –Al2 O3 –MgO is an
important basic slag system for metallurgical processes. Thanks to a large amount
of viscosity experimental data, Gan and Lai [42] established a slag viscosity model
with a wider range of content based on the Vogel-Fulcher-Tammann equation. These
models were fitted based on the viscosity-temperature dependence through existing
experimental data and only applied for specific slag systems. Many fitting parameters
were introduced to obtain higher model prediction accuracy with lacking physical
meaning.

5.3.1.3 Structural-Based Models

The viscosity of slag is closely related to its structure, and it is particularly impor-
tant to use appropriate parameters to characterize the slag structure and establish
the relationship between structure and viscosity. The commonly used characteriza-
tion parameters of structure are described in Chap. 3. The structural-based viscosity
models are summarized in Table 5.2.
5.3 Viscosity 111

Table 5.1 Summary of classic viscosity-composition fitting models


Sources η–T Description
A = 1.5183X Al2 O3 − 1.6030X CaO − 5.4936X MgO
Lakatos Vogel-Fulcher-Tammann
[37] + 1.4788X Na2 O − 0.8350X K2 O − 2.4550

B = 2253.4X Al2 O3 − 3919.3X CaO + 6285.3X MgO


− 6039.7X Na2 O − 1439.6X K2 O + 5736.4

C = 294.4X Al2 O3 + 544.3X CaO − 384.0X MgO


− 25.07X Na2 O − 321.0X K2 O + 471.3

ln A = −2.307 − 0.046X SiO2


Kim et al. Arrhenius − 0.07X CaO − 0.041X MgO − 0.185X Al2 O3
[38] + 0.035X CaF2 − 0.095X B2 O3

B = 6807 + 70.7X SiO2 + 32.58X CaO


+ 312.7X Al2 O3 − 34.8X Na2 O − 176.1X CaF2
− 167.4X Li2 O + 59.7X B2 O3
Urbain Weymann-Frenkel Glass formers (G): SiO2 , P2 O5 , etc
et al. [39, Network modifiers (N): CaO, MgO, TiO2 , etc
40] Amphoterics (A): Al2 O3 , Fe2 O3 , etc
η = AT exp(B/T )
ln A = −(0.293B + 11.571)
∗ + B (X ∗ )2 + B (X ∗ )3
B = B0 + B1 X G 2 G 3 G
Bi = ai + bi α + ci α 2 (i = 0, 1, 2, 3)
α = G/(N + A) in molar ratio
where, parameter A obtained from B, ai , bi , and ci is the
fitting constant, respectively
Riboud Weymann-Frenkel “SiO2 ”: SiO2 , P2 O5 , B2 O3, etc
et al. [5] “CaO”: CaO, MgO, FeO, MnO, etc
“Al2 O3 ”: Al2 O3 , TiO2, etc
“CaF2 ”: CaF2
“Na2 O”: Na2 O, K2 O etc
ln A = −19.81 + 1.73X CaO + 3.58X CaF2
+ 7.02X Na2 O − 35.76X Al2 O3

B = 31.140 − 23.896X CaO − 46.356X CaF2


− 39.159X Na2 O − 68.833X Al2 O3
where, A and B are the composition-dependent
constants; X i is the mole fraction of component “i”
(continued)
112 5 Estimation of High Alumina Blast Furnace Slag Properties

Table 5.1 (continued)


Sources η–T Description
Kondratiev Weymann-Frenkel ln A = −(m · B + n)

and Jak m = mi · Xi
[41]

3  2 
3 
CaO, j XC
B= bi0 · X SiO
i
+ bi ·
2
X CaO + X FeO
i=0 i=0 j=1

FeO, j X CaO
+bi · · α j · X SiO
i
X CaO + X FeO 2

where mi and X i are the model parameter for a pure


oxide and the mole fraction of the corresponding oxide,
respectively. bi is the parameter
 
Gan and Vogel-Fulcher-Tammann B = bi X i ; C = ci X i
Lai [42] n 
 
2
log ηmean − A + T −C B
j=1
where, bi and ci are the composition-dependent
constants; X i is the mole fraction of component “i”

5.3.1.4 Solid–Liquid Multiphase Viscosity Models

Many metallurgical processes are carried out in a multiphase slag system. For
example, the formed high melting point components can suspend in the slag, or
precipitated particles exist when the temperature is lower than the liquidus. The exis-
tence of the solid phase leads to a large deviation in the viscosity model developed
based on the fully molten state. The deviation depends on the solid phase volume
fraction, particle shape and size, solid phase distribution, etc. In order to describe the
viscous behavior of the solid-containing slag system, predecessors have made lots
of efforts, such as the Einstein [56] model and the Roscoe [57] model.
At the beginning of the twentieth century, Einstein discovered and deduced the
relationship between the viscosity of a solid–liquid mixture and the volume fraction
of solid particles, as shown in Eq.

ηe = ηL · (1 + 2.5θ ) (5.12)

where, ηe is the effective viscosity; ηL is the pure solvent viscosity; θ is the volume
fraction of solid particles in the slag.
This model is only suitable for a mixture with a very low fraction of solid particles
(less than 5%) and the solid phase is spherical particles and does not consider particle–
particle interaction. Whitmore pointed out that the factor may vary according to the
particle size distribution. Based on the Einstein model, Roscoe studied the viscosities
of a suspension containing spheres with very diverse sizes and developed a model
with a nonlinear relationship between the viscosity and volume fraction of solid
particles.
Table 5.2 Summary of structural-based models
Parameters Sources η–T Description
Basicity (B) Iida et al. [43, 44] Arrhenius A = 1.745 − 1.962 × 103 T + 7.00 × 10−7 T 2
E = 11.11 − 3.65 × 10−3 T
5.3 Viscosity


η0 = η0i · X i
 ∗
αi Wi B +αFe W
2 O3 Fe2 O3
Bi∗ =  ∗
αi Wi A +αAl O WAl O +αTiO ∗ W
2 3 2 3 2 TiO2

where Bi is modified basicity index; η0i is hypothetical viscosity of pure
component i; αi is specific coefficient; and Wi is mass percentage of
component i; The subscripts A and B represent acidic oxide and basic oxide,
respectively; αi∗ is the modified specific coefficient indicating the interaction
of the amphoteric oxide with other components
V
Optical basicity ( ) Mills and Sridhar [45] Arrhenius ln A = −232.22corr − 357.3corr − 144.2
ln(B/100) = −1.77 + 2.88/corr
Corrected optical basicity (corr ) is a measure of depolymerization
Concentrations of bridging oxygen Reddy et al. [6, 46] Weymann-Frenkel η = (2/3)Nh Rh (2π mkT )1/2 exp(E/RT )
(O0 ), non-bridging oxygen (O− ) and  
1/(n−1)
free oxygen (O2− ) E = E ∗ / A 1 − NO0 + 1
where N h is the number of holes per unit volume, Rh is the average radius of
the holes; π = 3.1416; m is mass weight of the basic building units (kg), k is
the Boltzmann constant (J/K), R is the gas constant [J/(mol K)]
NO0 is fraction of bridging oxygen
 3  2 
Zhang and Jahanshahi [47, 48] Weymann-Frenkel E η = a + b NO0 + c NO0 + d NO2−
ln A = a + bE η
where: a, b, c, and d are fitting parameters; NO0 and NO2− are mole
fractions of bridging and free oxygen, respectively
(continued)
113
Table 5.2 (continued)
114

Parameters Sources η–T Description


Tanaka et al. [49] Arrhenius EV = E
1+(αm )1/2

m−1
 
αm = Mi O · NO− + NO2−
i=1
m−1
  m α Mi + α M j
+ · NO2−
2
i=1 j=i+1

1 
Jak et al. [50, 51] Eyring 2RT · 2π m SU kT 2 · exp E α
η = E 2/3 RT
vap vSU

m SU = m O0 X O0 + m O− X O− + m O2− X O2−
vSU = vO0 X O0 + vO− X O− + vO2− X O2−
E α = E α,O0 X O0 + E α,O− X O− + E α,O2− X O2− + E αch/c

E vap = exp E v,O0 X O0 + E v,O− X O−

+E v,O2− X O2−
where: E vap is the energy of vaporization; E a is the activation energy; mSU
and vSU are the weight and volume of a structure unit of viscous flow,
respectively
where: X i , vi , and m i are the mole fractions, weights, and volumes of the
structural units i; E α,i is the partial molar activation energies, and E αch/c
represents the charge compensation effect; E v,i is the dimensionless partial
vaporization energy coefficients
(continued)
5 Estimation of High Alumina Blast Furnace Slag Properties
Table 5.2 (continued)
Parameters Sources η–T Description
Zhang et al. [52] Arrhenius ln A = k(E − 572,516) − 17.47
 
k= x i ki / xi
5.3 Viscosity

i, j =SiO2 i, j =SiO2
572,516×2
E=    i  i
n O +αAl n O + αi ·n O + αAl,i ·n O + αSi ·n i + αAl ·n i
Si Al i Al,i OSi OAl
where ki is model parameters; n O is number of different types of oxygen;
and βi describes deforming ability of bond around corresponding unit
NBO/T Senior and Srinivasachar [53] Weymann-Frenkel A = a0 + a1 B + a2 · NBO/T
Qn
B = b0 + b1 α + b2 α 2 + b3 N + b4 N α
+ b5 N α 2 + b6 N 2 + b7 N 2 α + b8 N 2 α 2
+ b9 N 3 + b10 N 3 α + b11 N 3 α 2
 
Avramov et al. [28] Avramov-Milchev η = η∞ / 4 n En
n=0 Q · exp − R·T
where: η∞ is the pre-exponential constant; Qn is the concentration of Si with
n “strong” oxygen bonds
(continued)
115
Table 5.2 (continued)
116

Parameters Sources η–T Description



FactSage [54, 55] Arrhenius Y Q 4,n = P n

A = AMeO X × (MeO X ) + A∗SiO × (SiO2 )


2
E
+ ASiO × (SiO2 )P 40 + AMe−Si × (MeO X ) × (SiO2 )
2

R
+ AMe−Si × (SiO2 ) P 4 − P 40


E = E MeO X × (MeO X ) + E SiO × (SiO2 )
2
E
+ E SiO × (SiO2 )P 40
2
+ E Me−Si × (MeO X ) × (SiO2 )

R
+ E Me−Si × (SiO2 ) P 4 − P 40

where: p is the probability to form bridging oxygen


 
G* KTH model Eyring η= Vh N · exp G ∗ = h Nρ · exp G ∗
RT M RT
m
Du and Seetharaman [9] 
G ∗ = X i j · G i∗j + G ideal +  E G ∗
where: Gij ∗ is the Gibbs energy of activation of pure component in liquid
state; the term Gideal represents the change in Gibbs energy resulting from
the ideal mixing of components; E G∗ represents the change in Gibbs
energy resulting from the mutual interactions between different species
5 Estimation of High Alumina Blast Furnace Slag Properties
5.3 Viscosity 117

ηe = ηL · (1 − θ )−2.5 (5.13)

This model can apply to the system with different solid particles size, and
theoretically, there is no limit to the volume fraction of solid particles.
These two similar models are summarized as the Einstein-Roscoe model,

ηe = ηL · (1 − R · θ )−n (5.14)

where: R and n are the empirical parameters obtained from experimental data.
Kondratiev and Jak stated that this model provided good performance for the CaO–
SiO2 –Al2 O3 –“FeO” system with partly precipitated solid phase by assuming the
empirical factors R = 2.04 and n = −1.29 [58]. Wright reported that those two
parameters are related to the particle size, but they did not give a specific relationship
[59]. The Einstein-Roscoe model is widely used due to its simple form, but R and
n vary with the studied systems. After the R and n parameters are combined with
functions such as solid phase particle size and shape, the application range of this
model can be greatly expanded [60–62].

5.3.1.5 Computer Simulation and Artificial Neural Network Model


(ANN)

In recent decades, with the development of computer technology, the calculation


of the physicochemical properties of the melt based on computer simulation calcu-
lation methods (density functional ab initio algorithm, first principles or molec-
ular dynamics, etc.) has also attracted the attention of researchers. Part researchers
have made effective attempts to calculate the viscosity via molecular dynamics
simulations. Statistical analysis of the trajectory of particles in the slag through
molecular dynamics simulation can obtain the relationship between the mean square
displacement (MSD) and time.

 1  
MSD = r (t)2 = ri(t) − ri(0)  (5.15)
N
where, ri(t) presents the location of atom i at the time of t and the bracket expresses
an ensemble average taken over many origin times. Combining statistical mechanics
and statistical thermodynamics, the MSD curve of the particles is used to calculate
the self-diffusion coefficient, D.
 
1 d r (t)2
D = lim (5.16)
t→∞ 6 dt
Then the viscosity η can be calculated by combining the self-diffusion coefficient
with the Einstein-Stokes equation.
118 5 Estimation of High Alumina Blast Furnace Slag Properties

KBT
η= (5.17)

where, K B is the Boltzmann constant, 1.38 × 10−23 J/K, λ is the step of ion diffusion.
Wu et al. [10] applied this method to CaO–SiO2 and CaO–Al2 O3 systems, and the
calculated results show a good agreement with the experimental data. However, as
far as the current molecular dynamics simulation is concerned, the system that can
be calculated is relatively simple, and the potential function is the key to the entire
simulation during the simulation.
The artificial neural network (ANN) has been increasingly used in various indus-
trial fields in recent years because it can handle complex and multi-scale nonlinear
relationships. Several researchers have attempted to predict slag viscosity via ANN.
Hou et al. [63] developed an ANN model for CaO–SiO2 –Al2 O3 –FeO–MgO system
but only 75 data points were used for learning. Duchesne et al. [64], Hanao et al.
[15], and Cheng et al. [65] established ANN models to predict the viscosity of coal
ash, mold flux, and blast furnace slag over a broad composition and temperature
range, respectively. Although these ANN-based models show quite good predictive
performance, their application scopes are limited by the composition of the simple
database and model architecture. Hence, Chen et al. [66, 67] proposed a structure
informed artificial neural network model to predict the viscosity for the composition
systems of CaO, SiO2 , MgO, Al2 O3 , Fe2 O3 , FeO, R2 O (Na2 O + K2 O), TiO2 , MnO,
P2 O5 , Cr2 O3 , V2 O5 and CaF2 in the temperature range of 530–2482 °C.

5.3.2 Structure-Based Viscosity Modeling

5.3.2.1 Hole Theory of Melts

The hole theory of melts considers that in the process of high-temperature thermal
movement, some of the ions constituting clusters are displaced relative to each other,
then holes between them are produced. motion fluid is considered to consist of
layers lying parallel to the direction of flow. The momentum transfer exists between
these layers when holes jump between neighboring layers, causing viscous drag, as
shown in Fig. 5.2. Based on Frenkel’s kinetic theory, the structure unit oscillates in
the energetic cells near average and results in the movement into an adjacent hole
formed in the liquid due to the fluctuations. No matter whether the fluid is in motion,
particles in the fluid continuously execute the random motion. The random walk
of particles carries out the momentum transfer between layers, which is visible to
the observer as the viscosity of the fluid. Thus, Bockris and Reddy [68] deduced an
equation to describe the viscosity based on the above-mentioned principles.
Assuming three adjacent parallel layers of A, B, and C in the fluid moving with
velocities V +( ∂∂ VZ )L, V , and V −( ∂∂ VZ )L, respectively. where Z is the direction normal
to the planes and L is the mean free path of the particles populating the layers, i.e.,
the mean distance traveled by the particles without undergoing collisions. In the
5.3 Viscosity 119

Fig. 5.2 Viscous forces


arise from transfer of
momentum between moving
layers in fluid [69]. Reprinted
from Ref. [69] CC BY 4.0

direction of motion of the moving layers, the momentums of the particles traveling
in the A, B, and C layers is m(V + ( ∂∂ VZ )L), mV , and m(V − ( ∂∂ VZ )L)(V − ( ∂∂ VZ )L),
respectively. When a particle jumps from the A layer to B layer, the momentum gained
by the B layer is −m(V + ( ∂∂ VZ )L), i.e., the momentum transported per particle in the
downward direction is −m(V + ( ∂∂ VZ )L). If one considers that there are N particles
per unit is, B layer is S cm2 and W is the mean velocity of particles in the direction
normal to the layers, the momentum transferred per second in the downward direction
owing to A–B jumps is −[N W Sm(V + ( ∂∂ VZ )L)]. This rate of change of momentum
is equal to a force (Newton’s second law of motion). Thus, the viscous force F is
given by

F = −(2N W m)S∂ V /∂ Z (5.18)

According to Newton’s law of viscosity, the viscous force is proportional to the


area of the layers and to the velocity gradient, and the proportionality constant is the
viscosity, i.e.,

F = −ηS∂ V /∂ Z (5.19)

Combining Eqs. 5.18 and 5.19, it is clear that

η = 2N mW (5.20)

For the fused slags, the hole model has been proposed by Fürth [35]. It is consid-
ered that the sizes and spatial location of the empty regions in the fused slags are
random. These randomly located and variable-sized vacancies are called holes. Holes
move at finite velocities and have an inertial resistance to motion, i.e., have masses
and momenta. According to the hole theory, holes play the role in pure ionic liquids.
120 5 Estimation of High Alumina Blast Furnace Slag Properties

Thus, it is considered that the random walk of hole between adjacent layers results
in momentum transfer and, therefore, viscous drag in moving fused slag exists. The
expression for the viscosity of an ionic liquid, on the basis of this model, is

η = 2Nh m h W (5.21)

where Nh and m h are the number of hole per unit volume and the mass of holes,
respectively. this equation shows a very sample expression for viscosity. For esti-
mating the viscosity of liquid systems, it is necessary to express the above terms for
slags. The velocity component Wh is given by the ratio of the mean time between
collisions.

Wh = L/t (5.22)

The viscosity can be written as follows



η = 2Nh t[m h Wh )2 (5.23)

The theorem of the equipartition of energy can be applied to the one-dimensional


motion referred to Wh .

1/2m h (Wh )2 = 1/2kT (5.24)

Substituting into, then obtains:

η = 2Nh kT t (5.25)

In a fused slag, k is Boltzmann constant. t is the mean lifetime of a hole, i.e., the
average time between creation and destruction of a hole through thermal fluctuations.
It is given by:
   21  
Rh 2π m E
t= exp (5.26)
3 kT RT

where Rh is the average radius of the hole, m is the mass of particle and E is the
activation energy of particle to overcome the barrier and move to an adjacent hole.
Combining Eqs. 5.25 and 5.26, the following equation can be obtained as:
   
2 1 E
η= Nh Rh (2π mkT ) exp
2 (5.27)
3 RT

This is a kinetic model to calculate viscosity. The activation energy E consists of


two parts, one is the energy required to form the holes, and the other is the energy
needed for the structure units moving to an adjacent hole. For aluminosilicate melts,
5.3 Viscosity 121

holes are easy to form but the structural units are hard to move since the T–O bond
must be broken before moving. Therefore, in the aluminosilicate slag system with a
network structure, the activation energy is determined by the energy required to break
the network bond. The number of holes (Nh ) is related to the non-bridging oxygen
(NO0 ) concentration. According to the atom pair model proposed by Yokokawa [8],
the bridging oxygen content in the slag can be calculated. Fürth [35] pointed out that
the radius of the holes (Rh ) in the melt is approximately the same as the radius of the
structural unit, that is, the minimum hole radius should be equal to the radius of the
tetrahedral structural unit.

5.3.2.2 Construction of Model

The viscosity model is constructed according to the relationship between the


slag viscosity, structure information, and temperature obtained by Eq. 5.27. The
main steps include calculation of different types of oxygen concentration in slag,
the establishment of an effective viscosity database, and optimization of model
parameters.
Three types of oxygen are suggested in a silicate melt: bridging oxygen (O0 ,
connects two tetrahedrons), non-bridging oxygen (O− , connects a tetrahedron to a
modifier), and free oxygen (O2− , links two network modifiers). In aluminosilicate
melts, charge compensation is required for [AlO4 ]5− tetrahedra due to the difference
in the valence compared with [SiO4 ]4− . On the contrary, in the Al2 O3 –CaO–MgO–
FeO–SiO2 system, there are three metal cations (Ca2+ , Mg2+ , and Fe2+ ) that could act
as charge compensators. Figure 5.3 shows a schematic diagram of the different types
of oxygen ions considered in the model construction. Zhang et al. [52] suggested that
there is a strict order for which cations carry out the charge compensation of the Al3+
ions with the highest priority being given to the cations with the lowest field strength.
Thus, priority for charge compensation is in the order: Ca2+ > Fe2+ > Mg2+ . When
the alkali metal atoms are not enough for charge balance, the local charge balance
of the [AlO4 ]5− tetrahedra will no longer be attained. In this case, the oxygen ions
bonded with non-compensated Al3+ ions are considered as a new type of oxygen
ion. Therefore, for CaO–SiO2 –Al2 O3 –MgO–FeO system, four types of oxygen ions
are present: bridging oxygen Oio , non-bridging oxygen O−j , free oxygen O2− , and
excess bridging oxygen O∗ . For the system, C mol% CaO–S mol% SiO2 –A mol%
Al2 O3 –M mol% MgO–F mol% FeO, the details of the calculation of the different
types of oxygen are given as follows.
(a) For Al2 O3 –SiO2 –MO (X MO < X Al2 O3 , MO = CaO + MgO + FeO):

3( A − (C + M + F)) 4(C + M + F)
NO∗ = ; NO0Al = ;
C + M + F + 2S + 3A C + M + F + 2S + 3A
2S
NO0Si = ; (5.28)
C + M + F + 2S + 3A
122 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.3 Schematic


illustrations of the structure
for different types of oxygen
atoms in the melts. Reprinted
from Ref. [69] CC BY 4.0

Specific examples: Pure SiO2: NO0Si = 1; Pure Al2 O3 : NO∗ = 1; For


Al2 O3 –SiO2 : NO0Si = 2S+3A
2S
; NO∗ = 2S+3A
3A
;
(b) For Al2 O3 –SiO2 –MO (X MO > X Al2 O3 , MO = CaO + MgO + FeO):
The charge and mass balance can be expressed as Eqs. 5.29 and 5.30, respectively
as:
1
nO0 = 4S + 8A − nO− (5.29)
2
1
nO2− = 2C + 2M + 2F − 2 A − nO− (5.30)
2
The total number of oxygens can be calculated as:

Total O = nO0 + nO− + nO2− = 2S + 3A + C + M + F (5.31)

where nO0 is the number of bridging oxygen ions. nO− is the number of non-bridging
oxygen ions. nO2− is the number of free oxygen ions. nO− further can be calculated
by Eq. 5.32. Based on those relations, the mole fractions of different types of oxygen
are the ratio with the total number of oxygens.
 
G 0
4S + 8A − nO− 2C + 2M + 2F − 2 A − nO− = exp (nO− )2 (5.32)
RT
5.3 Viscosity 123

G0 is the Gibbs energy of depolymerization reactions and can be expressed as



G 0 = NSi4+ NCa2+ G 01 + NMg2+ G 02 + NFe2+ G 03

+ NAl3+ NCa2+ G 04 + NMg2+ G 05 + NFe2+ G 06 (5.33)

where Gi 0 (i = 1, 2, 3, and 4) are the standard free energy change for the depoly-
merization reactions, which are shown in Table 5.3. N j are the mole fractions of
cations which can be calculated as
2S 3A
NSi4+ = ; NAl3+ = ; (5.34)
2S + 3A 2S + 3A
C−A
if X CaO > X Al2 O3 : NCa2+ = ;
C+M+F−A
M F
NMg2+ = ; N 2+ =
C + M + F − A Fe C+M+F−A
if X CaO < X Al2 O3 , X CaO + X FeO > X Al2 O3 : NCa2+ = 0;
C+F−A M
NMg2+ = ; NMg2+ = ;
C+M+F−A C+M+F−A
if X CaO + X FeO < X Al2 O3 , X CaO + X FeO + X MgO > X Al2 O3 : NCa2+ = 0;
C+M+F−A
NFe2+ = 0; NMg2+ = ;
C+M+F−A
if X CaO + X FeO + X MgO < X Al2 O3 : NCa2+ = 0; NFe2+ = 0; NMg2+ = 0. (5.35)

nO0
NO0 = ; NO0Si = NSi4+ NO0 ; NO0Al3+ = NAl3+ NO0 (5.36)
Total O
nO−
NO− = ; NO−
Ca2+
= NCa2+ NO− ; NO−
Mg2+
= NMg2+ NO− (5.37)
Total O
To establish an effective viscosity database, the experimental data sets of the
CaO–SiO2 –Al2 O3 –MgO–FeO system and its subsystems for using in the model
optimization also have been critically reviewed. Table 5.13 presents the sources of

Table 5.3 Standard free


Reaction G0 , J/mol
energy change for the
depolymerization reactions 2CaO + SiO2 = Ca2 SiO4 G1 0 = −118,712 − 11.3T
2MgO + SiO2 = Mg2 SiO4 G2 0 = −67,131 − 4.3T
2FeO + SiO2 = Fe2 SiO4 G3 0 = −44,182 + 18.07T
3CaO + Al2 O3 = Ca3 Al2 O6 G4 0 = −17,000 − 32.0T
3MgO + Al2 O3 = Mg3 Al2 O6 G5 0 = −21,582 − 7.8T
FeO + Al2 O3 = FeAl2 O4 G6 0 = −15,714 − 0.434T
124 5 Estimation of High Alumina Blast Furnace Slag Properties

the collected data (see Appendix). Three measurement methods are involved in the
table, and the rotating method as a reliable viscosity measurement technique is widely
used. The materials of crucible/spindle and the atmosphere would change the slag
composition. SiO2 will be reduced by carbon at high temperatures (T > 1520 °C,
calculated by FactSage), so the experimental data measured by graphite crucible and
spindle was carefully analyzed, and the viscosity data at high temperatures (more
than 1520 °C) was not used for model optimization. The Mo crucibles and spindle
may oxide to generate MoO3 when the equipment is not completely sealed, which
will affect the viscosity results. When measuring the viscosity of slag containing
FeO, it is necessary to control the oxygen potential or use slag-iron equilibrium to
ensure the FeO content in the slag. Therefore, the results with post-chemical analysis
were preferred. In addition, part of the reported viscosity data was conducted on
solid–liquid mixture which was not taken into account for optimization (liquidus
were predicted by FactSage). It should be noted that the viscosities measured by
different researchers at closed compositions were carefully compared to cross-check
the reliability of the data. According to the analysis of the above evaluation, more
than 4000 sets of effective viscosity data from nearly 60 literatures were used for the
optimization of model parameters.
1
During the model calculation, the term (2π mkT ) 2 in Eq. 5.27 can be rearranged
as,
 1
1 W 2
(2π mkT ) = 2π
2 kT (5.38)
RT

where W is the molecular weight of the basic building unit (kg/mol), and can be
expressed as:

W = NSi4+ WSiO4 + NAl3+ WAlO4 (5.39)

where WSiO4 = 0.092 kg/mol, and WAlO4 = 0.091 kg/mol.


The radius of the tetrahedral units (Rh ) can be expressed as:

Rh = NSi4+ RSiO4 + NAl3+ RAlO4 (5.40)

where RSiO4 = 2.99 Å, and RAlO4 = 3.14 Å.


The number of holes per unit volume (N h ) was expressed in terms of NO0 and
NO*. Thus

Nh = NO0 + NO∗ × Av (5.41)

where Av is Avogadro’s number (6.023 × 1023 g mol).


The value of E can be calculated based on the available experimental viscosities
by using Eq. 5.27 and a polynomial expression was deduced that is given by:
5.3 Viscosity 125

Table 5.4 Values of constant A and B for calculating the activation energy
a A B A B
−66.830 0.190
bi = Si4+ n=1 0.067 0.032 cj = Mg2+ n=1 651.948 −0.323
n=2 128.926 −0.227 n=2 −1425.673 0.702
n=3 458.532 0.030 n=3 1191.962 −0.574
bi = Al3+ n=1 74.873 −0.048 cj = Fe2+ n=1 −502.904 0.280
n=2 849.203 −0.397 n=2 3236.800 −1.983
n=3 −766.938 0.355 n=3 −3636.345 2.298
cj = Ca2+ n=1 768.001 −0.384 dk n=1 676.738 −0.309
n=2 −1306.970 0.629 n=2 −614.261 0.252
n=3 806.477 −0.370 n=3 126.667 −0.046


3

E =a+ (bi,n (NOio )n + c j,n (NO−j )n + dk,n NO∗ )n (5.42)
n=1

where i = Si4+ , or Al3+ ; j = Ca2+ , Mg2+ , or Fe2+ , a, b, c, and d are constants,


which were further a linear function of temperature and can be expressed by the
following general equation:

ξ = A + BT (5.43)

where ξ represents a, b, c, d. The values of A and B are optimized parameters by


available experimental viscosities data, as given in Table 5.4.

5.3.2.3 Performance of Viscosity Model

To evaluate the performance of the structural-based model (named as Reddy-Yan


model) on viscosity prediction, the average of relative errors  was used, as expressed
by:

N  
1  ηi,est − ηi,meas 
= × 100% (5.44)
N n=1 ηi,meas

where ηi,meas and ηi,est are the measured and estimated viscosity values, respectively.
N is the number of viscosity data.
126 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.4 Comparison between experimental data and calculated data for silica and alumina.
Reprinted from Ref. [69] CC BY 4.0

(a) Melton Silica and Alumina


Due to the critical role of silica in metallurgy, ceramics, and glass industries, its
physicochemical properties in the molten states have attracted much attention of
researchers, especially viscosity. As shown in Fig. 5.4, the calculated results from
the Reddy-Yan model presented here reproduce well the experimental viscosity data
for silica melts, except for the data of Bockris et al. [70]. Bruckner [71] figured
that the significant deviation is due to the pollution of silica by carbon vapors from
heating elements. The arrow in the figure shows the liquid phase region (FactSage
calculation), the solid lines are the model calculation results, and the points are
the experimental results from literature. The viscosity decreases with increasing
temperature, and the relationship between viscosity and temperature fellows to the
Arrhenius equation, and the calculated viscous activation energy is 512.5 kJ/mol.
It can be seen that the viscosity of pure silica is very high, even at 2500 °C with a
viscosity higher than 300 Pa s. when the temperature drops to the melting temperature,
the viscosity can reach 5.0 × 107 Pa s. The reason for the high viscosity is the complex
network structure of silica (Q4). Compared with silica, the number of viscosities data
of molten alumina is lesser due to its high melting temperature. The predicted value
of the melt viscosity of pure alumina by the Reddy-Yan model is in good agreement
with the measurement data. The viscous activation energy calculated according to the
Arrhenius equation is 111.8 kJ/mol. It should be noted that the viscosity of molten
5.3 Viscosity 127

Fig. 5.5 The viscosity for Al2O3-SiO2 system at various temperature (a), and comparison between
experimental data and calculated data (b). Reprinted from Ref. [69] CC BY 4.0

silica is much higher than that of alumina. Holes are formed and particles jump into
these holes during flowing. In molten silica and alumina, holes are easily formed
since the rate of hole formation is controlled by the vibrations of the atoms relative
to each other. Therefore, the bond-rupture step is the rate-determining process. The
strength of the Si–O–Si bond is stronger than that of Al–O–Al bond which holds the
tetrahedron together resulting in the higher viscosity.
(b) Binary Systems
For Al2 O3 –SiO2 system, there are no metal cations for charge compensation of
the [AlO4 ]-tetrahedra, and only bridging and excess bridging oxygens are present.
Figure 5.5a shows the effect of mole fraction of SiO2 on viscosity at 2126 K (1853 °C)
and 2226 K (1953 °C) for Al2 O3 –SiO2 system. The Reddy-Yan model has a good
prediction performance with using excess bridge oxygen. Comparing the calculated
viscosity data with the experimental data (Fig. 5.5b), although the average value of
the relative error is 21.90%, considering the high temperature and high viscosity
involved in the high SiO2 content, this relative error is acceptable.
Unlike Al2 O3 , CaO, MgO, and FeO can reduce the viscosity of silicate slag
systems as network modifiers. Figure 5.6 shows the viscosity comparison between
the calculated results and the experimental measurements of the MO–SiO2 (MO
= CaO, MgO). The initial addition of MO into SiO2 melt decreases the viscosity
sharply. With the further increase of MO content, the network structure of melt does
not change too much, resulting in the viscosity tending to be more gentle. It is should
be noted that the viscosity values have also been calculated for the compositions
outside the liquid region for hypothetical. The average relative error is 13.57% for
the CaO–SiO2 system and 15.02% for the MgO–SiO2 system.
As the basic slag system for copper smelting and oxygen converting, FeO–SiO2
system attracted the attention of many metallurgists, especially for the viscosity.
Since FeO can be oxidized, the oxygen potential during the experiment needs to be
strictly controlled. Most of the measurements are carried out by the spinning cylinder
method under the condition of slag-Fe (iron crucible and iron probe) equilibrium. I
128 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.6 The viscosity for CaO–SiO2 and MgO–SiO2 systems at various temperatures (a, b).
Reprinted from Ref. [69] CC BY 4.0, and comparison between experimental data and calculated
data (c, d)

the work of Kaiura et al. [72] and Kucharski et al. [73], Mo crucible and spindle were
used, and the oxygen partial pressures were controlled by CO/CO2 mixture (PO2 =
6 × 10–11 atm). In these cases, the Fe3+ /Fe2+ ratio is maintained in a low level and
the viscosities are expected to be closer to the FeO–SiO2 system. Figure 5.7a shows
the effect of SiO2 content on the calculated and experimental viscosity of FeO–SiO2
system. The calculated viscosity increases monotonically with increasing silica. It
should be noted that the measured viscosity exhibits a peak near the fayalite (Fe2 SiO4 )
composition observed by Shiraishi et al. [74], Kaiura et al. [72], and Urbain et al.
[40] The reasons for the local maximum viscosity near fayalite composition remain
controversial. Shiraishi et al. [74] reported that this phenomenon was attributed to the
formation of fayalite clusters, whereas Waseda and Shiraishi [75] explained that with
a combination of silicate anion polymerization effect and cation effect. However, in
the work of Chen et al. [76] and Kucharski et al. [73], the peak at the fayalite compo-
sition was not observed. Zhang and Jahanshahi [48] pointed that the presence of Fe3+
affects the structure of the melts at fayalite composition. According to the comprise
of measured data with calculated data, the estimated viscosity reproduces well at low
viscosity range but is not very good at high viscosity range, as shown in Fig. 5.7b.
5.3 Viscosity 129

Fig. 5.7 The viscosity for FeO–SiO2 system at different temperatures (a) and comparison between
experimental data and calculated data (b) [77]. Reprinted from Ref. [77]

The lack of valuable data at high viscosity range results in the underestimated the
ability of FeO on decreasing viscosity at high silica content. Despite that, the average
of relative errors for FeO–SiO2 system is in the vicinity of 18.72%.

(c) Ternary Systems

In the CaO–MgO–SiO2 system, the replacement of CaO with MgO increases the
viscosity, reported by Licko [78], which means that the non-bridge oxygen linked
with Ca2+ cations has different effect on the viscosity compared with that of Mg2+ .
In the Reddy-Yan model, the NO−j (j = Ca2+ , Mg2+ , Fe2+ ) is introduced to describe
the effect of different cations on the modification of the network structure. Figure 5.8
shows the replacement of CaO for MgO at a fixed SiO2 content at 50 mol% causes a
decrease in viscosity, and the calculated viscosities agree well with the experimental
data.

Fig. 5.8 CaO–MgO–SiO2 system: melts viscosity at 50 mol% SiO2 (a), Reprinted from Ref. [69]
CC BY 4.0; Comparison between experimental data and calculated data (b)
130 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.9 a Comparison between the calculated iso-viscosity cures and the experimental data for
the CaO–FeO–SiO2 system at 1673 K. b Comparison between experimental data and calculated
data in the CaO–FeO–SiO2 system. Reprinted from Ref. [77]

Five sets of data were found on the viscosity of CaO–FeO–SiO2 system. Chen
and Zhao [79], Ji et al. [80], Kucharski et al. [73], and Williams and Briggs [81] used
iron crucibles and spindles in inert gas atmosphere to maintain the slag composition.
While Kucharski et al. [73] did the measurement under a low oxygen partial pressure
(PO2 = 6 × 10–11 atm) with Mo sets. Due to the low oxygen partial pressure, it is
reasonable to treat the system as a ternary CaO–FeO–SiO2 . Johannsen and Wiese
[82] used falling sphere method to measure the viscosity in nitrogen gas atmosphere.
The predicted iso-viscosity cures at 1673 K (1400 °C) are shown in Fig. 5.9a, from
which the present model reproduces good of the viscosities for the system in compar-
ison with some data. The viscosity generally increases with increasing silica content.
The calculated viscosity from present model is compared in Fig. 5.9b with the exper-
imental data for CaO–FeO–SiO2 system, and the average of relative error is found
to be 16.93%.
The viscosity of MgO–FeO–SiO2 system has been reported by a few researchers,
and all of them used the rotational viscometer. Chen et al. [83], Kucharski et al. and
Toropov and Bryantsev [84] used Mo crucibles and spindles for viscosity experi-
ments, while Williams and Briggs [81] used iron sets in argon gas atmosphere. Ji
et al. [85] used iron crucibles and spindles for low-temperature measurements and
Mo sets for high temperature in Ar gas atmosphere. The predicted iso-viscosity
cures at two temperatures 1573 K and 1623 K (1400 °C and 1500 °C) are presented
in Fig. 5.10a, compared with the experimental data. A good agreement between
the calculation and experimental data, and the viscosity increases in principle with
increasing SiO2 concentrations. Figure 5.10b shows the comparison between exper-
imental data and calculated data. It should be noted that all the measurements by
Toropov and Bryantsev [84] at 1573 K and 1623 K were conducted below liquids
temperature (predicted by FactSage), and the oxygen partial pressure was not clari-
fied. As a result, these viscosity data are significantly higher than are expected ones
from fully liquid slags. Thus, these data were fully rejected in the viscosity model
optimization. Kucharski et al. [73] studied the effects of CaO and MgO on viscosity
5.3 Viscosity 131

Fig. 5.10 a Comparison between the calculated iso-viscosity cures and the experimental data for
the MgO–FeO–SiO2 system at 1673 K. b Comparison between experimental data and calculated
data in the MgO–FeO–SiO2 system. Reprinted from Ref. [77]

of MeO–FeO–SiO2 systems and suggested that the ability to decrease viscosity in


silicate melts is in the order of FeO > CaO > MgO. Figure 5.11 shows the calcu-
lated and measured viscosity results as a function of X MeO / X SiO2 (here MeO is CaO,
MgO, and FeO) with a fixed mass ratio of Fe/Si = 2.9 at 1573 K (1300 °C). The
same order is obtained from the present model indicating that the model can provide
a reasonable fit to the data and shows that the replacement of FeO by CaO or MgO
results in increasing the viscosity.

Fig. 5.11 Viscosity results as a function of various oxide additions to FeO–SiO2 system. Reprinted
from Ref. [77]
132 5 Estimation of High Alumina Blast Furnace Slag Properties

When Al2 O3 exists in the slag system, due to the charge compensation, the influ-
ence of network modifiers CaO, MgO, and FeO on the viscosity is different from
that of the slag system without Al2 O3 . Figure 5.12a, b well describe the viscosity
change of MO–Al2 O3 –67 mol% SiO2 slag system. With increasing of Al2 O3 content,
the viscosity of the system first increases and then decreases, and the maximum
viscosity appears in the fully charged compensation composition (MO/Al2 O3 = 1).
Kozakevitch [86] suggested that viscosity maxima occur on the Al2 O3 side at the
fully charge-compensated composition. The comparison of the calculated viscosities
with the experimental viscosities data of CaO–Al2 O3 –SiO2 and MgO–Al2 O3 –SiO2
systems are shown in Fig. 5.12c, and d. It should be noted that the measurements
by Matýsek [87] were performed in a graphite crucible, which may partially reduce
SiO2 and lead to experimental data unreliable. The measurements by Cukierman and
Uhlmann [88] were performed under non-equilibrium conditions, and Taniguchi [89]
used the falling ball method with high uncertainty. Therefore, these data were not
used for model optimization.

(d) Quaternary and Quinary Systems

Fig. 5.12 The viscosity for CaO–Al2 O3 –SiO2 and MgO–Al2 O3 –SiO2 systems at various temper-
ature (a, b) Reprinted from Ref. [69] CC BY 4.0, and comparison between experimental data and
calculated data (c, d)
5.3 Viscosity 133

CaO–MgO–SiO2 –Al2 O3 is an important slag system in the processes of iron making


and steel refining. Figure 5.13a shows the effect of basicity and Al2 O3 content on the
viscosity of CaO–MgO–SiO2 –Al2 O3 slag with fixed MgO content. The viscosity
decreases with increasing basicity and increases with increasing Al2 O3 content.
Figure 5.13b shows the effect of (wt% Al2 O3 )/(wt% SiO2 ) on the viscosity of
CaO–MgO–SiO2 –Al2 O3 slag with fixed CaO content. The slag viscosity decreases
with increasing of MgO content. With the increase of (wt% Al2 O3 )/(wt% SiO2 ), the
viscosity decreased at 5 wt% of MgO but almost no significant change at 10 wt% of
MgO. The strength of Al–O bond is weaker than that of Si–O bond. The replacing
SiO2 with Al2 O3 contributes to decrease the viscosity. With further addition of
Al2 O3 content, the basic oxides will be consumed to maintain the charge balance
in the [AlO4 ]-tetrahedron, which decreases the network modifiers and increases the
viscosity. Figure 5.13c compares the calculated data with the experimental data. The
Reddy-Yan model can describe the viscosity of the CaO–MgO–Al2 O3 –SiO2 system
well with temperature and composition, with an average relative error of 18.35%.

Fig. 5.13 Predicted viscosities of CaO–SiO2 –Al2 O3 –MgO melt as function of CaO/SiO2 wt%
ratio (a) Reprinted from Ref. [69] CC BY 4.0, (wt% Al2 O3 )/(wt% SiO2 ) wt% ratio (b) Reprinted
from Ref. [69] CC BY 4.0, and comparison between experimental data and calculated data (c)
134 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.14 Comparison between experimental data and calculated data of CaO–MgO–FeO–SiO2
system (a), and CaO–Al2 O3 –FeO–SiO2 system (b). Reprinted from Ref. [77]

Williams and Briggs [81] and Ji et al. [85] reported the experimental data on the
CaO–MgO–FeO–SiO2 systems. Figure 5.14a shows the comparison between exper-
imental data and calculated with about 12% relative error. Most of measured data by
Williams and Briggs [81] are lower than the calculated ones, but the measurements
from Ji et al. [85] were reproduced well by Reddy-Yan model. For the CaO–Al2 O3 –
FeO–SiO2 systems. Azuma et al. [90] and Johannsen and Wiese [82] used falling
sphere method to measure the viscosity in inter gas atmosphere. Hurst et al. [91, 92]
and Higgins [93] used Mo sets in N2 atmosphere by rotational technique. Gimmel-
farb [94] carried out his study by torsional method in Ar gas atmosphere, but most of
the measurements were conducted below liquids temperature. Figure 5.14b shows a
good agreement for the predicted viscosities with experimental results. although the
average of relative error is found to be 25%, it is close to the experimental scatter.
The CaO–MgO–Al2 O3 –FeO–SiO2 system is a fundamental slag system for steel
making process. Kim et al. [95], Lee et al. [96], Bills [97], and Higgins [93] reported
the viscosity measurements with iron saturation. Figure 5.15a illustrates the predicted
iso-viscosity cures and the experiment data of CaO–MgO–Al2 O3 –FeO–SiO2 system
with 10 wt% Al2 O3 and 5 wt% MgO at 1723 K. The estimated viscosity decreased
when CaO was initial added in the slag system with a fixed SiO2 content, which is
commonly due to that the Ca2+ has the highest priority for charge compensation and
FeO is the strongest modifier for this slag. When the CaO replaces the FeO and acts
as a modifier, the viscosity increases. The measures from Kim et al. [95] and Lee
et al. [96] at this temperature are obtained below the liquidus, which results that the
experimental data are much higher than expected values from fully liquid slags. The
measurements by Bills [97] are agree well with the present model. Figure 5.15b shows
the comparison between experimental data and calculated data, and a good agree-
ment for the predicted viscosities with experimental results. Figure 5.16 summarizes
the results of analyzing the average of relative errors for the Al2 O3 –CaO–MgO–
FeO–SiO2 and its subsystems, which the data for Al2 O3 –CaO–MgO–SiO2 and its
5.3 Viscosity 135

Fig. 5.15 a Comparison between the calculated iso-viscosity cures and the experimental data for
the CaO–MgO–Al2 O3 –FeO–SiO2 system at 1723 K. b Comparison between experimental data and
calculated data in the CaO–MgO–Al2 O3 –FeO–SiO2 system. Reprinted from Ref. [77]

Fig. 5.16 Average of


relative errors between the
calculated viscosities and
experimental data for
Al2 O3 –CaO–MgO–FeO–
SiO2 system. Reprinted from
Ref. [77]

subsystems from previous work. It can be seen that the accepted experimental viscosi-
ties are agreed well with the predicted data by present model; the average of relative
errors for this model is in the vicinity of 20%.

5.3.3 Iso-viscosity of High Alumina Blast Furnace Slag

The viscosity of CaO–SiO2 –9 wt% MgO–Al2 O3 –1 wt% TiO2 system was measured
(Chap. 4), and the valid viscosity data were used to verify the accuracy of the appli-
cable models. For the Reddy-Yan model calculation, 1 wt% of TiO2 was ignored.
Figure 5.17a shows the comparison between the models calculated value and the
136 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.17 Comparison between the measured viscosity and estimated viscosity for CaO–SiO2 –
Al2 O3 –9 wt% MgO–1 wt% TiO2 (a), iso-viscosity of the CaO–SiO2 –9 wt% MgO–Al2 O3 –1 wt%
TiO2 slag system at 1500 °C (b)

experimental data. Reddy-Yan model has the best prediction performance with about
10.8% of relative errors. Riboud model and Zhang model also give an acceptable
prediction result but shifted to one side. Iida model and Urbain model have large
errors, and the predicted results from FactSage are less than the measurement value,
which is not acceptable. The iso-viscosity distribution diagram can be obtained in
the diagram by using Reddy-Yan model calculation, as shown in Fig. 5.17b. The gray
area in the figure presents the liquid area calculated from the thermodynamic soft-
ware FactSage with a phase diagram module; each curve describes the distribution of
the viscosity related to the slag composition. The viscosity decreases with increasing
basicity at fixed alumina content. In high CaO content region, the viscosity shows
slightly increase when the SiO2 is substituted by Al2 O3 with fixed CaO content due
to the charge compensation effect. When Al2 O3 content is more than 24 wt% with
low SiO2 , CaO content ranges from 35 to 45 wt%, and the slag transform from
silicates-based to aluminates-based can still get a region with good fluidity. In other
words, there is no problem in BF operation with high alumina in the aluminates-based
region.

5.4 Density

5.4.1 Overview of Density Models

As one of the most important physical properties of molten slags, density plays a
crucial role in processes of pyrometallurgy, including metal droplets sedimentation,
metal–slag separation, slag foaming, and so on. However, it is difficult to get accurate
data from high-temperature measurements due to their complexity and variability.
5.5 Surface Tension 137

Therefore, many slag density models have been proposed in the publication over
decades.
The density of molten slag can be calculated via the following equation:

X i Mi
ρ= (5.45)
Vm

where ρ is the density of the slags, X i is the mole fraction of component i, Mi


is the molar mass of component i, and Vm is the molar volume of slag at the
measurement temperature. A brief summary of these models is presented in Table
5.5. It should be noted that all the partial molar volume of component i (V i ) is
temperature-dependence, known as temperature coefficient (dV i /dT ).

5.4.2 Density of High Alumina Blast Furnace Slag

In most density models, the calculation of slag molar volume is the core step. The
partial molar volume of CaO–SiO2 –MgO–Al2 O3 –TiO2 components at 1500 °C in
few models, and the recommended temperature coefficient values are listed in Table
5.6. According to the recommended values, the density of CaO–SiO2 –MgO–Al2 O3 –
TiO2 was calculated via the following models and compared with the experimental
data, as shown in Fig. 5.18. All the model calculations show that the slag density
increases with Al2 O3 content and Al2 O3 /SiO2 ratio, which is inconsistent with the
experimental results. These models still have some deficiencies in describing the
density of aluminosilicates, but the calculated results are still available for reference.
The density of these types of slag is less than 3.0 g/cm3 according to the calculation
and experimental results.

5.5 Surface Tension

5.5.1 Overview of Surface Tension Models

Surface tension is not a bulk property and tends not to be very structure-dependent.
Then the surface-active components such as B2 O3 , CaF2 , and SiO2 have a lower
surface tension and tend to occupy the surface layer. While, in silicate slags, since
the surface tension of SiO2 is much lower than that of CaO, the surface tension tends
to decrease with increasing degree of polymerization (DOP). The surface tension of
the slag is dependent upon the chemical activity of the surface-active component,
which determines the surface-active concentration in the surface layer. Consequently,
Table 5.5 Summary of slag density calculation models
138

Type Source Description Systems Reported


deviation
(%)

Additive models Bottinga and Vm = X i Vi CaO–SiO2 –MgO–Al2 O3 –FeO–Li2 O–Na2 O–K2 O 2–3
Weill [98] V i is the partial molar volume of component i
The partial volumes of the components at
1400 °C and the coefficients of thermal
expansion (1200–1700 °C) are recommended

Bottinga Vm = X i Vi + X A VA CaO–SiO2 –MgO–Al2 O3 –FeO–Li2 O–Na2 O–K2 O 2
et al. [99] V A is the composition-dependent apparent (30–90% SiO2 )
partial molar volume of Al2 O3 . Considering the
A13+ structure in four-fold and six-fold
coordination

Mills [100, Vm = X i Vi BOS slags 2
101] The partial molar volumes are different BF slags
compared with Bottinga’s work and have a good Mold flux
performance for metallurgical slags
Lange and V (T ) = X i (T )Vi (T ) + X Na2 O X TiO2 VNa2 O−TiO2 CaO–SiO2 –MgO–Al2 O3 –FeO–Fe2 O3 –TiO2 –Li2 O–Na2 O–K2 O 1
Carmichael
[102]

Excess molar Nakajima Vm = X i Vi + V EX Multicomponent silicate melt 2
volume model [103] −1
N N

V EX = i j X i X j
i=1 j=i+i

V EX is the excess molar volume due to mixture


(continued)
5 Estimation of High Alumina Blast Furnace Slag Properties
Table 5.5 (continued)
Type Source Description Systems Reported
deviation
(%)

Courtial and Vm = X i Vi + XS CaO–SiO2 –MgO–Al2 O3 /
Dingwell XS is the excess volume term corresponding to
5.5 Surface Tension

[104, 105] the interaction between SiO2 and CaO



Linard et al. Vm = X i Vi (T ) + V EX Silicate molten glass 4
[106] −1
N N

V EX = X i X j ViEX
j
i=1 j=i+i

ViEX
j is the excess volume contribution of
interactions between each pair of components

Gan et al. Vm = X i Vi + V EX CaO–SiO2 –MgO–Al2 O3 –FeO–Fe2 O3 1.62
[107, 108] EX

V = ai j X i X j + bT
i = j
Where ai j represents the interaction factor
between components i and j. b is the fitting
parameter
mix
Thermodynamic Hayashi and  VX V = K H RT Slags 2
i i i
models Seetharaman
[109] where H mix is the relative integral enthalpy of
mixing (determined with thermodynamic
software) and K is constant
(continued)
139
Table 5.5 (continued)
140

Type Source Description Systems Reported


deviation
(%)
M M
Persson V = λ HRT CaO–SiO2 –MgO–Al2 O3 –FeO–MnO 2–5
i X i Vi
[110, 111]
where VMis the relative integral molar volume

and i X i Vi represents the average molar
volume, λ is a constant, H M the relative integral
molar enthalpies for mixing. Values of H M and
λ were obtained from ThermoSlag software
Shu [112] Vm = X MO VMO + X SiO2 VSiO2 + Vmix MO–SiO2 systems 3

Vmix = λ X MO VMO + X SiO2 VSiO2 G mix /RT
where λ is a constant for a binary system and
values of G mix calculated by Toop-Samis
model

Structure models Shu and Vm = X i Vi + Vmix Slags (SiO2 , CaO, Al2 O3 , MgO, FeO, MnO, Na2 O, K2 O) 2
Chou [113] 
Vmix = λ X i Vi G mix /RT
Charge compensation of Al3+ was taken into
consideration. Optical basicity is used to
calculate the mixing molar volume Vmix
Zhang and ρ = ρ0 exp(E/RT ), CaO–SiO2 –MgO–Al2 O3 0.6
Chou [114] where E was expressed as a function of optical
basicity
(continued)
5 Estimation of High Alumina Blast Furnace Slag Properties
Table 5.5 (continued)
Type Source Description Systems Reported
deviation
(%)

Vm = n Q 4 aSiO2 + bSiO2 T
5.5 Surface Tension

3 
 X Si−i (ani + bni T )
+ n Qn i 
Thibodeau i X Si−i Li2 O–Na2 O–K2 O–MgO–CaO–MnO–PbO–Al2 O3 –SiO2 2
n=0
et al. [115, 
116] + nt X i− j ai− j + bi− j T
i j

where n Qn and n t are the amount of Qn species


and total number of oxygens in 1 mol of liquid,
and i and j are cations of network modifiers.
X Si−i and X i− j are bond fractions of
non-bridge oxygen (Si–i pairs) and free oxygen
(i–j pairs) in 1 mol of liquid. The molar volume

parameters of pure SiO2 aSiO2 + bSiO2 T
Q n (ani + bni T ) are obtained by the
optimization of the molar volume of binary
MO–SiO2 melt
Geometric Chou et al. The mass triangle model can use limited Ternary systems /
model [117, 118] experimental information to calculate the CaO–Al2 O3 –SiO2 and BaO–FeO–Fe2 O3
density within the homogenous phase region

ρ1673 kg/m3 = 2490 + 12(%FeO + %Fe2 O3
Numerical Mills and BOS slag 5
fitting model Keene [100] +%MnO + %NiO)
141
142 5 Estimation of High Alumina Blast Furnace Slag Properties

Table 5.6 The partial molar volume and temperature coefficient of components of few models
Component (i) CaO SiO2 MgO Al2 O3 TiO2
Mills model Vi 20.7 19.55 + 7.96X SiO2 16.1 28.31 + 32X Al2 O3 24
[119] − 2
31.45X Al 2 O3

dV i /dT 2.07 0.1Vi 1.61 0.1Vi 2.4


Linard model Vi 16.9 27.07 12.71 32.72 23.88
[106] dV i /dT 3.74 0.47 1.9 −9.4 7.24
Lange model Vi 16.84 26.88 11.85 37.52 23.98
[102] dV i /dT 4.22 −0.33 2.45 0.74 7.24
Dingwell model Vi 16.27 29.91 13.48 34.56 31.01
[120] dV i /dT 4.57 0.77 3.93 −3.94 −1.51

Fig. 5.18 Comparison between the measured density and calculated density

the models for calculating the slag surface tension are normally based on the thermo-
dynamics of the slag system. The proposed methods to calculate the surface tension
are summarized in Table 5.7.

5.5.2 Iso-surface Tension of High Alumina Blast Furnace


Slag

Tanaka and Nakamoto [125–129] developed a thermodynamic model for predicting


the surface tension of alloys and slags based on the Butler’s equation. Hanao [130],
and Liu [131] extended this model to multicomponent systems. For the slag systems
containing surface-active components (such as B2 O3 , CaF2, and Na2 O), the predicted
values of this model are also in good agreement with the experimental values. The
Tanaka model was used for prediction the surface tension of high alumina blast
furnace slag and the details are described as follows,
Table 5.7 Summary of models to calculate the slags surface tension
Source Description Systems Reported
deviation

γ (mN/m) = 271.2 + 148X Li2 O − 222X Li2 O


+ 196X K2 O + 334X MgO + 148X CaO
Kucuk et al. CaO–SiO2 –MgO–MnO–Al2 O3 –FeO–SrO–BaO–Li2 O–Na2 O–K2 O 1–3%
5.5 Surface Tension

[121] + 128X BaO + 332X SrO + 268X FeO


+ 292X MnO + 347X Al2 O3
Using multiple linear regression analysis to estimate
the surface tension parameters of the oxide
components in the silicate glass melt

Zhang et al. γ = X i γi + γ EX Binaries of molten slags and salts /
[122] EX
 2
 CaO–Al2 O3 –Fe2 O3
γ = X i X i A0 + A1 (X i − X i ) + A2 (X i − X i )
γ EX in the model has been designated as the excess
surface tension. A0 , A1 , and A2 are the interaction
coefficients independent of composition and vary
with temperature
Nakamoto et al. Neural network computation was applied to the Ternary silicate melts mold flux 4%
[15] estimation of surface tension in ternary silicate melts
Hanao et al. Hanao applied the ANN method to mold flux
[123]
(continued)
143
Table 5.7 (continued)
144

Source Description Systems Reported


deviation

Mills et al. [124] γ = X i γi + X j γ j Slags multicomponent 10%
γi is the recommended surface tension of component
i. γ j is the recommended surface tension of
surfactants j
whereas surfactants (j) are treated:
a: X j γ j = a + bX + C X 2 , for X j = N
b: X j γ j = a + bX, for X j > N
Overestimates the effect of surfactants when several
surfactants present
pure RT MiSurf
Nakamoto and γ = γi + Ai ln MiBulk
CaO–SiO2 –MgO–Al2 O3 –Na2 O–MnO–CaF2 –TiO2 10%
Tanaka
[125–129] R /R ·N P
MiP =  Ci Ai i P
Hanao et al. ( RC /RA ·N )
[130] where superscripts P = “Surf” and “Bulk” indicate
Lv et al. [131] the surface and bulk, respectively. RCi , RAi
correspond to the radii and anions of cations of
oxide i, respectively. NiP is the mole fraction of
oxide i in phase P
Zhang et al. Structure-based surface tension estimation model: CaO–SiO2 –Fex O system and its subsystems 6%
[132] Raman spectra were used to obtain the oxygen bond
database, and Butler’s equation was selected to build
the relationship between surface tension and
structure
5 Estimation of High Alumina Blast Furnace Slag Properties
5.5 Surface Tension 145

Table 5.8 Radii of cations and anions


Anions Radii (Å) Anions Radii (Å) Anions Radii (Å)
Si4+ 0.42 Mg2+ 0.66 Ti4+ 0.61
Ca2+ 0.99 Al3+ 0.51 O2- 1.44

Table 5.9 Molar volumes and surface tension of oxides as a function of temperature
Oxides Molar volumes (m3 /mol) Surfacetensions (mN/m)
SiO2 27.516{1 + 10 − 4(T − 1773)} × 10−6 243.2 + 0.031T
CaO 20.7{1 + 10 − 4(T − 1773)} × 10−6 791 − 0.0935T
MgO 16.1 {1 + 10 − 4(T − 1773)} × 10−6 1770 − 0.636T
Al2 O3 28.3{1 + 10 − 4(T − 1773)} × 10−6 1024 − 0.177T
TiO2 22.2{−4.689 × 10 − 5(T − 1023)} × 10−6 1384.3 − 0.6254T

pure RT M Surf
γ = γi + ln iBulk (5.46)
Ai Mi

RCi /R Ai · Ni
P
MiP = 
RSi4+ /RSiO4− · NSiO
P
2
+ RCa2+ /RO2− · NCaO
P +R
Mg2+ /RO2− · NMgO + RAl3+ /RO2− · NAl2 O3
P P
4
(5.47)

where the subscript i is the slag component (CaO, SiO2 , MgO, Al2 O3 , TiO2 ). The
[SiO4 4− ] is considered to be the ionic unit in which SiO2 component exists and
RSi4+ /RSiO4−
4
is usually taken as 0.5. The model assumes that the excess Gibbs free
energy in the Butler’s equation is ignored, and the ionic structure in ionic solutions
tends to depend on the ratio of radii between cations and anions. Therefore, the
anion-cation radius ratio is considered in the model to estimate the ionic structure
and chemical properties of the ionic melt. The radii of anions and cations in slag,
molar volume, and surface tension of pure components as a function of temperature
are shown in Tables 5.8 and 5.9.
The iso-surface tension curves of CaO–SiO2 –MgO–Al2 O3 –TiO2 slag system at
1500 °C are shown in Fig. 5.19. The yellow area within the red line in the figure is
the liquid phase area calculated by FactSage, the dotted line is the iso-surface tension
line under the corresponding composition calculated by the model, and the point is
the slag composition of the slag system reported in Chap. 4. It can be seen that the
surface tension of the slag gradually decreases with the increase of SiO2 content,
indicating that SiO2 acts as a surfactant in the slag. With increasing of SiO2 content,
more SiO2 is required for every 50 mN/m decreasing of the surface tension of the
slag, indicating that the effect of SiO2 as a surfactant is decreasing. CaO and Al2 O3
can increase the surface tension of slag due to their large surface tension values (Table
5.9). Furthermore, at a fixed SiO2 content, especially in the high-silicon region, the
replacement of Al2 O3 by CaO resulted in an increase in surface tension, indicating
146 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.19 Iso-surface


tension curves of CaO–
SiO2 –MgO–Al2 O3 –TiO2
slag system at 1500 °C.
Reprinted from Ref. [25]

that the effect of CaO on the surface tension is greater than that of Al2 O3 in the
silicate system.

5.6 Sulfide Capacity

5.6.1 Overview of Sulfide Capacity Models

After the sulfide capacity was proposed by Fincham and Richardson [133], metallur-
gists have developed various semi-empirical and empirical models for estimating the
sulfide capacity. Table 5.10 lists some of the models used to calculate the slag sulfide
capacity. Duffy et al. [134] first proposed the application of optical basicity to slag
capacity calculations, many sulfide capacity models were developed based on optical
basicity. Later, thanks to the abundance of thermodynamic data and experimental
data, thermodynamic models and neural network models were developed.

5.6.2 Structure-Based Sulfide Capacity Modeling

5.6.2.1 Theory of Model

According to the Eq. 1.6, the sulfide capacity can be expressed as:

ln Cs = ln K θ + ln aO2− − ln γS2− (5.48)

G de−s = −RT ln K θ (5.49)


Table 5.10 Summary of models to calculate the slags sulfide capacity
Type Source Description Systems
Optical basicity Duffy log CS = −11.9 + 12 Slags multicomponent
models et al.
[134]
Tsao and log CS = 14.2 − 9894
T − 7.55 CaO–SiO2 –MgO–Al2 O3
Katayama
5.6 Sulfide Capacity

 = 0.24 X CaO + 0.1X MgO − 0.8X Al2 O3 − X SiO2 + 0.67


[135]


Sosinsky log CS = 22,690−54,640


T + 43.6 − 25.2 Slags multicomponent
[136]
Young  < 0.8 : Slags multicomponent
et al.
2
[137] log CS = −13.913 + 42.84 − 23.82
 
11,710
− − 0.0223(wt% SiO2 )
T
− 0.2275(wt% Al2 O3 )
 ≥ 0.8 :
CS = −0.6261 + 0.4808 + 0.71972
1697 2587
+ − − 0.2275(wt% FeO)
T T

CS = 9.852 × 10−6 (wt% Al2 O3 )


Shankar 16.2933 High alumina BF slag
et al. + 0.010574 − + 0.002401
T
[138]
(continued)
147
Table 5.10 (continued)
148

Type Source Description Systems

log CS = 7.35 + 94.89 log 


11,710,051 + (−338(wt% MgO) + 287(wt% MnO))

Taniguchi T CaO–SiO2 –MgO–MnO–Al2 O3
et al. + 0.2284(wt% SiO2 ) + 0.1379(wt% Al2 O3 )
[139]
− 0.0587(wt% MgO) + 0.0841(wt% MnO)

4.49 15,864
Zhang log CS = −6.08 +  + 15,893 −  /T CaO–SiO2 –MgO–Al2 O3 –FeO–MnO–TiO2 –CaF2
et al.
[140]
−47,236.6+52,273.7corr
Ren et al. log CS = 16.89 − 19.46corr + T CaO–SiO2 –MgO–Al2 O3 –TiO2
[141] Where is the modified optical basicity by considering the
corr
charge compensation of Al3+
12,410
corr
−27,109 11.85
Hao et al. log CS = T + 19.45 − corr CaO–SiO2 –MgO–Al2 O3
[142] Secondary refining slag

100MS K M aMO 1−2X SiO2
Thermodynamic Reddy Cs = X SiO ≤ 1/3 Binary silicate
MMO +X SiO2 MSiO2 −MMO 2
models et al 
100MS K M aMO X SiO2 S
.[143] Cs = X SiO2 ≥ 1/3
MMO +X SiO2 MSiO2 −MMO aMS
where W i is the molecular weight of compound i
Nzotta RT ln Slags multicomponent
 0

et al. ξ
CS = exp −G RT exp RT
[144–147]

ξ= X i ξi + ξmix
WhereG 0 is the Gibbs energy change of desulfurization reaction.
ξ is a parameter to take account of the mutual interactions between
different oxides
Pelton A model for the solubility of gaseous species in molten slags within CaO–SiO2 –MgO–MnO–Al2 O3 –FeO–Fe2 O3 –TiO2 –Ti2 O3 multicomponent
and Kang the framework of the modified quasi-chemical model (MQM) in the slags
[148] quadruplet approximation
5 Estimation of High Alumina Blast Furnace Slag Properties

(continued)
Table 5.10 (continued)
Type Source Description Systems
 
 θ  935.45
Yang et al. log Cs = lg 16 K MS X MO n i + T − 1.3757 Slags multicomponent
[149–151] i

Zhang logC S = −4.638 + 0.506 log(%O2− ) − 3.116 log (%O0 ) CaO–SiO2 –Al2 O3 –MgO–TiO2
et al.
5.6 Sulfide Capacity

[152]
Neural network Derin Neural network approach has been used to predict sulfide capacities Multicomponent slags and flux
models (NN) et al. [16] of slag
Ma et al. Neural network approach has been used to predict sulfide capacities CaO–SiO2 –MgO–Al2 O3
[17] of slag
Xin et al. A regularized extreme learning machine (RELM) model was CaO–SiO2 –MgO–Al2 O3
[14] established to predict the sulfide capacity
149
150 5 Estimation of High Alumina Blast Furnace Slag Properties

Table 5.11 Standard Gibbs


Reactions Gibbs free energy (J/mol)
free energy change of
desulfurization reaction MgO + 1/2S2 = MgS +1/2 O2 G = 192068-13.854T
CaO +1/2 S2 = CaS + 1/2O2 G = 97301-7.299T

It is shown that the following two factors play important roles in desulfurization
process: the distribution of different types of oxygen, and the relative stability of
sulfide during the O2− ↔ S2− reaction. G de−s is the Gibbs free energy of desul-
furization reaction. To calculate the sulfide capacity via Eq. 5.48, it is necessary to
obtain the Gibbs free energy (G de−s ) of the desulfurization reaction, the activity of
free oxygen (aO2− ), and the activity coefficient S2− (γS2− ).
Considering that the activity coefficient of oxygen ions and the activity coefficient
of sulfur ions are not constant in the slag, then correction factors (A, B, and C) are
introduced:
G de−s
log Cs = A log NO2− − B −C (5.50)
2.303RT

where, NO2− is the concentration of free oxygen (O2− ) in the slag, which can be calcu-
lated as described in Sect. 5.3.2. The Gibbs free energy of desulfurization reaction of
CaO–SiO2 –Al2 O3 –MgO slag system can be approximated by the linear coupling of
the Gibbs free energies of the desulfurization reactions of each component. Specif-
ically, it can be calculated from the following equation. The standard Gibbs free
energy of desulfurization reactions is listed in Table 5.11.

G de−s = NCa2+ G θCa2+ + NMg2+ G θMg2+ (5.51)

To establish an effective sulfide capacity database, the experimental data sets


of the CaO–SiO2 –Al2 O3 –MgO system and its subsystems for using in the model
optimization also have been critically reviewed. Table 5.14 presents the sources
of the collected data (see Appendix). Slag-gas equilibrium method and slag-metal
equilibrium method are involved in the table. It should be noted that part of the
reported sulfide capacity data was conducted on solid–liquid mixture which was not
considered for optimization (liquidus were predicted by FactSage). The data reported
by different researchers at closed compositions were carefully compared to cross-
check the reliability of the data. According to the analysis of the above evaluation,
more than 500 sets of effective viscosity data from nearly 60 literatures were used
for the optimization of model parameters.
After calculating the concentration of free oxygen (O2− ), and the Gibbs free
energy of desulfurization reaction, the correction factors in Eq. 5.50 are obtained by
fitting with the data set.
5.6 Sulfide Capacity 151

Fig. 5.20 Comparison between experimental data and calculated data for CaO–SiO2 –Al2 O3 –MgO
system and its subsystems

5.6.2.2 Performance of Sulfide Capacity Model

To evaluate the performance of the structural-based sulfide capacity model, the


average of relative errors  was used, as expressed by:

1  |Csest − Csmeas |
N
= × 100% (5.52)
N n=1 Csmeas

where Csmeas and Csest are the measured and estimated sulfide capacity values,
respectively. N is the number of sulfide capacity data.
The CaO–MgO–Al2 O3 –SiO2 system is a fundamental slag system for ironmaking
process. The comparison between experimental data and calculated data is shown in
Fig. 5.20 and shows a good prediction performance with an average of relative errors
less than 6%.

5.6.3 Iso-sulfide Capacity of High Alumina Blast Furnace


Slag

According to the developed model, the iso-sulfide capacity diagram of CaO–SiO2 –


9 wt% MgO–Al2 O3 –1 wt% TiO2 slag system at 1500 °C was calculated, as shown in
Fig. 5.21. In this case, 1% of TiO2 was ignored. It is found that the sulfide capacity of
152 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.21 Iso-sulfide


capacity lines in the
CaO–SiO2 –MgO–Al2 O3
slag system at 1500 °C

slag increases with the increase of CaO content when SiO2 is constant. The sulfide
capacity increases with increasing Al2 O3 content at fixed basicity. In the case of fixed
CaO content, the replacement of SiO2 with Al2 O3 increases the sulfide capacity and
enhances the desulfurization ability of the slag. When Al2 O3 content is more than
24 wt% with low SiO2 , CaO content ranges from 35 to 45 wt%, and the sulfide
capacity (log(Cs)) is about −3.4 which is good for desulfurization.

5.7 Electrical Conductivity

5.7.1 Overview of Electrical Conductivity Models

The various models for estimating the electrical conductivities of slags are listed
in Table 5.12. It is should be noted that these reported models are mainly focused
on the silicate slag systems, where the ionic conduction is the dominant. Jiao and
Themelis [153] reported the relationship between electrical conductivity and slag
composition of SiO2 –CaO–MgO–MnO and SiO2 –CaO–MgO–FeO slag systems at
1500 °C; Wang [154] developed an electrical model based on the mass triangle
model. Zhang and Chou [155] gave the relationship between optical basicity and slag
conductivity according to the Arrhenius formula, which can effectively calculate the
conductivity of SiO2 –CaO–Al2 O3 (–MgO) slag. In addition, high viscosity results in
a high electrical resistance, Zhang and Chou [156] provided the relationship between
conductivity and viscosity for aluminosilicate melts.
5.7 Electrical Conductivity 153

Table 5.12 Summary of models to calculate the slags electrical conductivities


Source Description Systems Reported
deviation
Jiao and ln κ = CaO–SiO2 –MgO–MnO–Al2 O3 –FeO–Fe2 O3 /
Themelis a + bX FeO + c X CaO + X MgO
[153] where a is a constant. b and c
are the concentration
coefficients

X M2+ IM2+
Wang κ =a X Si4+ ISi4+ −b CaO–MgO–SiO2 , Na2 O–MgO–SiO2 , /
[154] CaO–FeO–SiO2
where a, and b are constants.
This model is based on the
mass triangle model
ln κ = ln A + mRT +n
corr
Zhang CaO–SiO2 –MgO–Al2 O3 14%
and where A, m, and n are constants
Chou κ as a function of corrected
[155] optical basicity
Zhang ln κ = A + B ln η MxO–SiO2 (M = Mg, Ca, Sr, Ba, Li, Na, K) 21.8%
and where A and B are constants MxO–Al2 O3 –SiO2
Chou κ as a function of viscosity
[156]
    
1 Q
log = a + k1 exp
κ t1
 
Mills Q Slags /
+k2 exp
et al. t2
[157]
− k3 N + k2 r 3
κ as a function of degree of
polymerization for the silicate
network (Q), the number of
available cations (N) and size
of cations (r) are taken into
consideration. a, and the
various k and t terms are
constants

5.7.2 Iso-electrical Conductivity of High Alumina Blast


Furnace Slag

Based on the experimental values of viscosity and conductivity, the relationship


between was established.

ln κ = −0.71 − 0.66 ln η (5.53)


154 5 Estimation of High Alumina Blast Furnace Slag Properties

Fig. 5.22 Iso-electrical conductivity lines in the CaO–SiO2 –MgO–Al2 O3 –TiO2 slag system at
1500 °C

According to Eq. 5.48, the iso-electrical conductivity diagram of CaO–SiO2 –


Al2 O3 –9 wt% MgO system was calculated as shown in Fig. 5.22. The viscosity was
calculated by using Reddy-Yan model, and more details can be found in sect. 5.3.
The gray area in the figure presents the liquid region, and the dotted line is the iso-
electrical conductivity. As shown in the figure, the electrical conductivity increases
with increasing CaO content and decreases with increasing SiO2 content. The elec-
trical conductivity decreases with the increasing Al2 O3 content at a fixed CaO/SiO2
ratio. There is no significant influence on the viscosity by using the Al2 O3 substitute
for SiO2 at fixed CaO content when Al2 O3 is less than 50 wt%, resulting in the same
trend in electrical conductivity.

Appendix

See Tables 5.13 and 5.14.


Table 5.13 Summary of the viscosity data for the CaO–MgO–FeO–SiO2 –Al2 O3 system and its subsystems
Slag system Resources Method PEA Materials Temperature (K) Data set Accepted data
Appendix

Crucible Spindle
Al2 O3 Kozakevitch RS; Ar; p/m, t/c N/A Mo, W Mo, W 2323–2373 2 2
et al. [1]
Elioutin et al. Vacuum; p/m Yes Mo – 2325–2775 6 6
[2]
Blomquist OM; vacuum; N/A W N/A 2394–2742 11 11
et al. [3] p/m
Urbain et al. [4] RC; Ar + H2 ; Yes Mo Mo 2328–2558 15 15
p/m
SiO2 Bockris et al. RC; Ar; p/m Yes Mo Mo 2173–2325 10 0
[5]
Bruckner et al. RS; Air; t/c N/A Ir Ir 1959–2279 4 4
[6]
Rossin et al. [7] RC; Ar; p/m Yes Mo Mo 2340–2392 8 8
Hofmaier et al. RC; Ar; p/m Yes Mo Mo 1878–2538 11 11
[8]
Leko et al. [9] RC; Ar; p/m Yes Mo Mo 1973–2275 4 4
Loryan et al. RC; Ar; p/m Yes Mo Mo 1973–2273 4 4
[10]
Urbain et al. [4] RC; Ar; p/m Yes Mo Mo 1465–2755 15 15
CaO–SiO2 Machin et al. OM; Air; t/c N/A Pt Pt 1723–1773 7 4
[11–14]
(continued)
155
Table 5.13 (continued)
156

Slag system Resources Method PEA Materials Temperature (K) Data set Accepted data
Crucible Spindle
Bockris et al. RC; Ar; p/m Yes Mo Mo 1723–2073 88 80
[15]
Kozakevitch RS; Ar; p/m, t/c Yes Mo, W Mo, W 1823–2373 21 20
et al. [16]
Urbain et al. [4] RC; Ar; p/m Yes Mo Mo 1825–2393 33 33
Licko et al. OM; Air; t/c N/A Pt-Rh Pt-Rh 1673–1973 8 5
[17]
Hu et al. [18] RS; Ar; t/c N/A Mo Mo 1973 6 6
Neuville et al. RS; Air; t/c Yes Pt-Rh Pt-Rh 1012–1809 22 11
[19]
MgO–SiO2 Bockris et al. RC; Ar; p/m Yes Mo Mo 1823–2073 30 25
[5]
Hofmaier et al. RC; Ar; p/m Yes Mo Mo 1873–2073 9 9
[8]
Urbain et al. RC; Ar; p/m Yes Mo Mo 1873–2073 9 9
[20]
Hu et al. [18] RS; Ar; t/c N/A Mo Mo 1973–2073 15 15
FeO–SiO2 Chen et al. [21] RS; Ar, EMF; Yes Mo Mo 1473–1773 30 14
t/c
Shiraishi et al. RS; Ar, EMF; Yes Fe Fe 1523–1673 40 33
[22] p/m, t/c
(continued)
5 Estimation of High Alumina Blast Furnace Slag Properties
Table 5.13 (continued)
Slag system Resources Method PEA Materials Temperature (K) Data set Accepted data
Appendix

Crucible Spindle
Kaiura et al. RS; CO/CO2 , No Mo Mo 1473–1623 28 23
[23] t/c
Kucharski et al. RS; CO/CO2 , No Mo Mo 1473–1623 192 176
[24] t/c
Urbain et al. [4] RC; Yes Fe Fe 1439–1733 54 49
H2 /CO/CO2 , t/c
Ji et al. [25] RS; Ar, EMF; Yes Fe Fe 1503–1753 9 9
t/c
Al2 O3 –SiO2 Kozakevitch RS; Ar; p/m, t/c Yes Mo, W Mo, W 2123–2373 15 15
et al. [16]
Urbain et al. [4] RC; Ar; p/m Yes Mo Mo 1926–2477 32 29
CaO–SiO2 –MgO Machin et al. OM; Air; t/c N/A Pt Pt 1623–1773 47 33
[11–14]
Urbain et al. [4] RC; Ar; p/m Yes Mo Mo 1675–2313 7 7
Scarfe et al. RS; Air; t/c Yes Pt-Rh Pt-Rh 1648–1873 6 5
[26]
Licko et al. OM; Air; t/c N/A Pt-Rh Pt-Rh 1673–1973 52 44
[17]
Kim et al. [27] RS; Ar; t/c Yes Pt-Rh Pt-Rh 1648–1773 6 4
CaO–SiO2 –Al2 O3 Machin et al. OM; Air; t/c N/A Pt Pt 1423–1773 282 151
[11–14]
Kozakevitch RS; Ar; p/m, t/c Yes Mo, W Mo, W 1723–2273 206 198
et al. [1]
Ohno et al. [28] RS; Ar; t/c N/A Mo Mo 1573–1773 13 9
157

(continued)
Table 5.13 (continued)
158

Slag system Resources Method PEA Materials Temperature (K) Data set Accepted data
Crucible Spindle
Skryabin et al. RS; Air; t/c Yes ZrO2 Pt 1673–1924 49 35
[29]
Cukierman RC; Ar; p/m N/A Mo Mo 1560–2340 22 0
et al. [30]
Urbain et al. [4] RC; Ar; p/m Yes Mo Mo 1578–2456 76 72
Scarfe et al. RS; Air; t/c Yes Pt-Rh Pt-Rh 1648–1898 11 5
[26]
Taniguchi et al. FS; AIr; t/c N/A Pt Pt 1118–1853 12 0
[31]
Solvang et al. RS; Air; t/c Yes Pt-Rh Pt-Rh 1558–1849 99 64
[32]
Toplis et al. RS; Air; t/c Yes Pt-Rh Pt-Rh 1573–1953 258 161
[33]
Matysek et al. RS; N2 ; t/c N/A C C 1803 9 0
[34]
Li et al. [35] RS; Ar; t/c N/A Mo Mo 1733–1823 3 3
CaO–SiO2 –FeO Chen et al. [36] RS; Ar, EMF; Yes Mo Mo 1423–1773 57 57
t/c
Ji et al. [25] RS; Ar, EMF; Yes Fe Fe 1423–1753 48 37
t/c
Kucharski et al. RS; CO/CO2 , No Mo Mo 1473–1623 464 464
[24] t/c
Williams et al. RS; Ar, EMF; No Fe Fe 1423–1673 10 10
[37] t/c
5 Estimation of High Alumina Blast Furnace Slag Properties

Johannsen FS; N2 , EMF; Yes Fe Pt 1423–1673 175 171


et al. [38] t/c
(continued)
Table 5.13 (continued)
Slag system Resources Method PEA Materials Temperature (K) Data set Accepted data
Appendix

Crucible Spindle
SiO2 –Al2 O3 –MgO Machin et al. OM; Air; t/c N/A Pt Pt 1523–1773 27 19
[11–14]
Riebling et al. FS; Ar; t/c Yes Pt-Rh Pt-Rh 1843–1993 68 68
[39]
Urbain et al. [4] RC; Ar; p/m Yes Mo Mo 1803–2388 29 28
Toplis et al. RS; Air; t/c Yes Pt-Rh Pt-Rh 1673–1913 96 72
[33]
Mizoguchi RS; Ar; t/c Yes Mo Mo 1773–1848 40 40
et al. [40]
Al2 O3 –SiO2 –FeO Chen et al. [41] RS; Ar, EMF; Yes Mo Mo 1473–1773 48 48
t/c
Rontgen et al. RS; N2 , EMF; Yes Fe Pt 1523 9 5
[42] t/c
Kurochkin Vibrational; Ar; No Al2 O3 / 1373–1723 16 0
et al. [43] t/c
Sheludyakov Vibrational; Ar; No Pt Pt 1653–1683 4 0
et al. [44] t/c
MgO–SiO2 –FeO Chen et al. [45] RS; Ar, EMF; Yes Mo Mo 1523–1773 52 52
t/c
Kucharski et al. RS; CO/CO2 , No Mo Mo 1543–1623 57 57
[24] t/c
Ji et al. [46] RS; Ar, EMF; Yes Fe/Mo Fe/Mo 1608–2031 32 28
t/c
Toropov et al. RS; Ar; t/c Yes Mo Mo 1523–1773 63 0
[47]
159

(continued)
Table 5.13 (continued)
160

Slag system Resources Method PEA Materials Temperature (K) Data set Accepted data
Crucible Spindle
Williams et al. RS; Ar, EMF; No Fe Fe 1523–1673 4 4
[37] t/c
CaO–SiO2 –Al2 O3 –MgO Machin et al. OM; Air; t/c N/A Pt Pt 1423–1773 512 339
[11–14]
Scarfe et al. RS; Air; t/c Yes Pt-Rh Pt-Rh 1448–1873 58 41
[26]
Forsbacka et al. RS; Ar-5%CO; Yes Mo Mo 1853–2023 52 49
[48] t/c
Saito et al. [49] RS; Air; t/c Yes Pt-Rh Pt-Rh 1673–1873 20 12
Lee et al. [50] RS; Ar; t/c Yes Pt-Rh Fe 1673–1773 12 10
CaO–SiO2 –Al2 O3 –MgO Shankar et al. RS; Ar; t/c Yes Mo Mo 1673–1873 30 25
[51]
Kim et al. [52] RS; Ar; t/c Yes Pt-Rh Pt-Rh 1723–1773 13 9
Song et al. [53] RS; Ar; t/c Yes Mo Mo 1727–1903 63 36
Tang et al. [54] RS; Ar; t/c N/A Mo Mo 1696–1798 60 59
Liao et al. [55] RS; Ar; t/c Yes Mo Mo 1748–1773 12 10
Kim et al. [27] RS; Ar; t/c Yes Pt-Rh Pt-Rh 1598–1773 72 59
Chen et al. [56] RS; Ar; t/c Yes Mo Mo 1673–1848 20 18
Gao et al. [57] RS; CO; t/c Yes C C 1763–1773 24 24
Li et al. [35] RS; Ar; t/c N/A Mo Mo 1733–1823 47 42
Yao et al. [58] RS; Ar; t/c Yes Mo Mo 1673–1773 107 57
Sun et al. [59] RS; Ar; t/c N/A Mo Mo 1723–1773 18 16
(continued)
5 Estimation of High Alumina Blast Furnace Slag Properties
Table 5.13 (continued)
Slag system Resources Method PEA Materials Temperature (K) Data set Accepted data
Appendix

Crucible Spindle
CaO–SiO2 –MgO–FeO Williams et al. RS; Ar, EMF; No Fe Fe 1473–1673 11 11
[37] t/c
Ji et al. [46] RS; Ar, EMF; Yes Fe Fe 1543–1773 36 36
t/c
CaO–SiO2 –Al2 O3 –FeO Azuma et al. FS; Ar; t/c Yes Fe Pt 1523–1573 30 30
[60]
Johannsen FS; N2 , EMF; Yes Fe Pt 1473–1673 54 54
et al. [38] t/c
Hurst et al. [61] RS; Reducing + Yes Mo Mo 1673–1773 156 111
N2 ; t/c
Gimmelfarb Torsional; Ar, Yes Al2 O3 Fe 1323–1623 417 149
et al. [62] EMF p/m
Higgins et al. RC; N2 , EMF; No Mo Mo 1473–1773 47 46
[63] t/c
CaO–SiO2 –Al2 O3 –MgO–FeO Kim et al. [64] RS; Ar, EMF; No Pt Pt 1723–1773 44 34
t/c
Bills et al. [65] RS; Ar, EMF; No Ir Ir 1523–1773 82 77
t/c
Lee et al. [66] RS; Ar, EMF; No Pt Fe 1673–1723 57 26
t/c
Higgins et al. RC; N2 , EMF; No Mo Mo 1423–1773 48 47
[63] t/c
PEA Post-experimental analysis; EMF Equilibrium with metallic Fe; FS Falling sphere method; RC or RS Rotational crucible or spindle; OM Oscillation method
t/c Thermocouple; p/m Pyrometer
161
Table 5.14 Summary of the sulfide capacity data for the CaO–MgO–FeO–SiO2 –Al2 O3 system and its subsystems
162

Slag system Resources Experiment Gas Gas Crucible Temperature Data set Accepted
partial (K) data
pressure
CaO–SiO2 Fincham et al. SG H2 –CO2 –SO2 –N2 N/Y Pt 1773 1873 8(5/3) 7
[67] 1923
Carter et al. SG CO–CO2 –SO2 Y Pt 1773 10 10
[68]
Abraham SG H2 –CO2 –SO2 –N2 N Pt 1773 1848 4 4
et al. [69–70] 1923
Brown et al. SG H2 –CO2 –SO2 N Pt 1773 8 8
[71]
Gornerup SG CO–CO2 –SO2 –Ar Y Pt 1823 1873 6 6
et al. [72] 1923
Park et al. SG CO–CO2 –SO2 –Ar Y Pt 1873 2 2
[73]
CaO–Al2 O3 Fincham et al. SG H2 –CO2 –SO2 –N2 Y Pt 1923 3 3
[67]
Carter et al. SG CO–CO2 –SO2 Y Pt 1773 14 14
[74]
Sharma et al. SG H2 –CO2 –SO2 –N2 N Pt 1773 5 5
[75]
Cameron SG CO–CO2 –SO2 /H2 –H2 O–H2 S–Ar Y Pt 1823 7 7
et al. [76]
Kor et al. [77] SG H2 –CO2 –SO2 –H2 CO2 + CO + Y Pt or Ir 1623 1773 54 3
SO2 1823
(continued)
5 Estimation of High Alumina Blast Furnace Slag Properties
Table 5.14 (continued)
Slag system Resources Experiment Gas Gas Crucible Temperature Data set Accepted
partial (K) data
Appendix

pressure
Hino et al. SG CO–CO2 –SO2 –Ar Y CaO 1928 1 1
[78]
Drakaliysky SG CO–CO2 –SO2 –Ar Y Pt 1773 1848 21 21
et al. [79]
Ohta et al. SM(Cu) – – Mo 1873 3 3
[80]
Choi et al. SG CO–CO2 –SO2 –Ar Y Pt 1873 5 5
[81]
MgO–SiO2 Fincham et al. SG H2 –CO2 –SO2 –N2 Y Pt 1923 1 1
[67]
Sharma et al. SG H2 –CO2 –SO2 –N2 N Ir 1923 7 7
[82]
Nzotta et al. SG CO–CO2 –SO2 –Ar Y Pt 1853 1873 13 13
[83] 1923
Al2 O3 –SiO2 Sharma et al. SG H2 –CO2 –SO2 –N2 N Ir 1923 11 11
[82]
CaO–SiO2 –Al2 O3 Fincham et al. SG H2 –CO2 –SO2 –N2 N Pt 1773 6 6
[67]
Thomas et al. SG CO–CO2 –SO2 Y Pt 1773 12 12
[84]
Carter et al. SG CO–CO2 –SO2 Y Pt 1773 2 2
[74]
(continued)
163
Table 5.14 (continued)
164

Slag system Resources Experiment Gas Gas Crucible Temperature Data set Accepted
partial (K) data
pressure
Kalyanram SG CO–CO2 –SO2 Y Pt 1773 39 39
et al. [85]
Cameron SG CO–CO2 –SO2 /H2 –H2 O–H2 S–Ar Y Pt or Ir 1823 27 27
et al. [76]
Brown et al. SG/SM H2 –CO2 –SO2 /– Y/– Pt 1773 2(1/1) 2
[86]
Kim Kirsrud SG CO–CO2 –SO2 –N2 N Pt 1773 1 1
[87]
Hino et al. SG CO–CO2 –SO2 –Ar Y Pt 1872 1873 10 10
[78]
Gornerup SG CO–CO2 –SO2 –Ar Y Pt 1823 1873 33 33
et al. [72] 1923
Drakaliysky SG CO–CO2 –SO2 –Ar Y Pt 1773 1848 20 20
et al. [79] 1873 1923
Ohta et al. SM(Cu) – – Pt 1873 3 3
[80]
Taniguchi SG CO–CO2 –SO2 –Ar Y Pt 1673 1 1
et al. [88]
Taniguchi SG CO–CO2 –SO2 –Ar Y Pt 1673 1723 18 18
et al. [89] 1773
Adolfo et al. SM(Cu) – – Mo 1873 22 22
[90]
(continued)
5 Estimation of High Alumina Blast Furnace Slag Properties
Table 5.14 (continued)
Slag system Resources Experiment Gas Gas Crucible Temperature Data set Accepted
partial (K) data
Appendix

pressure
Choi et al. SG CO–CO2 –SO2 –Ar Y Pt 1873 52 52
[81]
CaO–SiO2 –MgO Abraham SG H2 –CO2 –SO2 –N2 N Pt 1923 5 5
et al. [69–70]
Kalyanram SG CO–CO2 –SO2 Y Pt 1773 42 42
et al. [91]
Nzotta et al. SG CO–CO2 –SO2 –Ar Y Pt 1773 1823 26 26
[83] 1873
O’Neill et al. SG CO–CO2 –SO2 N Pt 1673 1 0
[92]
CaO–Al2 O3 –MgO Hino et al. SG CO–CO2 –SO2 –Ar Y Pt or CaO 1822 1823 38 38
[93] 1866 1867
1868 1922
1923 1924
1928
Nzotta et al. SG CO–CO2 –SO2 –Ar Y Pt 1773 1823 10 10
[94] 1873
Ohta et al. SM(Cu) – – Pt 1873 1 1
[80]
SiO2 –Al2 O3 –MgO Sharma et al. SG H2 –CO2 –SO2 –N2 N Ir 1923 26 26
[82]
Nzotta et al. SG CO–CO2 –SO2 –Ar Y Pt 1773 1823 17 17
[94] 1873 1923
(continued)
165
Table 5.14 (continued)
166

Slag system Resources Experiment Gas Gas Crucible Temperature Data set Accepted
partial (K) data
pressure
CaO–SiO2 –MgO–Al2 O3 Abraham SG H2 –CO2 –SO2 N Pt 1773 4 4
et al. [95]
Kalyanram SG CO–CO2 –SO2 Y Pt 1773 27 27
et al. [85]
Kim Kirsrud SG CO–CO2 –SO2 –N2 N Pt 1773 3 3
[87]
Drakaliysky SG CO–CO2 –SO2 –Ar Y Pt 1773 1823 13 13
et al. [96]
Tang et al. SM – – C 1773 3 3
[97]
Nzotta et al. SG CO–CO2 –SO2 –Ar Y Pt 1773 1823 10 10
[98] 1923
Seo et al. [99] SG CO–CO2 –SO2 –Ar Y Pt 1773 10 10
O’Neill et al. SG CO–CO2 –SO2 N Pt 1673 47 32
[92]
M Ohta et al. SM(Cu) – – Pt 1873 16 16
[80]
Shankar et al. SG CO–CO2 –SO2 –Ar Y Pt 1773 1823 27 27
[100] 1873
Hayakawa SG CO–CO2 –SO2 –Ar Y Pt (1673, 1823, 28 28
et al. [101] 25)*
Taniguchi SG CO–CO2 –SO2 –Ar Y Pt 1673 1723 7 7
et al. [88] 1773
(continued)
5 Estimation of High Alumina Blast Furnace Slag Properties
Table 5.14 (continued)
Slag system Resources Experiment Gas Gas Crucible Temperature Data set Accepted
partial (K) data
Appendix

pressure
Yang et al. SM – – C 1773 13 7
[102]
Park et al. SG CO–CO2 –SO2 –Ar Y Pt 1873 3 3
[103]
Allertz et al. SM(Cu) – – Mo 1823 1873 26 26
[104]
Wang et al. SG CO–CO2 –SO2 –Ar Y Pt 1823 1848 6 6
[105] 1873 1898
Ma et al. SG CO–CO2 –SO2 –Ar Y Pt 1773 1823 32 32
[106]
Adolfo et al. SM(Cu) – – Mo 1823 1873 17 17
[90]
Trinath et al. SM – – C 1773 8 8
[107]
Adolfo et al. SM(Cu) – – Mo 1713 1743 39 39
[108] 1773
SG Slag-Gas equilibrium method; SM Slag-Metal equilibrium method; (1673, 1823, 25)*: 1673 is the start temperature, 1823 is the end temperature, 25 is the
step.
Y Yes (with gas partial pressure information); N No (no gas partial pressure information)
167
168 5 Estimation of High Alumina Blast Furnace Slag Properties

References (appendix)
1. Kozakevitch, P. (1961). Viscosity of lime-aluminasilica melts between 1600°
and 2100° C. Proceedings Metallurgical Society Conferences, 7, 97–116.
2. Elioutin, B. C. M. V. P., & Nagibin, Y. A. (1972). Properties of liquid
aluminum oxide. Fizika Aerodispersnykh Sistem, 7, 129–135.
3. Blomquist, R. A., Fink, J. K., & Leibowitz, L. (1978). Viscosity of molten
alumina. Technical Report Archive & Image Library, 57.
4. Urbain, G., Bottinga, Y., & Richet, P. (1982). Viscosity of liquid silica,
silicates and alumino-silicates. Geochimica et Cosmochimica Acta, 46(6),
1061–1072.
5. Bockris, J. O. M., Mackenzie, J. D., & Kitchener, J. A. (1955). Viscous flow
in silica and binary liquid silicates. Transactions of the Faraday Society, 51,
1734–1748.
6. Bruckner, R. (1964). Characteristic physical properties of the principal glass-
forming oxides and their relation to glass structure. Glass Science and
Technology , 37, 413–425.
7. Rossin, R., Bersan, J., & Urbain, G. (1964). Viscosity of fused silica and
slags belonging to the SiO2 -Al2 O3 system. Revue Internationale des Hautes
Temperatures et des Refractaires, 1, 159–170.
8. Hofmaier, U. G. (1968). Viscosity of pure silica. Science of Ceramics, 4,
25–32.
9. Leko, V. K., Meshcheryakova, E. V., Gusakova, N. K., & Lebedeva, R. B.
(1974). Optiko-Mekhanicheskiy Promyshlennost, 12, 42–45.
10. Loryan, S. G., Kostanyan, K. A., Saringyulyan, R. S., Kafyrov, V. M., &
Bogdasaryan, E. Kh. (1976).Elektron. Tekh. Ser. 6, Mater, 2, 53.
11. Machin, J. S., & Yee, T. B. (1954). Viscosity studies of system CaO-MgO-
Al2 O3 -SiO3 : IV, 60 and 65% SiO2 . Journal of the American Ceramic Society,
37(4), 177–186.
12. Machin, J. S., Yee, T. B., & Hanna, D. L. (1952). Viscosity studies of system
CaO-MgO-Al2 O3 -SiO2 : III, 35, 45, and 50% SiO2 . Journal of the American
Ceramic Society, 35(12), 322–325.
13. Machin, J. S., & Yee, T. B. (1948). Viscosity studies of system CaO-MgO-
Al2 O3 -SiO2 : II, CaO-Al2 O3 -SiO2 . Journal of the American Ceramic Society,
31(7), 200–204.
14. Machin, J. S., & Hanna, D. L. (1945). Viscosity studies of system CaO–
MgO–Al2 O3 –SiO2 : 1, 40% SiO2 . Journal of the American Ceramic Society,
28(11), 310–316.
15. Bockris, J. O. M., & Lowe, D. C. (1954). Viscosity and the structure of molten
silicates. Proceedings of the Royal Society A, 226(1167), 423–435.
16. Kozakevitch, P. (1960). Viscosité et éléments structuraux des aluminosili-
cates fondus: Laitiers CaO-AI2 O3 -SiO2 entre 1600 et 2100 °C. Revue de
Métallurgie, 57, 149–160.
17. Licko, V. D. T. (1986). Viscosity and structure of melts in the system CaO-
MgO-SiO2 . Physics and Chemistry of Glasses, 27.
Appendix 169

18. Hu, H., & Reddy, R. G. (1990). Modeling of viscosities of binary alkaline
earth metal oxide and silicate melts. High Temperature Science, 28, 195–202.
19. Neuville, D. R. (2006). Viscosity, structure and mixing in (Ca, Na) silicate
melts. Chemical Geology, 229(1), 28–41.
20. Urbain, G. (1974). Viscosity and structure of liquid aluminosilicates:
Measurement methods and results. Revue Internationale des Hautes Temper-
atures et des Refractaires, 11, 133–145.
21. Chen, M., Raghunath, S., & Zhao, B. (2013). Viscosity measurements of
“FeO”-SiO2 slag in equilibrium with metallic Fe. Metallurgical & Materials
Transactions B, 44(3), 506–515.
22. Shiraishi, Y., Ikeda, K., Tamura, A., & Saito, T. (1978). On the viscosity and
density of the molten FeO-SiO2 system. Transactions of the Japan Institute
of Metals, 19(5), 264–274.
23. Kaiura, G. H., Toguri, J. M., & Marchant, G. (1977). Viscosity of fayalite-
based slags. Canadian Metallurgical Quarterly, 16(1), 156–160.
24. Kucharski, M., Stubina, N. M., & Toguri, J. M. (1989). Viscosity measure-
ments of molten Fe–O–SiO2 , Fe–O–CaO–SiO2 , and Fe–O–MgO–SiO2 slags.
Canadian Metallurgical Quarterly, 28(1), 7–11.
25. Ji, F. Z., Du, S., & Seetharaman, S. (1997). Experimental studies of
the viscosities in the CaO-FenO-SiO2 slags. Metallurgical & Materials
Transactions B, 28(5), 827–834.
26. Scarfe, C. M., Cronin, D. J., Wenzel, J. T., & Kauffman, D. A. (1983).
Viscosity-temperature relationships at 1 atm in the system diopside-anorthite.
American Mineralogist, 68(11–12), 1083–1088.
27. Kim, H., Matsuura, H., Tsukihashi, F., Wang, W., Min, D., & Sohn, I. (2013).
Effect of Al2 O3 and CaO/SiO2 on the viscosity of calcium-silicate–based
slags containing 10 mass Pct MgO. Metallurgical and Materials Transactions
B, 44(1), 5–12.
28. Ohno, A., & Ross, H. (1963). Optimum slag composition for the blast-furnace
smelting of titaniferous ores. Canadian Metallurgical Quarterly, 2(3), 259–
279.
29. Skryabin, V. G., & Novokhatskii, I. A. (1972). Cheminform abstract: Einfluss
von CO2 auf die viskositaet von schmelzen des syst. CaO-Al2 O3 -SiO2 .
Chemischer Informationsdienst, 3(43).
30. Cukierman, M., & Uhlmann, D. R. (1973). Viscosity of liquid anorthite.
Journal of Geophysical Research, 78(23), 4920–4923.
31. Taniguchi, H. (1992). Entropy dependence of viscosity and the glass-
transition temperature of melts in the system diopside-anorthite. Contribu-
tions to Mineralogy & Petrology, 109(3), 295–303.
32. Solvang, M., Yue, Y. Z., Jensen, S. L., & Dingwell, D. B. (2004). Rheological
and thermodynamic behaviors of different calcium aluminosilicate melts with
the same non-bridging oxygen content. Journal of Non-crystalline Solids,
336(3), 179–188.
33. Toplis, M. J., & Dingwell, D. B. (2004). Shear viscosities of CaO-Al2 O3 -
SiO2 and MgO-Al2 O3 -SiO2 liquids: Implications for the structural role of
170 5 Estimation of High Alumina Blast Furnace Slag Properties

aluminium and the degree of polymerisation of synthetic and natural alumi-


nosilicate melts. Geochimica et Cosmochimica Acta, 68(24), 5169–5188.
34. Řeháčková, L., Rosypalová, S., Dudek, R., Ritz, M., Matýsek, D., Smetana,
B., Dobrovská, J., Zlá, S., & Kawuloková, M. (2015). Effect of chem-
ical composition and temperature on viscosity and structure of molten
CaO-Al2 O3 -SiO2 system. Archives of Metallurgy and Materials, 60(4),
2873–2878.
35. Li, W., Cao, X., Jiang, T., Yang, H., & Xue, X. (2016). Relation between
electrical conductivity and viscosity of CaO–SiO2 –Al2 O3 –MgO system. ISIJ
International, 56(2), 205–209.
36. Chen, M., & Zhao, B. (2015). Viscosity measurements of SiO2 -“FeO”-
CaO system in equilibrium with metallic Fe. Metallurgical & Materials
Transactions B, 46(2), 577–584.
37. Williams, M. S. P., & Briggs, G. (1983). Viscosities of synthetic slags in
the system CaO-“FeO”-SiO2 -MgO. Institution of Mining and Metallurgy,
Section C, 105–109.
38. Johannsen, F., & Wiese, W. (1958). Das Absetzverhalten on Kupfer-
stein und Kupfer in fliissigen Schlacken. Zeitschrift fiir Erzbergbau und
Metallhiittenwesen, 1(48), 1–15.
39. Riebling, E. F. (1964). Structure of magnesium aluminosilicate liquids at
1700°C. Canadian Journal of Chemistry, 42(12), 2811–2821.
40. Mizoguchi, K., Okamoto, K., & Suginohara, Y. (1982). Oxygen coordi-
nation of Al**3** plus ion in several silicate melts studied by viscosity
measurements. Journal of the Japan Institute of Metals, 46(11), 1055–1060.
41. Chen, M., Raghunath, S., & Zhao, B. (2013). Viscosity of SiO2 -“FeO”-
Al2 O3 system in equilibrium with metallic Fe. Metallurgical and Materials
Transactions B, 44(4), 820–827.
42. Rontgen, H. W. P., & Kammel, R. (1956). Struktur und Eigenschaften von
Schlacken der Metallhuttenprozesse, I. Viskositätsmessungen an Schlacken
des Systems Eisenoxydul-Tonerde-Kieselsäure. Zeitschrift für Erzbergbau
und Metallhüttenwesen (5), 207–214.
43. Kurochkin, A., Pavlov, A., & Kvyatkovskii, A. (1983). Izuchenie vyazkosti
vysokokremnezemistykh oksidnykh rasplavov. Kompleksnoe ispol’zovanie
mineral’nogo syr’ya, 12, 33–36.
44. Sheludyakov, E. S. L., & Vakhitov, A. (1967). Vyazkost’alyuminosilikatnykh
rasplavov sistemy Mex Oy -Al2 O3 -SiO2 . Trudy Instituta Khimicheskikh Nauk,
158–164.
45. Chen, M., Raghunath, S., & Zhao, B. (2014). Viscosity measurements of
SiO2 -“FeO”-MgO system in equilibrium with metallic Fe. Metallurgical &
Materials Transactions B, 45(1), 58–65.
46. Ji, F. Z., Du, S., & Seetharaman, S. (1998). Experimental studies of the
viscosities in the FeO MgO SiO2 and FeO MnO SiO2 slags. Ironmaking &
Steelmaking, 25(4), 309–316.
47. Toropov, N. A., & Bryantsev, B. A. (1965). Physiochemical properties and
crystallization of melts in the system magnesium oxide-ferrous oxide-silica.
Appendix 171

In: The structure of glass, Vol. 5. Structural transformations in glasses at high


temperatures, ed. by N. A. Toropov and E. A. Porai-Koshits, tr. by E. B.
Uvarov, New York, Consultants Bureau. p. 178–200.
48. Forsbacka, L. (2007). Experimental study and modelling of viscosity of
chromium containing slags. Steel Research International, 78(9), 676–684.
49. Saito, N., Hori, N., Nakashima, K., & Mori, K. (2003). Viscosity of blast
furnace type slags. Metallurgical and Materials Transactions B, 34(5), 509–
516.
50. Lee, Y. S., Kim, J. R., Yi, S. H., & Min, D. J. (2004). Viscous behaviour of
CaO-SiO2 -Al2 O3 -MgO-FeO slag. In II International Conference on Molten
Slags Fluxes and Salts, The South African Institute of Mining and Metallurgy
(pp. 225–230).
51. Shankar, A., Görnerup, M., Lahiri, A. K., & Seetharaman, S. (2007). Exper-
imental investigation of the viscosities in CaO-SiO2 -MgO-Al2 O3 and CaO-
SiO2 -MgO-Al2 O3 -TiO2 slags. Metallurgical and Materials Transactions B,
38(6), 911–915.
52. Ro, K. J., Lee, Y. S., Min, D. J., Jung, S. M., & Yi, S. H. (2004). Influence of
MgO and Al2 O3 contents on viscosity of blast furnace type slags containing
FeO. ISIJ International, 44(8), 1291–1297.
53. Song, M., Shu, Q., & Sichen, D. (2011). Viscosities of the quaternary Al2 O3 -
CaO-MgO-SiO2 slags. Steel Research International, 82(3), 260–268.
54. Tang, X., Zhang, Z., Zhang, M., & Wang, X. (2011). Viscosities behavior of
CaO-SiO2 -MgO-Al2 O3 slag with low mass ratio of CaO to SiO2 and wide
range of Al2 O3 content. Journal of Iron and Steel Research , International,
18(2), 1–17.
55. Liao, J., Zhang, Y., Sridhar, S., Wang, X., Zhang, Z., Gao, Y., Kim, H. G.,
Sohn, H. Y., & Kim, C. W. (2012). Effect of Al2 O3 /SiO2 ration on the viscosity
and structure of slags. ISIJ International, 52(5), 753–758.
56. Chen, M., Zhang, D., Kou, M., & Zhao, B. (2014). Viscosities of iron blast
furnace slags. Transactions of the Iron & Steel Institute of Japan, 54(9),
2025–2030.
57. Gao, Y. M., Wang, S. B., Hong, C., Ma, X. J., & Yang, F. (2014). Effects
of basicity and MgO content on the viscosity of the SiO2 -CaO-MgO-
9wt%Al2 O3 slag system. International Journal of Minerals, Metallurgy and
Materials, 21(4), 353–362.
58. Yao, L., Ren, S., Wang, X., Liu, Q., Dong, L., Yang, J., & Liu, J. (2016). Effect
of Al2 O3 , MgO, and CaO/SiO2 on viscosity of high alumina blast furnace
slag. Steel Research International, 87(2), 241–249.
59. Sun, C. Y., Liu, X. H., Li, J., Yin, X. T., Song, S., & Wang, Q. (2017). Influence
of Al2 O3 and MgO on the viscosity and stability of CaO-MgO-SiO2 -Al2 O3
slags with CaO/SiO2 =1.0. ISIJ International, 57(6).
60. Azuma, K., Okamura, K., Nakabayashi, H., & Lee, S. Y. (1973). The effect
of Al2 O3 addition on the viscosity of the FeO-CaO-SiO2 slags. Journal of
the Mining and Metallurgical Institute of Japan. 89(1025), 467–472.
172 5 Estimation of High Alumina Blast Furnace Slag Properties

61. Hurst, H. J., Novak, F., & Patterson, J. H. (1999). Viscosity measurements
and empirical predictions for fluxed Australian bituminous coal ashes. Fuel,
78(15), 1831–1840.
62. Gimmelfarb, A. A. (1968). Viscosity of slags of the four-component CaO-
SiO2 -FeO-Al2 O3 system. Izvestia Akademii Nauk SSSR Metally, 2, 59–70.
63. Higgins, T. J. R. (1963). Viscosity characteristics of Rhodesian copper
smelting slags. Bulletin of Institution of Mining and Metallurgy, 72, 825–864.
64. Hua, H., Shin, J., & Kim, J. (2014). Level set, phase-field, and immersed
boundary methods for two-phase fluid flows. Journal of Fluids Engineering,
136(2).
65. Bills, P. M. (1963). Viscosities in silicate slag systems. Journal of the Iron
and Steel Institute, 133–140.
66. Lee, Y. S., Min, D. J., Jung, S. M., & Yi, S. H. (2004). Influence of basicity
and FeO content on viscosity of blast. ISIJ International. 44(8), 1283–1290.
67. Fincham, C., & Richardson, F. D. (1954). The behaviour of sulphur in silicate
and aluminate melts. Proceedings of the Royal Society of London. Series A.
Mathematical and Physical Sciences, 223(1152), 40–62.
68. Carter, P. T., & Macfarlane, T. G. (1957). Thermodynamics of slag systems:
Part II—The thermodynamic properties of CaO-SiO2 slags. The Journal of
the Iron and Steel Institute, 185, 62–66.
69. Abraham, K. P., & Richardson, F. D. (1960). Sulphide capacities of silicate
melts PART II. The Journal of the Iron and Steel Institute, 196, 313–317.
70. Abraham, K. P., & Richardson, F. D. (1960). Sulphide capacities of silicate
melts PART I. The Journal of the Iron and Steel Institute, 196, 309–312.
71. Brown, R. J. R. S. D., Ghita, I., & Bell, H. B. (1982). Sulphide capacity of
titania-containing slags. Ironmaking Steelmaking, 9, 163–167.
72. Görnerup, M., & Wijk, O. (1996). Sulphide capacities of CaO-Al2 O3 -SiO2
slags at 1550, 1600 and 1650 C. Scandinavian Journal of Metallurgy, 25(3),
103–107.
73. Park, G.-H., Kang, Y.-B., & Park, J. H. (2011). Sulfide capacity of the CaO–
SiO2 –MnO slag at 1873K. ISIJ International, 51(9), 1375–1382.
74. Carter, P. T., & Macfarlane, T. G. (1957). Thermodynamics of slag systems:
Part I—The thermodynamic properties of CaO-Al2 O3 slags. The Journal of
the Iron and Steel Institute, 185, 54–61.
75. Sharma, R., & Richardson, F. (1961). Activities in lime-alumina melts. The
Journal of the Iron and Steel Institute, 198, 386–390.
76. Cameron, J., Gibbons, T., & Taylor, J. (1966). Calcium sulphide solubilities
and lime activities in lime-alumina-silica system. Journal of the Iron and
Steel Institute, 204, 1223.
77. Kor, G., & Richardson, F. (1968). Sulphur in lime-alumina mixtures. Journal
of the Iron and Steel Institute, 206, 700.
78. Hino, M., Kitagawa, S., & Ban-Ya, S. (1993). Sulphide capacities of CaO-
Al2 O3 -MgO and CaO-Al2 O3 -SiO2 slags. ISIJ International, 33(1), 36–42.
Appendix 173

79. Drakaliysky, E., Nilsson, R., Sichen, D., & Seetharaman, S. (1996). Sulphide
capacities of CaO-Al2 O3 slags in the temperature range 1773-1848 K. High
Temperature Materials and Processes, 15(4), 263–272.
80. Ohta, M., Kubo, T., & Morita, K. (2003). Effects of CaF2 , MgO and SiO2
addition on sulfide capacities of the CaO-Al2 O3 slag. Tetsu-to-Hagane, 89(7),
742–749.
81. Choi, J. S., Park, Y., Lee, S., & Min, D. J. (2018). Cationic effect of charge
compensation on the sulfide capacity of aluminosilicate slags. Journal of the
American Ceramic Society, 101(7), 2856–2867.
82. Sharma, R. A., & Richardson, F. D. (1965). Activities of manganese oxide,
sulfide capacities, and activity coefficients in aluminate and silicate melts.
Transactions of the Metallurgical Society of AIME, 233, 1586–1592.
83. Nzotta, M. M., Nilsson, R., Sichen, D., & Seetharaman, S. (1997). Sulphide
capacities in MgO-SiO2 and CaO-MgO-SiO2 slags. Ironmaking and Steel-
making, 24(4), 300–305.
84. ProQuest LLC (1955).
85. Kalyanram, M., Macfarlane, T., & Bell, H. (1960). The activity of calcium
oxide in slags in the systems CaO–MgO–SiO2 , CaO–Al2 O3 –SiO2 , and CaO–
MgO–Al2 O3 –SiO2 at 1500° C. Journal of the Iron and Steel Institute London,
195, 58–64.
86. Brown, S., Roxburgh, R., Ghita, I., & Bell, H. (1982). Sulphide capacity of
titania-containing slags. Ironmaking and Steelmaking, 9, 163–167.
87. Karsrud, K. (1984). Sulfide capacities of synthetic blast furnace slags at 1500
deg C. Scandinavian Journal of Metallurgy, 13(3), 144–150.
88. Taniguchi, Y., Sano, N., & Seetharaman, S. (2009). Sulphide capacities of
CaO–Al2 O3 –SiO2 –MgO–MnO slags in the temperature range 1673–1773K.
ISIJ International, 49(2), 156–163.
89. Taniguchi, Y., Wang, L., Sano, N., & Seetharaman, S. (2012). Sulfide capac-
ities of CaO-Al2 O3 -SiO2 slags in the temperature range 1673 K to 1773 K
(1400 °C to 1500 °C). Metallurgical and Materials Transactions B, 43(3),
477–484.
90. Condo, A. F. T., Qifeng, S., & Sichen, D. (2018). Sulfide capacities in the
Al2 O3 -CaO-MgO-SiO2 system. Steel Research International, 89(8).
91. Kalyanram, M. R., Macfarlane, T. G., & Bell, H. B. (1960). The activity
of calcium oxide in slags in the systems CaO-MgO-SiO2 , CaO-AI2 O3 -SiO2
and CaO-MgO-AI2 O3 -SiO2 at 1500°C. Journal of the Iron and Steel Institute,
195, 58–63.
92. O’neill, H. S. C., & Mavrogenes, J. A. (2002). The sulfide capacity and the
sulfur content at sulfide saturation of silicate melts at 1400 C and 1 bar.
Journal of Petrology, 43(6), 1049–1087.
93. Hino, M., et al. (1993). Sulphide capacities of CaO-AI2 O3 -MgO and CaO-
AI2 O3 -SiO2 slags. ISIJ International, 33, 36–42.
94. Nzotta, M. M. (1997). Experimental determination of sulphide capacities in
the Al2 O3 -MgO-SiO2 , Al2 O3 -MnO-SiO2 and Al2 O3 -CaO-MgO slags in the
174 5 Estimation of High Alumina Blast Furnace Slag Properties

temperature range 1773-1923 K. Scandinavian Journal of Metallurgy, 26(4),


169–177.
95. Abraham, K. P., et al. (1960). Sulphide capacities of silicate melts, Part 2.
The Journal of the Iron and Steel Institute, 196, 4.
96. Drakaliysky, E., Srinivasan, N., & Staffansson, L.-I. (1991). Sulfide capac-
ities of CaO-MgO-SiO2 -Al2 O3 slags containing small amounts of Cr2 O3 .
Scandinavian Journal of Metallurgy, 20(4), 251–255.
97. Tang, X., & Xu, C. (1995). Sulphur distribution between CaO-SiO2 -TiO2 -
Al2 O3 -MgO slag and carbon-saturated iron at 1773K. ISIJ International,
35(4), 367–371.
98. Nzotta, M. M., Sichen, D., & Seetharaman, S. (1998). Sulphide capacities in
some multi component slag systems. ISIJ International, 38(11), 1170–1179.
99. Seo, J.-D., & Kim, S.-H. (1999). The sulphide capacity of CaO-SiO2 -Al2 O3 -
MgO(-FeO) smelting reduction slags. Steel Research, 70(6), 203–208.
100. Shankar, A., Görnerup, M., Seetharaman, S., & Lahiri, A. K. (2006). Sulfide
capacity of high alumina blast furnace slags. Metallurgical & Materials
Transactions B, 37(6), 941–947.
101. Hayakawa, H., Hasegawa, M., Oh-nuki, K., Sawai, T., & Iwase, M. (2006).
Sulphide capacities of CaO-SiO2 -Al2 O3 -MgO slags. Steel Research Interna-
tional, 77(1), 14–20.
102. Yang, X., Jiao, J., Ding, R., Shi, C., & Guo, H. (2009). A thermodynamic
model for calculating sulphur distribution ratio between CaO-SiO2 -MgO-
Al2 O3 ironmaking slags and carbon saturated hot metal based on the ion and
molecule coexistence theory. ISIJ International, 49(12), 1828–1837.
103. Park, J. H., & Park, G.-H. (2012). Sulfide capacity of CaO–SiO2 –MnO–
Al2 O3 –MgO slags at 1873 K. ISIJ International, 52(5), 764–769.
104. Allertz, C., & Sichen, D. (2015). Sulfide capacity in ladle slag at steel-
making temperatures. Metallurgical and Materials Transactions B, 46(6),
2609–2615.
105. Wang, L., Wang, Y., Chou, K., & Seetharaman, S. (2016). Sulfide capac-
ities of CaO-MgO-Al2 O3 -SiO2 -CrOx slags. Metallurgical and Materials
Transactions B, 47(4), 2558–2563.
106. Ma, X., Chen, M., Xu, H., Zhu, J., Wang, G., & Zhao, B. (2016). Sulphide
capacity of CaO–SiO2 –Al2 O3 –MgO system relevant to low MgO blast
furnace slags. ISIJ International, 56(12), 2126–2131.
107. Talapaneni, T., Yedla, N., & Sarkar, S. (2018). Study on desulfurization
capacity of high alumina blast furnace slag at 1773 K using slag-metal
equilibrium technique. Metallurgical Research & Technology, 115(5).
108. Condo, A. F. T., Allertz, C., & Sichen, D. (2019). Experimental determina-
tion of sulphide capacities of blast furnace slags with higher MgO contents.
Ironmaking & Steelmaking, 46(3), 207–210.
References 175

References

1. www.factsage.com (2020).
2. https://thermocalc.com/ (2021).
3. https://www.phase-trans.msm.cam.ac.uk/mtdata/index.html (2021).
4. Urbain, G. (1987). Viscosity esitimation of slags. Steel Research, 58(3), 111–116.
5. Riboud, P., Roux, Y., Lucas, L., & Gaye, H. (1981). Improvement of continuous casting
powders. Fachberichte Huttenpraxis Metallweiterverarbeitung, 19(8), 859–869.
6. Hebbar, K., & Reddy, R. G. (1991). Prediction of viscosities of binary silicate melts. In EPD
Congress, TMS1991 (pp. 523–540).
7. Zhang, Z., & Reddy, R. G. (2004). Experimental determination and modeling of Na2 O-SiO2 -
B2 O3 melts viscosities. High Temperature Materials and Processes, 23(4), 247–259.
8. Yokokawa, T., & Niwa, K. (1969). Free energy of solution in binary silicate melts. Transactions
of the Japan Institute of Metals, 10(1), 3–7.
9. Du, S., Bygd’en, J., & Seetharaman, S. (1994). A model for estimation of viscosities of
complex metallic and ionic melts. Metallurgical and Materials Transactions B, 25(4), 519–
525.
10. Wu, T., He, S., Liang, Y., & Wang, Q. (2015). Molecular dynamics simulation of the structure
and properties for the CaO–SiO2 and CaO–Al2 O3 systems. Journal of Non-crystalline Solids,
411, 145–151.
11. Zhang, L., Sun, S., & Jahanshahi, S. (2001). Molecular dynamics simulations of silicate slags
and slag–solid interfaces. Journal of Non-crystalline Solids, 282(1), 24–29.
12. Šob, M., Friák, M., Legut, D., Fiala, J., & Vitek, V. (2004). The role of ab initio electronic struc-
ture calculations in studies of the strength of materials. Materials Science and Engineering:
A, 387, 148–157.
13. Sajid, M., Bai, C., Yu, W., You, Z., Tan, M., & Aamir, M. (2020). First principle study of
electronic structural and physical properties of CaOSiO2 Al2 O3 ternary slag system by using
Ab Initio molecular and lattice dynamics. Journal of Non-crystalline Solids, 550, 120384.
14. Xin, Z.-C., Zhang, J.-S., Lin, W.-H., Zhang, J.-G., Jin, Y., Zheng, J., Cui, J.-F., & Liu, Q. (2021).
Sulphide capacity prediction of CaO–SiO2 –MgO–Al2 O3 slag system by using regularized
extreme learning machine. Ironmaking & Steelmaking, 48(3), 275–283.
15. Nakamoto, M., Hanao, M., Tanaka, T., Kawamoto, M., Holappa, L., & Hämäläinen, M.
(2007). Estimation of surface tension of molten silicates using neural network computation.
ISIJ International, 47(8), 1075–1081.
16. Derin, B., Suzuki, M., & Tanaka, T. (2010). Sulphide capacity prediction of molten slags by
using a neural network approach. ISIJ International, 50(8), 1059–1063.
17. Ma, A., Mostaghel, S., & Chattopadhyay, K. (2017). Development of an artificial neural
network to predict sulphide capacities of CaO–SiO2 –Al2 O3 –MgO slag system. ISIJ Interna-
tional, 57(1), 114–122.
18. Hanao, M., Kawamoto, M., Tanaka, T., & Nakamoto, M. (2006). Evaluation of viscosity of
mold flux by using neural network computation. ISIJ International, 46(3), 346–351.
19. Sridhar, S., Mills, K. C., Afrange, O. D. C., Lörz, H. P., & Carli, R. (2000). Break temperatures
of mould fluxes and their relevance to continuous casting. Ironmaking & Steelmaking, 27(3),
238–242.
20. Arrhenius, S. (1887). Über die innere Reibung verdünnter wässeriger Lösungen. Zeitschrift
für Physikalische Chemie, 1(1), 285–298.
21. Vogel, H. (1921). Das Temperaturabhängigkeitsgesetz der Viskosität von Flüssigkeiten.
Physikalische Zeitschrift, 22, 645–646.
22. Fulcher, G. S. (1925). Analysis of recent measurements of the viscosity of glasses. Journal
of the American Ceramic Society, 8(6), 339–355.
23. Tammann, G., & Hesse, W. (1926). Die Abhängigkeit der Viscosität von der Temperatur bie
unterkühlten Flüssigkeiten. Zeitschrift Für Anorganische Chemie, 156(1), 245–257.
24. Richet, P. (1984). Viscosity and configurational entropy of silicate melts. Geochimica et
Cosmochimica Acta, 48(3), 471–483.
176 5 Estimation of High Alumina Blast Furnace Slag Properties

25. Eyring, H., & Urry, D. W. (1938). The theory of absolute reaction rates. Transactions of the
Faraday Society, 34, 41–48.
26. Weymann, H. D. (1962). On the hole theory of viscosity, compressibility, and expansivity of
liquids. Kolloid-Zeitschrift und Zeitschrift für Polymere, 181(2), 131–137.
27. Milchev, A., & Avramov, I. (1983). On the influence of amorphization on atomic diffusion in
condensed systems. Physica Status Solidi A, 120(1), 123–130.
28. Avramov, I., Rüssel, C., & Keding, R. (2003). Effect of chemical composition on viscosity of
oxide glasses. Journal of Non-crystalline Solids, 324(1–2), 29–35.
29. Porter, D. A., Easterling, K. E., & Sherif, M. (2009). Phase transformations in metals and
alloys (Revised Reprint). CRC Press.
30. Urbain, G. (1985). Viscosity of silicate melts: Measure and estimation. Transactions & Journal
of the British Ceramic Society, 80(9), 139–141.
31. Mauro, J. C., Yue, Y., Ellison, A. J., Gupta, P. K., & Allan, D. C. (2009). Viscosity of glass-
forming liquids. Proceedings of the National Academy of Sciences of the United States of
America, 106(47), 19780–19784.
32. Richet, P., Robie, R. A., & Hemingway, B. S. (1986). Low-temperature heat capacity of
diopside glass (CaMgSi2 O6 ): A calorimetric test of the configurational-entropy theory applied
to the viscosity of liquid silicates. Geochimica et Cosmochimica Acta, 50(7), 1521–1533.
33. Yan, B., Richet, P., & Sipp, A. (1995). Viscosity regimes of homogeneous silicate melts.
American Mineralogist, 80(3–4), 305–318.
34. Fürth, R. (1941). On the theory of the liquid state: I. The statistical treatment of the ther-
modynamics of liquids by the theory of holes. Mathematical Proceedings of the Cambridge
Philosophical Society, 37(3), 252–275.
35. Fürth, R. (1941). In Mathematical Proceedings of the Cambridge Philosophical Society
(pp. 281–290). Cambridge University Press.
36. Mills, K. C., Franken, M., Machingawuta, N., Green, P., Broardbent, C., Urbain, G., Scheel,
R., Hocevoir, J., & Pontoire, J. (1991). Commission report on Standard Reference Material
(SRM) for high temperature viscosity measurements. National Physical Laboratory.
37. Lakatos, C. (1972). Viscosity temperature relations in the glass system SiO2 -Al2 O3 -Na2 O-
K2 O-CaO-MgO in the composition range of technical glasses.
38. Kim, J. W., Choi, J., Kwon, O. D., Lee, I. R., Shin, Y. K., & Park, J. S. (1992). In Forth
International Conference on Molten Slags and Fluxes. ISIJ.
39. Urbain, G., Cambier, F., Deletter, M., & Anseau, M. R. (1981). Viscosity of silicate melts.
Transactions & Journal of the British Ceramic Society, 80(9), 139–141.
40. Urbain, G., Bottinga, Y., & Richet, P. (1982). Viscosity of liquid silica, silicates and alumino-
silicates. Geochimica et Cosmochimica Acta, 46(6), 1061–1072.
41. Kondratiev, A., & Jak, E. (2001). Review of experimental data and modeling of the viscosi-
ties of fully liquid slags in the Al2 O3 -CaO-‘FeO’-SiO2 system. Metallurgical & Materials
Transactions B, 32(6), 1015–1025.
42. Gan, L., & Lai, C. (2014). A general viscosity model for molten blast furnace slag.
Metallurgical & Materials Transactions B, 45(3), 875–888.
43. Iida, T., Sakai, H., Kita, Y., & Murakami, K. (2000). Equation for estimating viscosities of
industrial mold fluxes. High Temperature Materials and Processes, 153.
44. Iida, T., Sakai, H., Kita, Y., & Shigeno, K. (2000). An equation for accurate prediction of
the viscosities of blast furnace type slags from chemical composition. ISIJ International, 40,
S110–S114.
45. Mills, K., & Sridhar, S. (1999). Viscosities of ironmaking and steelmaking slags. Ironmaking &
Steelmaking, 26(4), 262–268.
46. Zhang, Z., & Reddy, R. G. (2005). Structure model and properties of alkali borate melts.
Mineral Processing and Extractive Metallurgy Review, 114(3), 192–199.
47. Zhang, L., & Jahanshahi, S. (1998). Review and modeling of viscosity of silicate melts: Part
I. Viscosity of binary and ternary silicates containing CaO, MgO, and MnO. Metallurgical
and Materials Transactions B, 29(1), 177–186.
References 177

48. Zhang, L., & Jahanshahi, S. (1998). Review and modeling of viscosity of silicate melts: Part
II. Viscosity of melts containing iron oxide in the CaO-MgO-MnO-FeO-Fe2 O3 -SiO2 system.
Metallurgical and Materials Transactions B, 29(1), 187–195.
49. Nakamoto, M., Lee, J., & Tanaka, T. (2005). A model for estimation of viscosity of molten
silicate slag. ISIJ international, 45(5), 651–656.
50. Suzuki, M., & Jak, E. (2013). Quasi-chemical viscosity model for fully liquid slag in the
Al2 O3 -CaO-MgO-SiO2 system—Part I: Revision of the model. Metallurgical and Materials
Transactions B, 44(6), 1435–1450.
51. Suzuki, M., & Jak, E. (2013). Quasi-chemical viscosity model for fully liquid slag in the
Al2 O3 -CaO-MgO-SiO2 system. Part II: Evaluation of slag viscosities. Metallurgical and
Materials Transactions B, 44(6), 1451–1465.
52. Zhang, G.-H., Chou, K.-C., & Mills, K. (2014). A structurally based viscosity model for oxide
melts. Metallurgical and Materials Transactions B, 45(2), 698–706.
53. Senior, C., & Srinivasachar, S. (1995). Viscosity of ash particles in combustion systems for
prediction of particle sticking. Energy & Fuels, 9(2), 277–283.
54. Grundy, A. N., Liu, H., Jung, I.-H., Decterov, S. A., & Pelton, A. D. (2008). A model to
calculate the viscosity of silicate melts Part I: Viscosity of binary SiO2 –MeOx systems (Me
= Na, K, Ca, Mg, Al). International Journal of Materials Research, 99, 1185–1194.
55. Grundy, A. N., Jung, I.-H., Pelton, A. D., & Decterov, S. A. (2008). A model to calculate the
viscosity of silicate melts Part II: The NaO0. 5–MgO–CaO–AlO1.5 –SiO2 system. International
Journal of Materials Research, 99, 1195–1209.
56. Einstein, A. (1906). A new determination of molecular dimensions. Annalen der Physik, 19,
289–306.
57. Roscoe, R. (1952). The viscosity of suspensions of rigid spheres. British Journal of Applied
Physics, 3(8), 267.
58. Kondratiev, A., & Jak, E. (2001). Modeling of viscosities of the partly crystallized slags
in the Al2 O3 -CaO-“FeO”-SiO2 system. Metallurgical & Materials Transactions B, 32(6),
1027–1032.
59. Wright, S., Zhang, L., Sun, S., & Jahanshahi, S. (2001). Viscosities of calcium ferrite slags and
calcium alumino-silicate slags containing spinel particles. Journal of Non-crystalline Solids,
282(1), 15–23.
60. Chong, J., Christiansen, E., & Baer, A. (1971). Rheology of concentrated suspensions. Journal
of Applied Polymer Science, 15(8), 2007–2021.
61. Hirai, M., Takebayashi, K., Yoshikawa, Y., & Yamaguchi, R. (1992). Apparent viscosity of
semi-solid metals. Tetsu-to-Hagane, 78(6), 902–909.
62. Mendoza, C. I., & Santamaria-Holek, I. (2009). The rheology of hard sphere suspensions at
arbitrary volume fractions: An improved differential viscosity model. The Journal of Chemical
Physics, 130(4), 044904.
63. Hou, L.-M., Liu, L.-L., Zhang, Z.-T., & Wang, X.-D. (2012). Experimental research and
prediction model on viscosity of high FeO content slag. Iron and Steel/Gangtie, 47(7), 20–25.
64. Duchesne, M. A., Macchi, A., Lu, D. Y., Hughes, R. W., McCalden, D., & Anthony, E.
J. (2010). Artificial neural network model to predict slag viscosity over a broad range of
temperatures and slag compositions. Fuel Processing Technology, 91(8), 831–836.
65. Cheng, R., Ni, H., Li, X., Zhu, W., Xiong, J., He, H., & Zhang, H. (2012). Predicting viscosity
of blast furnace slag based on BP neural network model: An experimental study. Journal of
Wuhan University of Science and Technology, 06.
66. Chen, Z., Wang, M., Wang, H., Liu, L., & Wang, X. (2022). ANN-based structure-
viscosity relationship model of multicomponent slags for production design in mineral wool.
Construction and Building Materials, 319.
67. Chen, Z., Wang, M., Meng, Z., Wang, H., Liu, L., & Wang, X. (2021). Develop-
ment of structure-informed artificial neural network for accurately modeling viscosity of
multicomponent molten slags. Ceramics International, 47(21), 30691–30701.
68. Reddy, A. K., & Bockris, J. O. M. (1977). Modern electrochemistry (6th Printing, pp. 513–
622). Plenum Press.
178 5 Estimation of High Alumina Blast Furnace Slag Properties

69. Yan, Z., Reddy, R. G., & Lv, X. (2019). Structure based viscosity model for aluminosilicate
slag. ISIJ International, 59(6), 1018–1026.
70. Bockris, J. O. M., Mackenzie, J. D., & Kitchener, J. A. (1955). Viscous flow in silica and
binary liquid silicates. Transactions of the Faraday Society, 51, 1734–1748.
71. Bruckner, R. (1964). Characteristic physical properties of the principal glass-forming oxides
and their relation to glass structure. Glass Science and Technology, 37, 413–425.
72. Kaiura, G. H., Toguri, J. M., & Marchant, G. (1977). Viscosity of fayalite-based slags.
Canadian Metallurgical Quarterly, 16(1), 156–160.
73. Kucharski, M., Stubina, N. M., & Toguri, J. M. (1989). Viscosity measurements of molten Fe–
O–SiO2 , Fe–O–CaO–SiO2 , and Fe–O–MgO–SiO2 slags. Canadian Metallurgical Quarterly,
28(1), 7–11.
74. Shiraishi, Y., Ikeda, K., Tamura, A., & Saito, T. (1978). On the viscosity and density of the
molten FeO-SiO2 system. Transactions of the Japan Institute of Metals, 19(5), 264–274.
75. Waseda, Y., & Shiraishi, Y. (1980). The structure of the molten FeO–Fe2 O3 –SiO2 system by
X-ray diffraction. Transactions of the Japan Institute of Metals, 21(1), 51–62.
76. Chen, M., Raghunath, S., & Zhao, B. (2013). Viscosity measurements of “FeO”-SiO2 slag in
equilibrium with metallic Fe. Metallurgical & Materials Transactions B, 44(3), 506–515.
77. Yan, Z., Reddy, R. G., Lv, X., Pang, Z., & Bai, C. (2019). Viscosity of iron oxide
aluminosilicate melts. Metallurgical and Materials Transactions B, 50(1), 251–261.
78. Licko, T., & Danek, V. (1986). Viscosity and structure of melts in the system CaO-MgO-SiO2 .
Physics and Chemistry of Glasses, 27.
79. Chen, M., & Zhao, B. (2015). Viscosity measurements of SiO2 -“FeO”-CaO system in
equilibrium with metallic Fe. Metallurgical and Materials Transactions B, 46(2), 577–584.
80. Ji, F. Z., Du, S., & Seetharaman, S. (1997). Experimental studies of the viscosities in the
CaO-FenO-SiO2 slags. Metallurgical & Materials Transactions B, 28(5), 827–834.
81. Williams, M. S. P., & Briggs, G. (1983). Viscosities of synthetic slags in the system CaO-
“FeO”-SiO2 -MgO. Institution of Mining and Metallurgy, Section C, 105–109.
82. Johannsen, F., & Wiese, W. (1958). Das Absetzverhalten on Kupferstein und Kupfer in
fliissigen Schlacken. Zeitschrift fiir Erzbergbau und Metallhiittenwesen, 1(48), 1–15.
83. Chen, M., Raghunath, S., & Zhao, B. (2014). Viscosity measurements of SiO2 -“FeO”-MgO
system in equilibrium with metallic Fe. Metallurgical & Materials Transactions B, 45(1),
58–65.
84. Toropov, N. A., & Bryantsev, B. A. (1965). Physiochemical properties and crystallization
of melts in the system magnesium oxide-ferrous oxide-silica. Structural Transformations in
Glasses at High Temperatures, 178–200.
85. Ji, F. Z., Sichen, D., & Seetharaman, S. (1998). Experimental studies of the viscosities in the
FeO MgO SiO2 and FeO MnO SiO2 slags. Ironmaking & Steelmaking, 25(4), 309–316.
86. Kozakevitch, P. (1960). Viscosité et éléments structuraux des aluminosilicates fondus: Laitiers
CaO-AI2 O3 -SiO2 entre 1600 et 2100 °C. Revue de métallurgie, 57, 149–160.
87. Řeháčková, L., Rosypalová, S., Dudek, R., Ritz, M., Matýsek, D., Smetana, B., Dobrovská,
J., Zlá, S., & Kawuloková, M. (2015). Effect of chemical composition and temperature
on viscosity and structure of molten CaO-Al2 O3 -SiO2 system. Archives of Metallurgy and
Materials, 60(4), 2873–2878.
88. Cukierman, M., & Uhlmann, D. R. (1973). Viscosity of liquid anorthite. Journal of
Geophysical Research, 78(23), 4920–4923.
89. Taniguchi, H. (1992). Entropy dependence of viscosity and the glass-transition temperature
of melts in the system diopside-anorthite. Contributions to Mineralogy & Petrology, 109(3),
295–303.
90. Azuma, K., Okamura, K., Nakabayashi, H., & Lee, S. Y. (1973). The effect of Al2 O3 addition
on the viscosity of the FeO-CaO-SiO2 slags. 89(1025), 467–472.
91. Hurst, H. J., Novak, F., & Patterson, J. H. (2000). Viscosity measurements and empirical
predictions for some model gasifier slags. Fuel, 79(14), 1797–1799.
92. Hurst, H. J., Novak, F., & Patterson, J. H. (1999). Viscosity measurements and empirical
predictions for fluxed Australian bituminous coal ashes. Fuel, 78(15), 1831–1840.
References 179

93. Higgins, T. J. R. (1963). Viscosity characteristics of Rhodesian copper smelting slags. Bulletin
of Institution of Mining and Metallurgy, 72, 825–864.
94. Gimmelfarb, A. A. (1968). Viscosity of slags of the four-component CaO-SiO2 -FeO-Al2 O3
system. Izvestia Akademii Nauk SSSR Metally, 2, 59–70.
95. Kim, J. R., Lee, Y. S., Min, D. J., Jung, S. M., & Yi, S. H. (2004). Influence of MgO and
Al2 O3 contents on viscosity of blast furnace type slags containing FeO. ISIJ International,
44(8), 1291–1297.
96. Lee, Y. S., Kim, J. R., Yi, S. H., & Min, D. J. (2004). Viscous behaviour of CaO-SiO2 -
Al2 O3 -MgO-FeO slag. In II International Conference on Molten Slags Fluxes and Salts
(pp. 225–230). The South African Institute of Mining and Metallurgy.
97. Bills, P. M. (1963). Viscosities in silicate slag systems. Journal of the Iron and Steel Institute,
133–140.
98. Bottinga, Y., & Weill, D. F. (1970). Densities of liquid silicate systems calculated from partial
molar volumes of oxide components. American Journal of Science, 269(2), 169–182.
99. Bottinga, Y., Weill, D., & Richet, P. (1982). Density calculations for silicate liquids. I. Revised
method for aluminosilicate compositions. Geochimica et Cosmochimica Acta, 46(6), 909–
919.
100. Mills, K. C., & Keene, B. J. (1987). Physical properties of BOS slags. International Materials
Reviews, 32(1), 1–120.
101. Mills, K. C., Karagadde, S., Lee, P. D., Yuan, L., & Shahbazian, F. (2016). Calculation of
physical properties for use in models of continuous casting process-part 1: Mould slags. ISIJ
International, 56(2), 264–273.
102. Lange, R. A., & Carmichael, I. S. (1987). Densities of Na2 O-K2 O-CaO-MgO-FeO-
Fe2 O3 -Al2 O3 -TiO2 -SiO2 liquids: New measurements and derived partial molar properties.
Geochimica et Cosmochimica Acta, 51(11), 2931–2946.
103. Nakajima, K. (1994). Estimation of molar volume for multicomponent silicate melts. Tetsu-
to-hagané, 80(8), 593–598.
104. Courtial, P., & Dingwell, D. B. (1995). Nonlinear composition dependence of molar volume of
melts in the CaOAl2 O33 SiO2 system. Geochimica et Cosmochimica Acta, 59(18), 3685–3695.
105. Courtial, P., & Dingwell, D. B. (1999). Densities of melts in the CaO-MgO-Al2 O3 -SiO2
system. American Mineralogist, 84(4), 465–476.
106. Linard, Y., Nonnet, H., & Advocat, T. (2008). Physicochemical model for predicting molten
glass density. Journal of Non-crystalline Solids, 354(45), 4917–4926.
107. Xin, J., Gan, L., Wang, N., & Chen, M. (2019). Accurate density calculation for molten slags
in SiO2 -Al2 O3 -CaO-MgO-‘FeO’-‘Fe2 O3 ’systems. Metallurgical and Materials Transactions
B, 50(6), 2828–2842.
108. Xin, J., Gan, L., Jiao, L., & Lai, C. (2017). Accurate density calculation for molten slags in
SiO2 –Al2 O3 –CaO–MgO systems. ISIJ International, ISIJINT-2017-070.
109. Hayashi, M., & Seetharaman, S. (2003). In Current advances in materials and processes.
Report of the ISIJ Meeting (pp. 860–863).
110. Persson, M., Matsushita, T., Zhang, J., & Seetharaman, S. (2007). Estimation of molar volumes
of some binary slags from enthalpies of mixing. Steel Research International, 78(2), 102–108.
111. Persson, M., Zhang, J., & Seetharaman, S. (2007). A thermodynamic approach to a density
model for oxide melts. Steel Research International, 78(4), 290–298.
112. Shu, Q. (2007). A density estimation model for molten silicate slags. High Temperature
Materials and Processes, 26(5–6), 341–348.
113. Shu, Q. F., & Chou, K. C. (2013). Calculation for density of molten slags using optical basicity.
Ironmaking & Steelmaking, 40(8), 571–577.
114. Zhang, G.-H., & Chou, K.-C. (2010). Model for evaluating density of molten slag with optical
basicity. Journal of Iron and Steel Research, International, 17(4), 1–4.
115. Thibodeau, E., Gheribi, A. E., & Jung, I.-H. (2016). A structural molar volume model
for oxide melts part I: Li2 O-Na2 OK2 O-MgO-CaO-MnO-PbO-Al2 O3 -SiO2 melts—Binary
systems. Metallurgical and Materials Transactions B, 47(2), 1147–1164.
180 5 Estimation of High Alumina Blast Furnace Slag Properties

116. Thibodeau, E., Gheribi, A. E., & Jung, I.-H. (2016). A structural molar volume model for
oxide melts Part II: Li2 O-Na2 OK2 O-MgO-CaO-MnO-PbO-Al2 O3 -SiO2 melts—Ternary and
multicomponent systems. Metallurgical and Materials Transactions B, 47(2), 1165–1186.
117. Wang, L.-J., Chen, S., Chou, K.-C., & Chang, Y. A. (2005). Calculation of density in a ternary
system with a limited homogenous region using a geometric model. Calphad, 29(2), 149–154.
118. Chou, K.-C., Zhong, X., & Xu, K. (2004). Calculation of physicochemical properties in
a ternary system with miscibility gap. Metallurgical and Materials Transactions B, 35(4),
715–720.
119. Mills, K. (1995). Slag Atlas.
120. Knoche, R., Dingwell, D., & Webb, S. (1995). Melt densities for leucogranites and granitic
pegmatites: Partial molar volumes for SiO2 , Al2 O3 , Na2 O, K2 O, Li2 O, Rb2 O, Cs2 O, MgO,
CaO, SrO, BaO, B2 O3 , P2 O5 , F2 O–1, TiO2 , Nb2 O5 , Ta2 O5 , and WO3 . Geochimica et
Cosmochimica Acta, 59(22), 4645–4652.
121. Kucuk, A., Clare, A., & Jones, L. (1999). An estimation of the surface tension for silicate
glass melts at 1400 C using statistical analysis. Glass Technology, 40(5), 149–153.
122. Zhang, J., Shu, Q., & Wei, S. (2000). In Proceedings of the 6th International Conference on
Molten Slags, Fluxes and Salts, Stockholm.
123. Hanao, M., Kawamoto, M., Nakamoto, M., & Tanaka, T. (2009). In CDROM, Molten 2009,
Proceedings of the VIII International Conference on Molten Slags, Fluxes and Salts, Gecamin,
Santiago, Chile.
124. Mills, K. (2011). The estimation of slag properties. Southern African Pyrometallurgy, 52.
125. Nakamoto, M., Tanaka, T., Holappa, L., & Hamalainen, M. (2007). Surface tension evalua-
tion of molten silicates containing surface-active components (B2 O3 , CaF3 or Na2 O). ISIJ
International, 47(2), 211–216.
126. Nakamoto, M., Kiyose, A., Tanaka, T., Holappa, L., & Hamalainen, M. (2007). Evaluation
of the surface tension of ternary silicate melts containing Al2 O3 , CaO, FeO, MgO or MnO.
ISIJ International, 47(1), 38–43.
127. Tanaka, T., Hara, S., Ogawa, M., & Ueda, T. (2006). Evaluation of surface tension of molten
salt mixtures. ISIJ International, 46(3), 400–406.
128. Tanaka, T. (2003). Thermodynamics of surface: Relationship between surface properties and
thermodynamic properties in bulk. Iron & Steel Institute of Japan, 8, 223–229.
129. Tanaka, T., Hack, K., Iida, T., & Hara, S. (1996). Application of thermodynamic databases
to the evaluation of surface tensions of molten alloys, salt mixtures and oxide mixtures.
Zeitschrift fuer Metallkunde, 87(5), 380–389.
130. Hanao, M., Tanaka, T., Kawamoto, M., & Takatani, K. (2007). Evaluation of surface tension
of molten slag in multicomponent systems. ISIJ International, 47(7), 935–939.
131. Liu, Y., Lv, X., Bai, C., & Yu, B. (2014). Surface tension of the molten blast furnace slag
bearing TiO2 : Measurement and evaluation. Transactions of the Iron & Steel Institute of Japan,
54(10), 2154–2161.
132. Zhang, R., Wang, Z., Meng, Y., Jiao, S., Jia, J., Min, Y., & Liu, C. (2021). Structural anal-
ysis and evaluation of surface tension of silicate melts containing CaO and FexO. Chemical
Engineering Science, 245, 116870.
133. Fincham, C., & Richardson, F. D. (1954). The behaviour of sulphur in silicate and aluminate
melts. Proceedings of the Royal Society of London. Series A. Mathematical and Physical
Sciences, 223(1152), 40–62.
134. Duffy, J. A., Ingram, M. D., & Sommerville, I. D. (1978). Acid–base properties of molten
oxides and metallurgical slags. Journal of the Chemical Society. Faraday Transactions, 74,
1410–1419.
135. Tsao, T., & Katayama, H. G. (1986). Sulphur distribution between liquid iron and CaO-MgO-
Al2 O3 -SiO2 slags used for ladle refining. Transactions of the Iron and Steel Institute of Japan,
26(8), 717–723.
136. Sosinsky, N. J., & Sommerville, I. D. (1986). The composition and temperature dependence
of the sulfide capacity of metallurgical slags. Metallurgical Transactions B, 17(2), 331–337.
References 181

137. Young, R. W., Duffy, J. A., Hassall, G. J., & Xu, Z. (1992). Use of optical basicity concept for
determining phosphorus and sulfur slag metal partitions. Ironmaking & Steelmaking, 19(3),
201–219.
138. Shankar, A., Gornerup, M., Lahiri, A. K., & Seetharaman, S. (2006). Sulfide capacity of high
alumina blast furnace slags. Metallurgical and Materials Transactions B: Process Metallurgy
and Materials Processing Science, 37(6), 941–947.
139. Taniguchi, Y., Sano, N., & Seetharaman, S. (2009). Sulphide capacities of CaO–Al2 O3 –
SiO2 –MgO–MnO slags in the temperature range 1673–1773K. ISIJ International, 49(2),
156–163.
140. Zhang, G.-H., Chou, K.-C., & Pal, U. (2013). Estimation of sulfide capacities of multicom-
ponent slags using optical basicity. ISIJ International, 53(5), 761–767.
141. Ren, Z.-S., Hu, X.-J., & Chou, K.-C. (2013). Calculation and analysis of sulfide capacities for
CaO-Al2 O3 -SiO2 -MgO-TiO2 slags. Journal of Iron and Steel Research International, 20(9),
21–25.
142. Hao, X., & Wang, X. (2016). A new sulfide capacity model for CaO-Al2 O3 -SiO2 -MgO slags
based on corrected optical basicity. Steel Research International, 87(3), 359–363.
143. Reddy, R. G., & Blander, M. (1987). Modeling of sulfide capacities of silicate melts.
Metallurgical Transactions B, 18(3), 591–596.
144. Nzotta, M., Sichen, D., & Seetharaman, S. (1999). A study of the sulfide capacities of iron-
oxide containing slags. Metallurgical and Materials Transactions B, 30(5), 909–920.
145. Nzotta, M. M., Sichen, D. U., & Seetharaman, S. (1999). Sulphide capacities of “FeO”-SiO2 ,
CaO-“FeO”, and “FeO”-MnO slags. ISIJ International, 39(7), 657–663.
146. Nzotta, M. M., Sichen, D. U., & Seetharaman, S. (1998). Sulphide capacities in some multi
component slag systems. ISIJ International, 38(11), 1170–1179.
147. Nzotta, M. M., Nilsson, R., Sichen, D., & Seetharaman, S. (1997). Sulphide capacities in
MgO-SiO2 and CaO-MgO-SiO2 slags. Ironmaking and Steelmaking, 24(4), 300–305.
148. Kang, Y.-B., & Pelton, A. D. (2009). Thermodynamic model and database for sulfides
dissolved in molten oxide slags. Metallurgical and Materials Transactions B, 40(6), 979–994.
149. Yang, X.-M., Zhang, M., Shi, C.-B., Chai, G.-M., & Zhang, J. (2012). A sulfide capacity
prediction model of CaO-SiO2 -MgO-Al2 O3 slags during the LF refining process based on
the ion and molecule coexistence theory. Metallurgical and Materials Transactions B, 43(2),
241–266.
150. Yang, X., Shi, C., Zhang, M., & Zhang, J. (2012). A thermodynamic model for prediction of
iron oxide activity in some FeO-containing slag systems. Steel Research International, 83(3),
244–258.
151. Yang, X.-M., Shi, C.-B., Zhang, M., Chai, G.-M., & Wang, F. (2011). A thermodynamic
model of sulfur distribution ratio between CaO–SiO2 –MgO–FeO–MnO–Al2 O3 slags and
molten steel during LF refining process based on the ion and molecule coexistence theory.
Metallurgical and Materials Transactions B, 42(6), 1150–1180.
152. Zhang, J., Lv, X., Yan, Z., Qin, Y., & Bai, C. (2016). Desulphurisation ability of blast furnace
slag containing high Al2 O3 and 5 mass% TiO2 at 1773 K. Ironmaking & Steelmaking, 43(5),
378–384.
153. Jiao, Q., & Themelis, N. J. (1988). Correlations of electrical conductivity to slag composition
and temperature. Metallurgical Transactions B, 19(1), 133–140.
154. Wang, L. (2009). Estimations of electrical conductivities in molten slag systems. Steel
Research International, 80(9), 680–685.
155. Zhang, G. H., & Chou, K. C. (2010). Simple method for estimating the electrical conductivity
of oxide melts with optical basicity. Metallurgical & Materials Transactions B, 41(1), 131–
136.
156. Zhang, G. H., & Chou, K. C. (2012). Correlation between viscosity and electrical conductivity
of aluminosilicate melts. Metallurgical & Materials Transactions B, 43(4), 849–855.
157. Mills, K. C., Yuan, L., Li, Z., & Zhang, G. (2013). Estimating viscosities, electrical & thermal
conductivities of slags. High Temperatures—High Pressures, 42(3).
Chapter 6
The Revolution of High Alumina Slag
in Blast Furnace Process

6.1 The Inevitability of High Alumina Blast Furnace Slag

6.1.1 Iron and Steel Industry in the World

The iron and steel industry is an important industry in any country, the big country and
each developed country all have the strong steel industry because it plays an essential
role during the modernization process. Infrastructure, buildings, and autos require
huge amounts of steel in the developing process of a country, the most outstanding
example is China, as shown in Fig. 6.1. The steel production gradually increased
with the time, and the annual crude steel production was 100 million tons in 1996.
Currently, China has established a gigantic production capacity for iron and steel and
is the largest steel producer in the world. In 2021, the crude steel output was close
to 1030 million tons, accounting for over 50% of the global crude steel production.
Crude steel production per capita and steel cumulative per capita can be used as
representative indicators of industrialization and urbanization. Figure 6.2 shows the
crude steel production per capita and steel cumulative per capita in some developed
countries and China. The crude steel production per capita in developed countries
such as the United Kingdom and the United States has grown for a long time, reaching
a peak of around 650 kg. In Japan and South Korea, the peak of crude steel output
per capita is even higher, which is more than 1100 kg due to their advanced auto-
mobile manufacturing. The capita crude steel production of China reached 734 kg in
2020, exceeding the historical value of the United States, but still exists a certain gap
compared with Japan and South Korea. Due to the stringent environmental require-
ments and production capacity control policies of the Chinese government. The
capita crude steel production of China is not going to continuously increase, but it
will maintain a high level for a long time. From the perspective of capita steel cumu-
lative, the developed countries have a continuous growth process before reaching a
certain value. The steel cumulative will become stable when the annual consumption
and depreciation are roughly in balance. The United Kingdom and France completed
their industrialization with nearly 8.0 tons of the capita steel cumulative, while the
© The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd. 2022 183
X. Lv and Z. Yan, High Temperature Physicochemical Properties of High Alumina Blast
Furnace Slag, https://doi.org/10.1007/978-981-19-3288-5_6
184 6 The Revolution of High Alumina Slag in Blast Furnace Process

Fig. 6.1 China crude steel production from 1978 to 2020. Data from World Steel Association [1]

Fig. 6.2 Steel output (a) and steel cumulant per capita (b) for some developed countries and China

value was 11 tons for the USA and more than 12 tons for Canada and Japan. Therefore,
for a country to complete the industrialization and the infrastructure construction,
the capita steel cumulative should reach 8–10 tons. According to the current growth
rate, China is expected to reach this level after 2030.
Due to the development of world and the wave of global urbanization, the demand
for steel has increased year by year [1]. With infrastructure plans in developed and
developing countries, demand for steel remains strong. Today, about 70% of steel
is made through the blast furnace process. Although scrap steel can be recycled
through electric arc furnace process, it is not enough to meet the strong demand. It
6.1 The Inevitability of High Alumina Blast Furnace Slag 185

is foreseeable that the blast furnace ironmaking process will certainly be used in the
future.

6.1.2 Iron Ore Resources

The steel industry consumes about 1.6 times the iron ore while producing steel
productions for global society and industrial development. As of 2021, the whole
world has produced about 53 billion tons of steel and consumed about 85 billion tons
of iron ore. The distribution of the main iron ore resources in the world is shown
in Fig. 6.3. According to data from the United States Geological Survey (USGS),
in 2021, the global iron ore reserves are about 170 billion tons, and the reserves
calculated on the basis of metallic iron are about 83 billion tons. Therefore, if the
global production of pig iron and direct reduced iron is considered statically on a scale
of 1 billion tons per year, the iron ore resources will be consumed within 100 years.
China has relatively rich iron ore resources, but the grade is low, with an average
iron content of about 31 wt%, 11% lower than that of the average iron ore grade of
the world. Low-grade iron ore (high phosphorus, high alumina, Vanadium-titanium
magnetite) accounts for about 97% of the total reserves. Brazil and Australia are
rich in iron ore resources and are the main exporters of iron ore resources. After

Fig. 6.3 The distribution of the main iron ore resources in the world (data form USGS Mineral
Commodity Summaries). Data source US Geological survey: mineral commodity summaries,
January 2019
186 6 The Revolution of High Alumina Slag in Blast Furnace Process

decades of mining, steel companies have felt that the deterioration of raw mate-
rials has affected the normal smelting of blast furnaces. Table 6.1 shows the typical
iron ore composition in some countries. As the report from Rio Tinto, the Al2 O3
content in iron ore exported from Australia has increased from less than 0.5 wt%
in 2000 to the current about 2–3 wt%. It is foreseeable that with the consumption
of high-quality iron ore resources, the content of Al2 O3 in iron ore will continue to
increase. In addition, the iron ore in Brazil, India, and Sierra Leone also contains high
Al2 O3 content. For example, the Al2 O3 content in the iron ore produced by a mining
company in Sierra Leone is as high as 7–8 wt%. In addition, many high Al2 O3
iron ore deposits were found in India. For some high alumina iron ore resources,
because the geochemical and crystal-chemical behaviors of Al and Fe are similar, it
is easy to form isomorphic substitution in the ore, making the relationship between
aluminum and iron in the ore complicated. The physical beneficiation process has
a poor separation ability on the complex ore with aluminum and iron embedded,
resulting in a high content of Al2 O3 in the iron concentrate, which is difficult to meet
the current production requirements of the iron and steel industry [2]. At present, the
most feasible methods for processing high alumina iron ore are combined methods of
beneficiation and metallurgy, including sodium salt roasting-dissolving, and sodium
reduction-magnetic separation. However, these processes consume large amounts of
chemicals, cause serious environmental pollution, and are still far away to large-scale
industrial applications [3–5].
With the consumption of high-grade iron ore, resources have gradually become
depleted, and the iron ore available to steel industry has gradually shifted to low-
grade raw materials. It is foreseeable that the use of low-grade is an alternative
path to alleviate the deterioration of iron ore resources in the future. The ore with

Table 6.1 Typical iron ore composition in some countries (wt%)


Country TFe CaO SiO2 MgO Al2 O3
China 66.35 0.27 5.56 0.25 0.17
65.99 0.81 5.91 0.38 0.43
51.2 1.56 5.2 3.33 4.23
63.48 1.03 3.22 1.68 4.26
47.28 4.68 16.14 0.46 4.61
50.18 1.01 3.02 0.16 9.65
Brazil 65.75 0.22 3.79 0.08 0.81
65.6 0.15 2.82 0.08 1.03
66.7 – 1.6 – 1.9
India 66.85 – 1.52 – 1.41
64.67 – 0.57 – 2.3
Australia 62.3 0.17 1.55 0.07 1.98
62.55 0.05 3.44 0.12 2.17
Sierra Leone 56.96 0.07 1.7 0.08 7.09
6.2 Current Technical Routes for High Alumina Blast Furnace Slag 187

high alumina will inevitably become an important usable resource for future blast
furnace ironmaking production, resulting in the increase of alumina content in the
blast furnace slag.

6.2 Current Technical Routes for High Alumina Blast


Furnace Slag

The impact of high alumina on blast furnace operation has already attracted the
attention of the metallurgical industry. In recent years, many iron and steel enterprises
in China have used a large amount of iron ore with high Al2 O3 content to reduce
production costs. The use of such raw materials undoubtedly increased the Al2 O3
content in blast furnace slag. As shown in Table 6.2, the Al2 O3 content in blast
furnace slag of some plants has exceeded 15 wt%.
Aiming at the problems caused by the high Al2 O3 content of the slag in the prac-
tical production of blast furnaces, which referred to the increase in melting temper-
ature, viscosity, and the decrease of desulfurization capacity, some countermeasures
are suggested as the following according to the respective smelting conditions.

6.2.1 Basicity Control

According to the industrial experience, if the alumina content in the slag is low
(<14 wt%), and its physicochemical properties can fully meet the requirements of
blast furnace smelting when the basicity is about 1.2. With the increase of alumina
content, the slag composition area (region A) crosses the 1400 °C liquidus and
approaches the 1450 °C liquidus when the basicity is fixed, indicating that the melting
temperature of the slag increases, as shown in Fig. 6.1. Reducing the slag basicity

Table 6.2 BF slag


Plant MgO wt% Al2 O3 wt% Basicity
composition of some steel
(CaO/SiO2 )
enterprises in China
Taiyuan steel 8–10 <15 1.0–1.15
Meishan steel 8–10 <15.5 <1.2
Tangshan steel 11.5 <16 1.1–1.15
Wuhan steel 10–12 12.5–16 1–1.15
Jinan steel 9.5–10 15–15.5 1.2
Baoshan steel 8.1–9.1 14.22–16.81 1.01–1.24
Nanjing steel 10–12 15–17 1.05–1.1
Shao steel 10–12 16–18 1–1.1
Rizhao steel 10.52–13.81 14.91–18.34 1.08–1.13
188 6 The Revolution of High Alumina Slag in Blast Furnace Process

Fig. 6.4 Phase diagram of


CaO–SiO2 –9 wt%
MgO–Al2 O3 –1 wt% TiO2
slag system

is a method to reduce the melting temperature, as shown in the area B of Fig. 6.4.
Peng et al. [6] pointed out that when the alumina content in the slag is 17–20 wt%,
its melting temperature and fluidity meet the smelting requirements even through the
basicity decreased from 1.2 to 0.95. Cen et al. [7] reported that when the alumina
content in the slag is 10–14 wt%, it has no obvious effect on the desulfurization and
fluidity of the slag, but the increase of alumina to 17 wt% would worsen the stability
of the slag, and the basicity needs to be appropriately reduced to maintain the slag
stability.
However, the decrease in basicity increases the viscosity of the slag. Although
the results [6, 7] show that the fluidity still meets the smelting requirements, it
may weaken the kinetic conditions of desulfurization. On the other hand, due to the
decrease of CaO content and the increase of alumina content, the thermodynamic
conditions of desulfurization are also weakened, which has a great influence on
the quality of hot metal. This method can be used when the alumina content is low
(<17 wt%), otherwise, the change in fluidity would affect the blast furnace operation.

6.2.2 MgO/Al2 O3 Control

The reduction of slag basicity would not be the perfect route according to the reason
mentioned above. As the basic oxide, MgO and CaO are similar, therefore, adjusting
the MgO content in the slag becomes an effective slagging approach. On the basis of
theoretical analysis and experimental, Jiang and Shen [8] proposed a segmented
control technology on (MgO wt%)/(Al2 O3 wt%). The specific implementation
methods are as follows:
• Al2 O3 : <14 wt%
The alumina content is always lower than 14 wt% if the high alumina iron ore
is not used. The phase diagram of CaO–35 wt% SiO2 –MgO–Al2 O3 is shown in
6.2 Current Technical Routes for High Alumina Blast Furnace Slag 189

Fig. 6.5 Phase diagram of CaO–35 wt% SiO2 –MgO–Al2 O3 slag system

Fig. 6.5. When the alumina content is 14 wt%, the basicity in the range from 0.9
to 1.2, the change of MgO has no obvious effect on the liquidus temperature of
the slag, and the melting temperature of the slag system is less than 1400 °C. In
this case, there is no requirement for adjusting the content of MgO in smelting
operations.
• Al2 O3 : 15–17 wt%
The alumina content can reach 15–17 wt% in the blast furnace slag if the high
alumina iron ore is partially used. At present, the alumina content in the blast
furnace slag in most Chinese steel plants is in this range. It is recommended that
the lower limit of the (MgO wt%)/(Al2 O3 wt%) ratio be 0.3. As shown in Fig. 6.5,
increasing the MgO in the slag is beneficial to reduce the liquidus temperature of
the slag. However, MgO addition also makes the slag volume increase as well,
resulting in an increase in coke ratio and energy consumption. From the view
of fluidity and melting point, for slag with alumina content of 15–17 wt%, the
recommended range of (MgO wt%)/(Al2 O3 wt%) is 0.4–0.5, and the basicity is
1.15–1.2.
• Al2 O3 : 18–20 wt%
There are few blast furnace operations in China with alumina content in slag
exceeding 18 wt%, but the alumina content in blast furnace slag of Durgapur and
Bokaro steel companies in India reaches 18–20 wt%. With the increase in the use
of high alumina iron ore, the alumina in the slag may approach or exceed 20 wt%.
For this case, the (MgO wt%)/(Al2 O3 wt%) in the slag should be appropriately
190 6 The Revolution of High Alumina Slag in Blast Furnace Process

Fig. 6.6 Liquid phase


diagram of the
CaO–SiO2 –Al2 O3 –9 wt%
MgO–1 wt% TiO2 slag
system

increased and controlled between 0.45 and 0.55. The (MgO wt%)/(Al2 O3 wt%)
of Durgapur and Bokaro are 0.45 and 0.51, respectively. The slag volume in this
operation is slightly larger, but the operation is stable, and the technical indicators
are good, which proves the correctness and applicability of this range.

6.3 Feasibility of Revolution High Alumina Blast Furnace


Slag

As mentioned above, for the case the alumina in the blast furnace slag is between
15 and 20 wt%, there are two main control methods: reducing the basicity and
maintaining a suitable MgO content. But if the alumina content continues to increase,
these two methods will no longer work. For the case of the high alumina blast furnace
slag, the increase of MgO content result in magnesia-alumina spinel becoming the
first precipitation phase, and the increase of the melting temperature of slag. In order
to ensure the smooth operation of the blast furnace, the physicochemical properties
of the slag are required to be kept in a relatively stable state by reducing the silica.
In this case, according to the experimental measurements and model calculation
results, a revolution slag system (substitution of silica by alumina) is proposed, and
its feasibility will be analyzed from the slag-iron separation and slag desulfurization
capabilities below.
6.3 Feasibility of Revolution High Alumina Blast Furnace Slag 191

6.3.1 Slag-Metal Separation

The slag-metal separation requires that the slag and hot metal phases are completely
molten in the hearth. Generally, the hearth temperature of a blast furnace is 1450–
1500 °C which depends on the slag melting temperature since the liquidus tempera-
ture of hot metal (carbon-saturated) is about 1200 °C. The liquid phase diagram of the
CaO–SiO2 –Al2 O3 –9 wt% MgO–1 wt% TiO2 slag system is shown in Fig. 6.6. The
large-scale use of high alumina iron ore is gonging to inevitably lead to the gradual
transformation of blast furnace slag composition from silicate-based to aluminate-
based, that is, the slag composition changes from the current silicate blast furnace
slag system in region 1 to region 2 or region 3. Theoretically, these three regions are in
a molten state above 1450 °C. According to the experimental data, the free-running
temperature increases from 1390 to 1445 °C when the slag composition changes
from region 1 to region 2 (Fig. 4.6). The free-running temperature first increases
from 1390 to 1440 °C then decreases to 1397 when the slag composition changes
from region 1 to region 3 (Fig. 4.9), indicating that there is an operating composition
range that meets the temperature requirement of blast furnace practice in region 3
after overcoming the barrier.
The slag-metal separation requires that the iron droplets can quickly settle from
the molten slag, and the main factors affecting the sedimentation are the density
difference between the molten slag and the hot metal and the viscosity of the molten
slag. The greater density difference between hot metal and slag, the better separation
effect is. The density of hot metal is about 7200 kg/m3 , and the density of slag is
less than 3000 kg/m3 , which meets the density requirements of slag-metal separation.
According to the Stokes equation (Eq. 1.3), the smaller viscosity of the slag is benefit
to the separation of slag and hot metal. However, if the viscosity of the slag is too
low, the erosion of the refractory material of the furnace lining will be aggravated.
Therefore, the viscosity of the slag is required to be within a certain range in blast
furnace operation. According to the iso-viscosity diagram (Fig. 5.17) and Fig. 6.6,
the viscosity of actual composition from industry at 1500 °C is around 4.0 dPa s,
which is good for the blast furnace practice. Use this as a baseline, when the slag
composition transitions from region 1 to region 2, the viscosity increases rapidly;
when the slag composition transitions from region 1 to region 3, the viscosity is about
5.0 dPa s, indicating that there is an operating range in region 3 where the viscosity
meets the requirements of blast furnace fluidity and slag-metal separation.
The interfacial tension between the slag and iron droplet is also a factor that
affects the separation of the slag-metal. Only when the interfacial tension between
the two phases is large enough, are the iron droplets dispersed in the slag are easy to
aggregate and sink from the slag. The shape of the slag droplet on the liquid metal
is shown in Fig. 6.7. When the slag droplet is stabilized on the metal surface, the
resultant force of the three forces at point O is zero,

γMS = γM2 + γS2 − 2γM γS cos α (6.1)
192 6 The Revolution of High Alumina Slag in Blast Furnace Process

Fig. 6.7 The surface tension


and interfacial tension of
gas-slag-metal

where, γM , γS , and γMS are the surface tensions of slag, metal, and slag-metal, respec-
tively, mN/m; α is the contact angle, degree. The value of surface energy per unit area
and surface tension is the same, but the physical meanings of the two are different.
Surface energy is a measure of the breaking of chemical bonds between molecules
when creating a surface. In physical theory, surface atoms have more energy than
atoms inside the substance. Therefore, according to the principle of minimum energy,
atoms will spontaneously tend to the interior of the substance. Therefore, as the
surface tension of the slag increases, the contact angle α increases, and the wetta-
bility decreases, which is beneficial to the separation of slag and metal. In addition,
since the surface tension (γ ) of the slag increases, it means that the energy required
to generate bubbles in the slag increases, and it is difficult to generate bubbles.
In summary, when the slag composition changes from region 1 to region 3, the slag
system transformed from silicate to aluminate. The slag properties such as the free-
running temperature, density, viscosity, and interfacial tension meet the requirements
of blast furnace operation for slag-metal separation.

6.3.2 Hot Metal Quality Control

In order to ensure the quality of hot metal, blast furnace slag is required to have good
desulfurization ability in the process. The sulfur in the blast furnace comes from the
raw materials entering the furnace. About 80–90% of the sulfur keep by the slag and
hot metal, and only about 10–20% of the sulfur escapes out of the furnace with the
off-gas. The total distribution relationship of sulfur in the furnace can be expressed
by the following formula:
 
0.1 · w[S]load − w[S]gas
w[S]% = L S ·Q slag
(6.2)
1+ 1000

where, w[S]% is the sulfur content in hot metal, wt%, w[S]load is the sulfur load,
kg/t, w[S]gas is the sulfur escapes with off-gas, kg/t, L S is the sulfur distribution ratio
between the slag and metal, and Q slag is the slag volume, kg/t. The larger value of
the L S means the higher the concentration of sulfur in the slag, which has the same
meaning as the sulfide capacity (C s ). According to Eq. 6.2, the control of sulfur in
hot metal is based on the following:
6.4 Summary 193

• Lower sulfur load by using low-sulfur raw materials.


• Increase in slag volume.
• Increase in the sulfur distribution ratio of the slag.
Reducing the sulfur load is an effective method to obtain low-sulfur hot metal,
but this is determined by the sulfur content in raw materials. Since the blast furnace
is a reduction furnace, the amount of sulfur released by gas is limited. Increasing
the amount of slag should lower the sulfur content in hot metal. But increase in slag
volume also decreases the sulfur distribution ratio, which worse the sulfur diffu-
sion. The total effect of slag volume on the desulfurization is various depends on
the operation conditions. In addition, high slag volume means more flux and energy
consumption, reducing the production efficiency. Therefore, in blast furnace opera-
tion, an effective method is increasing the sulfur distribution ratio of the slag. The
distribution ratio of sulfur depends on the thermodynamic equilibrium of the slag
desulfurization reactions and the kinetic conditions of the reaction.
According to the Figs. 5.20 and 6.6, when the slag composition changes from
region 1 to region 2, the sulfide capacity gradually decreases, and the viscosity
increases rapidly, both thermodynamic and kinetic conditions deteriorate. When the
slag composition changes from region 1 to region 3. The sulfide capacity increases,
while the viscosity did not change significantly. This indicates that the slag system
in region 3 has a better ability to remove the sulfur from metal, which meets the
requirement of blast furnace slag on hot metal quality control.
Based on the analysis of slag-metal separation and hot metal quality, a slag
system suitable for blast furnace operation was found in the aluminosilicate slag
area (region 3 in Fig. 6.6). The composition range of the slag is: (38–42 wt%) CaO–
(25–30 wt%) Al2 O3 –(20–25 wt%) SiO2 –9 wt% MgO–1 wt% TiO2. Nakamoto et al.
[9] also reported that the 43.1 wt% CaO–35 wt% Al2 O3 –7.5 wt% MgO–14.4 wt%
SiO2 slag has a low melting temperature, and low viscosity that satisfies the fluidity
in blast furnace operation.

6.4 Summary

Due to the increase in steel demand and the decrease in high-quality iron ore
resources, high alumina iron ore will inevitably become the raw material of blast
furnaces, and the alumina content in the slag will increase accordingly. Currently,
when the Al2 O3 in slag is less than 20 wt%, the main methods to deal with it include
reducing the basicity and controlling the MgO content. When the alumina is more than
20 wt% in slag, a potential slag composition area in the range of (38–42 wt%) CaO–
(25–30 wt%) Al2 O3 –(20–25 wt%) SiO2 –9 wt% MgO–1 wt% TiO2 was proposed,
where the physicochemical properties meet the requirements for slag-metal sepa-
ration and hot metal quality. However, more research is needed on the granulation
and agglomeration, charge structure, reduction behavior, ventilation behavior, and
primary slag properties.
194 6 The Revolution of High Alumina Slag in Blast Furnace Process

References

1. World steel in figures 2021 (2022). World Steel Association. https://worldsteel.org/steel-by-


topic/statistics/world-steel-in-figures/
2. Zhang, S. (2013). Research on aluminum-iron separation technology of high alumina iron ore
(Ph.D. thesis), Northeastern University.
3. Jiang, T., Liu, M., Li, G., Sun, N., Zeng, J., & Qiu, G. (2010). A new process for the treatment
of high-alumina limonite by sodium reduction. Chinese Journal of Nonferrous Metals, 20(3),
565–571.
4. Jiang, T., Liu, M., Li, G., Sun, N., Zeng, J., & Qiu, G. (2010). Effect of sodium salt on the
separation of aluminum and iron by reduction roasting of high-alumina limonite. Chinese Journal
of Nonferrous Metals, 20(6), 1226–1233.
5. Li, G., Dong, H., Xiao, C., Fan, X., Guo, Y., & Jiang, T. (2006). Process mineralogy of high-
speed iron bauxite and separation technology of aluminum and iron. Journal of Central South
University, 37(2), 235–240.
6. Peng, Q., Yang, C., & Li, G. (2005). Study on the viscosity of high Al2 O3 blast furnace slag of
Xianggang. Journal of Wuhan University of Science and Technology. 28(1), 11–13.
7. Cen, M., Sun, L., Ye, B., Wen, Z., Xia, Y., & Qin, C. (1999). Research on the properties of
WISCO blast furnace slag. WISCO Technology. 37(6), 11-15.
8. Jiang, X., Shen, F., Han, H., Long, F., Zheng, H., & Gao, Q. (2019). Analysis and application of
blast furnace slag suitable for segmental control of magnesium-aluminum ratio. Iron and Steel,
54(10), 1–10.
9. Nakamoto, M., Tanaka, T., Lee, J., & Usui, T. (2004). Evaluation of viscosity of molten SiO2 -
CaO-MgO-Al2 O3 slags in blast furnace operation. ISIJ International, 44(12), 2115–2119.

You might also like