ISIJ International
Online ISSN : 1347-5460
Print ISSN : 0915-1559
ISSN-L : 0915-1559
Regular Article
Calculation of Physical Properties for Use in Models of Continuous Casting Process-Part 1: Mould Slags
Kenneth C. Mills Shyamprasad KaragaddePeter David LeeLang YuanFatemeh Shahbazian
Author information
JOURNAL OPEN ACCESS FULL-TEXT HTML

2016 Volume 56 Issue 2 Pages 264-273

Details
Abstract

Physical properties of both steels and mould slags are needed as input data for the mathematical modelling of the continuous casting process. Routines for calculating the properties of mould slags and for estimating steel properties have been developed and are described in Parts 1 and 2, respectively. Many mould powders, with differing compositions, are used in casting practice and their properties vary significantly. Reliable models have been developed to calculate these property values as a function of temperature from their chemical composition since this is available on a routine basis. Models have been developed to calculate the following properties: heat capacities, enthalpies, thermal expansion coefficient, density, viscosity, thermal conductivity and surface tension. Solid mould slags can exist as glassy or crystalline phases or as mixtures of the two (i.e. slag films) and the properties for the various phases can vary considerably; methods have been developed to calculate property values for these various states. The software used to calculate the properties is available via the link (i) http://www.mxif.manchester.ac.uk/resources/software (ii) https://sites.google.com/site/shyamkaragadde/software/thermophysical-properties.

1. Introduction

Recent mathematical models1,2,3) have shown that that they are capable of providing (i) accurate predictions of both the heat flux and the powder consumption (a measure of the lubrication supplied to the shell) and (ii) valuable insights into the mechanisms underlying problems and defects (e.g. slag entrapment and oscillation mark formation). These models are being further developed to identify the causes of defects. However, models require accurate values for the physical properties of both the steel and mould slag (as input data). Properties of both mould slag and steel vary considerably with composition and a large number of steels and mould powders are in use. Consequently, a model of the continuous casting process requires a large amount of property data for the steel and mould powder used in the casting. The most satisfactory solution is to develop routines to calculate the required properties from their chemical composition, since the latter is available on a routine basis. The relationships developed to estimate the physical properties of mould slags from their chemical composition are described here in Part 1 and those used for steels are described in Part 2.The two models are linked to calculate the metal/slag interfacial tension.

Mould powders play an important role in the continuous casting process. They are added to the top of the mould and as they heat up the following events occur, sequentially (i) carbonates decompose to form CO2(g) (ii) Carbon particles in the powder start to oxidise (iii) the oxide components sinter and (iv) then melt when the carbon is consumed, and molten globules mix to form a molten slag pool. Slag from the molten pool then infiltrates into the channel between the steel shell and the mould to form a slag film (Fig. 1); this consists of (i) a liquid layer (ca. 0.1 mm thick) and (ii) a solid layer (ca. 1–2 mm thick). The heat flux between shell and mould is determined, principally, by the thickness of this solid layer (and the amount of crystalline phase present) and the lubrication supplied to the shell is determined by the thickness of the liquid layer. Lubrication involves the following properties: viscosity (η), density (ρ) surface tension (γ) and break (or solidification) temperature (Tbr). The heat flux involves the following properties of the mould slag: thermal conductivity (k), Tbr, the fraction of crystalline phase formed in the slag film (fcrys) and the optical properties of the mould slag.

Fig. 1.

Schematic drawings showings showing the various layers formed in the mould for (a) the meniscus area and (b) half-section of the mould.1,3)

The objective of this study was to develop routines to provide reliable values of the thermo-physical properties of mould slags from their chemical compositions, for input into the model of the continuous casting process. The other aim was to make these routines available as Excel software; open access is provided at (i) http://www.mxif.manchester.ac.uk/resources/software (ii) https://sites.google.com/site/shyamkaragadde/software/thermophysical-properties.

1.1. Effect of Structure on Slag Properties

The properties of slags are principally determined by the structure of the slag. In silicate slags the basic building block is the Si–O tetrahedron, in which one Si4+ cation is surrounded by 4 O ions. Furthermore, in pure silica, each of these O ions bonds to Si4+ ions in other tetrahedra to form a 3-dim. structure (these bonds are known as bridging O s (BOs)). When a cation, such as Na+, is introduced into SiO2, the Na+ breaks the bridging O bond and forms a non-bridging O (NBO); this results in de-polymerisation of the slag. In highly-depolymerised slags some O ions are not bonded to Si ions, these are referred to as free oxygens (FOs).

In alumino- silicates, Al3+ ions are readily introduced into the Si4+ chain or ring but a Na+ or 0.5 Ca2+ is needed for charge-balancing (i.e. to form NaAl4+); cations on charge-balancing duties are not available for network-breaking. Thus the introduction of Al3+ ions result in an increase in the polymerisation of the slag, however, the Al–O bond is weaker than the Si–O bond.

Some properties, such as viscosity, electrical resistivity, diffusion coefficient and thermal conductivity are very dependent upon the degree of polymerisation. Several parameters have been used to represent the degree of polymerisation. In this paper the parameter, Q, is used; it can be calculated from Eqs. (1) and (2) where MO=CaO, MnO etc. and M2O=Na2O etc. Summation index NMO etc. represents the number of different oxides. Other properties, such as density, are affected by the structure but to a much smaller extent than that in the above properties.   

NBO/T = 2( 1 N MO X MO + 1 N M2O X M2O - 1 N M2O3 X M2O3 ) ( 1 N MO2 X MO2 -2 1 N M2O3 X M2O3 ) (1)
  
Q=4-NBO/T (2)

The O ions bonded to cations tend to be arranged into an octahedral array (i.e. they have 6- fold coordination) but this varies with cation size.4) In general, the effect of different cations on the physical properties is much smaller than the effect of polymerisation. Nevertheless, the nature of the cations affects both the structure and the physical properties. The parameter usually used to represent the effect of cations is the field strength (=z/r2 where z and r are the charge and radius of the cation). Increasing field strength has been reported to bring about (i) a wider spread of polymerised silicate anions (i.e. a wider spread of Qn species) (ii) a decrease in coordination number which affects the concentrations of BOs and NBOs (iii) increasing disorder in the melt or glass and (iv) a lower probability that they will carry out charge- balancing duties (which are usually carried out in the hierarchy K+>Na+>Li+ ≈Ba2+ etc). Cations affect properties (i) directly or (ii) indirectly through their effect on the structure; two examples of direct effects are (a) the electrical conductivities (κ) and diffusion coefficients (D) are affected by the number of available cations and for XNa2O=XCaO there are twice as many Na+ ions as Ca2+ cations (so κ Na2O slag>κ CaO slag and DNa2O slag>DCaO slag) and (b) the mixed alkali effect occurs where some properties for slags are significantly affected by cations with differing size (e.g. K+ and Li+).

Increasing temperature has a similar effect to that of increasing the number of cations, i.e. it reduces the degree of polymerisation; thus increasing temperature reduces viscosity and increases electrical conductivity and diffusion rates.

1.2. Crystalline, Glassy and Liquid Phases

In the solid state, slags can exist as crystalline or glassy phases or as a mixture of the two phases (e.g. slag film). In crystalline solids, the ions occupy regular positions and the amount of disorder (or configurational entropy) is low. In contrast, in glasses the ions are more disordered and the configurational entropy is much higher than that in crystals.

The formation of glasses is promoted by (i) slags with high Q (degree of polymerisation) values (i.e. high SiO2 contents) and (ii) fast cooling rates. When a frozen glass is heated, it transforms at the glass transition temperature (Tg) into a supercooled liquid (scl). This transition is accompanied by step-increases in both Cp and thermal expansion coefficient (α) and results in a concomitant increase in configurational entropy.

When a liquid slag is cooled, if crystals are precipitated, there is a sharp increase in viscosity at a certain temperature, often denoted as the break temperature (Tbr). In contrast, if a scl is formed the viscosity increases smoothly with decreasing temperature.

2. Model Details

2.1. Composition, Modelling Parameters

The mould slag compositions covered by the model include conventional F- containing powders and F- free powders but do not include, at this time, the mould slags (based on calcium aluminates) now being used to cast high- Al, high- Mn steels. Furthermore, B2O3 additions have been reported to both reduce and increase slag viscosities of mould slags by various investigators; for this reason the effect of B2O3 on viscosity has been excluded.

As –received, casting powders contain both free carbon and carbonate; these constituents are lost on ignition (LOI) in the process of forming a casting slag. The fraction of powder forming slag (f*) can be calculated from the chemical composition using either Eqs. (3) or (4)   

f*=( 100-%LOI ) /100 (3)
  
f*=(100-% C free -( 44/12 ) ( % C total -% C free ) (4)

The composition of the casting slag can then be calculated by (i) multiplying by (1/f*) and normalising the composition. The software (covering these models) has the option to take into account typical pick-up values for Al2O3, FeO and MnO during casting. If this option is used the slag composition is re-normalised.

Mole fractions (X) are then calculated from these normalised compositions using Eq. (5) (where i refers to the constituents e.g. CaO, SiO2 etc.); the %F present is converted into XCaF2 and the Ca2+ associated with the F- ions is deducted from XCaO.   

X i =( X i / M i ) /Σ( X i / M i ) (5)

The mean molecular weight (M) is calculated using Eq. (6)   

M=Σ( X i M i ) (6)

The parameter (NBO/T) is used as a measure of the de-polymerisation of the slag. It is calculated from Eq. (1) where Σ XMO=XCaO+XMgO+XFeO etc. and Σ XM2O=XNa2O+XK2O etc. The parameter, Q is a measure of the polymerisation of the slag and is calculated from (NBO/T) using Eq. (2). It should be noted that both XCaF2 and XTiO2 are ignored in Eq. (5) (on the basis of structural evidence) and the calculated Q value is that of the remaining slag (Qrem).

The fraction of crystalline phase in the slag film (fcrys) was calculated using the relation proposed by Li et al.5) where (NBO/T)* indicates that the parameter differs from that calculated with Eq. (1), in that (i) CaF2 is included in network breakers and (ii) when %MgO and %MnO are >7% and >4%, respectively, they are included in the denominator of Eq. (1).   

f crys =-2.84+1.41( NBO/T ) * (7)
  
% crys =100.  f crys (8)

2.2. Temperatures (Tg, Tcrit, Tliq and Tbr)

These temperatures are calculated from chemical composition by the software and the temperature scale in the program is adjusted automatically.

The glass transition temperature (Tg) is the temperature where, on heating, a frozen glass transforms into a super-cooled liquid (scl). This transition is accompanied by step-increases in Cp and thermal expansion coefficient (TEC) (i.e. 3α<Tgα>Tg). The viscosity of a glassy slag is taken to have a value of 1013.4 (d Pas) at Tg (i.e. log10 η (d Pas)=13.4)).

Numerical analysis of a database of Tg values (840–910 K) obtained from Cp and TEC data6) and chemical composition data for 23 mould slags was carried out in this study. Inspection of Tg values for mould slags indicates that values of mould slags tend to have similar values (Tg=870±30 K). The best fit results are shown in Eq. (9) and are used in the software. The uncertainty in the calculation is ca. ±20 K.   

T g ( K ) =906-330.5 X SiO2 +190 X CaO +440 X Al2O3 - 449 X N2O+K2O -11 X MgO +154 X CaF2 -309 X MnO -1391 X FeO (9)

The critical (or deformation) temperature (Tcrit) is the temperature, above which, the thermal conductivity, decreases rapidly with increasing temperature (this can be seen in Fig. 6). Thermal conductivity- viscosity plots reveal that Tcrit occurs when log10 η Tcrit (dPas)=6 i.e. between the softening and flow temperatures.7) Values of the thermal conductivity as a function of temperature for a number of mould slags indicate that Tcrit occurs at 1040±10 K.8,9)

Fig. 6.

Thermal conductivity of mould slags as a function of temperature; Glassy phase:=faint line=kTHW7) and ●,○ kLP values for solid and liquid calculated from 107a=4.5 and 4 m2s−1,respectively;16,18,28,29,30,31) partially- crystalline slags= X, + and dashed, bold line and bold line: Δ,□=partially-melted powders32) values recorded on cooling after melting.

The liquidus temperature (Tliq) is the temperature where the slag becomes fully molten. A numerical analysis of Tliq values (from DSC experiments) and composition data gave Eq. (10); the calculated values are subject to uncertainties of ±30 K.   

T liq ( K ) =1464-+11.4%SiO2-11%CaO+4.2%A l 2 O 3 +5.7%MgO-10.1%Na2O-15.8% K 2 O+10%L i 2 O  +1.9%F+8.3%FeO+11.6%MnO (10)

More recently, numerical analysis was carried out on a database of Tliq and chemical composition data for 23 mould slags6) and the Eq. (11) was obtained and is preferred. Values are subject to uncertainties of ±25 K.   

T liq ( K ) =1473-1.518%Si O 2 +2.59%CaO+1.56%A l 2 O 3 -17.1%MgO-9.06%N a 2 O-6.0% K 2 O+18%L i 2 O +4.8% F-9.87%FeO-2.12%MnO (11)

The solidification temperature (Tsol) is the temperature where solids are first precipitated on cooling. It is frequently represented by the break temperature (Tbr) which is the temperature below which the viscosity shows a sharp increase in viscosity during cooling. The break temperature decreases with increasing cooling rate and often Tbr is cited for a cooling rate of 10 Kmin−1. The thickness of the solid slag layer (Fig. 1) is partially- determined by the break temperature. The following equation was obtained from numerical analysis of Tbr and chemical composition data;10) these values are subject to uncertainties of ±30 K.   

T br ( K ) =1393-3.3%Si O 2 +8.65%CaO-8.45%A l 2 O 3 -17.1%MgO-3.2%N a 2 O-2.2% K 2 O-6.6%L i 2 O -6.47% F-18.4%FeO-3.2%MnO (12)

In general, super-cooling will ensure that Tbr<Tliq. However, the calculation of both Tbr and Tliq involve an uncertainty of ±25–30 K, so cases where Tliq<Tbr do occur occasionally; in these cases a warning is issued on the Collected Worksheet.

2.3. Heat Capacity (Cp) Enthalpy (HT−H298)

The Cp- T relationships are different for different phases. Consequently, the mould slags are first sorted into individual phases i.e. (i) crystalline (b) glass (iii) slag film and (iv) liquid. The Cp and enthalpy of mould slags are little affected by the slag structure and it has been shown11,6) that they can be calculated using routines based on partial molar quantities   

P= i=1 N ( X i P i ) (13)
where P is a property, or coefficient, X is the mole fraction and i=1,2,3,…N denotes the different components (e.g. CaO, SiO2 etc.).The values of Cp and (HT−H298) are calculated in units of JK−1mol−1 and kJK−1 kg−1, respectively; values were converted from mol−1 to kg−1 by multiplying by (1000/M) where M, the average molecular weight of the slag.

The temperature dependence of Cp for crystalline solids is usually expressed in the form:   

C p =a*+b*T-c*/ T 2 . (14)
where a*, b* and c* are constants (given in Table 1). In the model the parameters “a*” “b*” and “c*” are treated as partial molar quantities:10) e.g.   
a * = i=1 N ( X i a * i ) (15)
Table 1. The coefficients used in the calculation of Cp for crystalline phase Cp (JK−1mol−1)=a*+b*T-c*/T2; Cp for liquids (JK−1mol−1): and ΔSfus: (JK−1mol−1).
SiO2CaOAl2O3MgONa2OK2OLi2OFeOMnOCaF2ZrO2TiO2B2O3
a*56.048.811542.765.765.765.74946.459.869.675.2
103b*15.44.5211.87.4522.622.622.68.378.130.457.531.17
10−5c*14.46.5335.16.20002.83.88−1.9714.118.2
Cp(l)8790.8146.490.492.174.196.276.679.996.2113111.7
ΔSfus4.624.750.224.733.933.933.99.618.81825.126.8

The parameters b* and c* are calculated in a similar manner to a*.

The enthalpy (HT−H298) is given by:   

( H T - H 298 ) = 298 T C p dT =a*( T-298 ) +0.5b*( T 2 - 298 2 ) +( c*/T ) -( c*/298 ) (16)

The enthalpy of fusion, ΔHfus, is calculated by assuming the entropy of fusion (ΔSfus) can also be calculated from partial molar terms.   

Δ S fus = i=1 N ( X i Δ S i fus ) (17)
  
Δ H fus = T liq Δ S fus (18)

The Cp for the liquid phase is also calculated from partial molar terms11) using Eq. (19) and the enthalpy for the liquid is determined from Eq. (20).   

C p = i=1 N ( X i C pi ) (19)
  
( H T - H 298 ) = ( H Tliq - H 298 ) sol +Δ H fus + C p liq ( T- T liq ) (20)

The Cp for the glass phase is close to that of the crystalline phase for temperatures below Tg. At Tg the frozen glass transforms to a super-cooled liquid (scl); this is accompanied by a step- increase in Cp. Measured values of Cp for the scl (Cp scl) were between 1400 and 1500 JK−1kg−1.6,12) These values are in close agreement with the value calculated for the liquid.

Consequently it was assumed that Cp scl=Cp liq for the range between Tg and Tliq. Values of the enthalpy for the scl in the temperature range between Tg and Tliq are given in Eq. (21).   

( H T - H 298 ) scl =( H Tg - H 298 ) + C p   scl ( T- T g ) (21)

It should be noted that:

(i) Crystallisation occurs around 50–100 K above Tg (this is an exothermic reaction and results in an apparent decrease in Cp in DSC measurements); thus the state of the sample remains largely undefined in this temperature range.

(ii) When (HTliq−H298) is calculated for both crystalline and scl it was found that (HTliq−H298)cryst>(HTliq−H298)scl which suggests that ΔHfus ≠ 0 for the scl; high temperature Cp measurements are needed to resolve this anomaly.

The slag film consists of a mixture of glass and crystalline phases; the Cp and enthalpy for a slag film is calculated using the rule of mixtures (Eqs. (22) and (23), respectively).   

C p slag film = f cryst C p cryst +( 1- f cryst ) C p gl (22)
  
( H T -  H 298 ) slag film = f cryst ( H T - H 298 ) cryst +( 1- f cryst ) ( H T - H 298 ) gl . (23)

There are only a few measurements of the Cp and (HT−H298) available for mould slags6,12) and these are restricted to temperatures below 1000 K. there are no reported measurements for the liquid. Thus the model is based on a generic data for oxides and fluorides. Although, the predicted Cp and (HT−H298) values are within ±2% of the experimental values for T<1000 K, the uncertainty is probably ±5% for T>1000 K. It should also be noted that when a glass is heated to a temperature of ca. (Tg +80 K) the sample will partially crystallise and the exothermic enthalpy released causes a sudden apparent decrease in Cp values measured by DSC; this is not a true effect and the values calculated by the model will not show this effect.

2.4. Density (ρ) Thermal Expansion Coefficient (α)

Thermal expansion coefficients reported for 10 glassy samples over the temperature (298−Tg) range between 9 and 11×10−6 K−1 and for one sintered mould flux (298–1100 K).6) These measurements indicated that α decreased slightly with increasing Q (106 α=15.2–2.14 Q) but the scatter was such (R2=0.41) that a constant value of α=10×10−6 K−1 has been preferred in the calculations. There are few data available for partially-crystallised samples of mould slag; one sample6) exhibited a value of α(298–1100 K)=10×10−6 K−1; this value was adopted.

The thermal expansion coefficient data for binary silicates in the liquid indicate that it is dependent upon both Q and (z/r2).13) However, there are few reported ρ- T data for liquid mould slags and individual ρ values are subject to uncertainty. Consequently, thermal expansion values for the liquid phase were obtained with a generic model for oxide slags11) containing CaF2.

The densities of silicate and alumino-silicate slags exhibit only a small dependence on the slag structure. Reasonable estimates can be obtained using partial molar volumes (V) for the various slag constituents but special procedures are needed for SiO2 and Al2O3.11)

There are few reported data for the density of mould slags. Values have been reported for the glassy phase14,15,16) the liquid state17,18) and for some slag films.12)

The density of the liquid mould slag is calculated from partial molar volumes (V=M/ρ where M=molecular weight) as shown in Eqs. (24), (25), (26) (using the data given in Table 2) but special treatment is given to SiO2 (VSiO2=19.55 +7.966X SiO2) and Al2O3 (VAl2O3= 28.31+32XAl2O3 +31.45 (XAl2O3)2) to account for the effect of structure on the molar volume and hence the density (ρ).11) These routines were calculated from studies of density values for a wide range of slag compositions.11)   

V 1773 = i=1 N ( X i V i ) (24)
  
V T = V 1773 +0.01( T-1773 K ) (25)
  
V T =M/ ρ T (26)
Table 2. Partial molar volumes (V1773) used in the calculation of the density of the liquid; a denotes for VSiO2=19.55 +7.966X SiO2) b for Al2O3 (VAl2O3=28.31+32XAl2O3 +31.45 (XAl2O3)2) and V298 for the solid; c for SiO2, VSiO2=23.76+3.5X SiO2.; V1873 and (dV/dT) in the 4th and fifth row are from Stebbins21); * indicates calculated here.
SiO2CaOAl2O3MgONa2OK2OLi2OFeOMnOCaF2ZrO2TiO2B2O3
V1773(l)[11]a20.7b16.13351.81615.815.631.32410
V298c14.440.412.520.233.51116.51728.5
V1673 *[21]26.7516.937.711.9529.548.517.7*13.713.3*31*24.2527*
103dV/dT[21]−0.423.971.022.227.421.5*4.154*5.1*0.84*

Stebbins et al.19) have also reported values of V1873 and dV/dT based on regressions of V-T and (dV/dT) data for most of the components shown in Table 2. These data were used to calculate both V1873 K and (dV/dT). Values for the missing components in Table 2 were derived here from ρ-T and V-T data.

The recommended values of VT were calculated using the V1773 value and Eq. (27) and are preferred to values calculated by Eq. (25). The values calculated using the Stebbins value for V1873 are given on the Density Worksheet. Although VT values in Table 2 vary, the density values calculated with the Stebbins parameters lie within 1% of those calculated by Eq. (27).   

V T =Σ X i V 1773 +Σ X i (dV/dT)(T-1773   K) (27)

A similar approach to that used for the liquid is taken for solid phase with different values for the molar volume (Table 2) and (VSiO2=23.76+3.5X SiO2) and (VAl2O3=40.4). However, when a glassy sample crystallises it is accompanied by shrinkage (since ρcrys>ρgl). This results in porosity in the sample; the porosity levels were not determined for the density measurements on slag films12) but would be expected to lower the density values.

There are very few data reported for mould slags. The glassy phase has a lower density than the crystalline phase but there are no experimental data for mould slags to validate this statement. Thus, it has been assumed that ρgl=ρcrys. for temperatures between Tg and Tliq; the density of the super-cooled liquid was calculated by assuming that (i) there was no density change associated with the (scl → liquid) transition (ii) ρscl is a linear function over this temperature range (i.e. ρT=ρTg+(T−Tg)(ρTliqρTg)/(Tliq−Tg).

Values of ρT for solids (both the glassy and sintered samples) were calculated using a value of α=10−5 K−1.Normally, the calculations produce a density decrease at Tliq but the uncertainty in the calculations of ρT for both solid and liquid phases can lead (with certain compositions) to an apparent increase in density at Tliq; this is not a true effect and is caused by the combined uncertainties. In such circumstances a warning is given on the Collected Worksheet.

There are few reported density data for mould slags. Nevertheless, the estimated values for the density for temperatures below 1000 K, the uncertainties are probably 2–3% and are within ±5% for T>1000 K. Data are needed for the density change associated with glass→ crystalline phase change.

2.5. Viscosity (η)

The principal factor affecting the viscosities of slags is the degree of polymerisation present in the slag; this can be clearly seen in Fig. 2 where both ln η1900 K and the parameter, Bη exhibit considerable sensitivity to the parameter, Q (which is a measure of the polymerisation). The scatter of individual points in Fig. 2(a) is due to the effect of different cations (with viscosity increasing with increasing cation size or decreasing field strength (z/r2)).

Fig. 2.

(a) Viscosity (ln η1900 K) and (b) the parameter, B η, as functions of Q; note the upper curve in (b) is for MO-silicates and the lower curve is for M2O-silicates.7)

The introduction of Al3+ into the Si4+ structure for slags with Q>2.8 results in a small linear increase in viscosity with increasing Q for alumina-silicates, in contrast with the dramatic increase in viscosity with increasing Q for silicates shown in Fig. 2(a).

Viscosity measurements on mould slags have been reported by several workers.20,21,22,23,24,25) Experimental uncertainties can be up to 25%; McCauley21) cites six sets of viscosity measurements on the same slag and the reported viscosities were found to vary by ±25%. Another source of uncertainty is that few workers cite chemical analysis on the post-measurement sample and fluorine and sodium losses can be significant during the experiment.

The new model to calculate the viscosities of mould slags is based on an earlier model proposed for slags in general7) but which has required certain modifications in applying it to mould slags. In the original model, ln(η1900 K) and the activation energy term, Bη, are represented as functions of Q as double exponential relations (Fig. 2). The small deviations from the curve due to cation effects can be accommodated in terms of the field strength for the cations (i.e. mean z/r 2).The activation energy term, Bη can also represented as a function of Q by two different double- exponential equations for MO-silicates and M2O-silicates (Fig. 2(b) and Eqs. (28) and (29)). In mould slags, the charge balancing duties are carried out by K+ and Na+ ions, and this coupled with the fact that XMO>>XM2O implies that the upper curve in Fig. 2(b) and Eq. (28) is more appropriate for casting slags. The viscosities at temperatures other than the reference temperature can be calculated from the Eq. (30).   

MO-silicates:  B η =795830.8+8.476 × 10 -6 exp( Q/0.2611 ) +795841.4exp( Q/211587.4 ) (28)
  
M 2 O-silicates:  B η =-24.6+6.122 × 10 -13 exp( Q/0.1258 ) +28.465exp ( Q/7.727 ) (29)
  
ln η T =ln η 1900 +( B η /T ) -( B η /1900 ) (30)

The above model must be modified to apply it to mould slags since:

(i) the reference temperature, 1573 K, is far removed from 1900 K; this would have little effect on Bη but would have a dramatic effect on the ln η term since viscosity is sensitive to temperature changes.

(ii) both Bη and ln η1573 K are affected by both Q and the CaF2 content, simultaneously.

Consequently, it is necessary to separate the effects of CaF2 and Q. In the case of Bη (little affected by temperature) a value for a CaF2-free slag can be obtained from Eq. (28); this is denoted as a solid curve in Fig. 3(a). It can be seen that Bη (i) increases with increasing Q and (ii) departs from the curve with deviations increasing with increasing CaF2.Since Bη is dependent upon two factors (Q and XCaF2), the various slags were then sorted into individual groups according to their XCaF2.values (e.g. (0–0.05); (0.05–0.1); etc.) and allocated a mean XCaF2 value (e.g. 0.025; 0.075, respectively). The departure of experimental values (Bη expt) from the curve (i.e. from Eq. (28), calculated using the remaining slag composition, Bη rem Eq. (28)) is denoted ΔBη, (defined in Eq. (31)) and was then plotted as a function of XCaF2 for specific groups of similar Q (e.g.{1.5±0.25}; {2±0.25} etc.); the results are shown in Fig. 3(b).   

Δ B η = B η rem    Eq 28 - B η expt (31)
  
B η =( B η rem Eq 28 -Δ B η ) = B η rem Eq 28 -32.7 X CaF2 (32)
Fig. 3.

(a) The parameter, Bη, as a function of Q for slags containing varying XCaF2: for following XCaF2 ranges: line=0; o=0–0. 05; ◊=0.05–0.1; Δ=0.1–0.15; ♦=0.15–0.2; ■=0.2–0.25; ●=>0.25 and (b) ΔBη as a function of XCaF2.

The same approach can not be adopted to treat the ln η1573 K data because of the large differences in ln η values for 1900 K and 1573 K. Consequently, it is difficult to calculate a value of ln η1573, CaF2=0 for CaF2-free slags, since both Q and XCaF2 affect the viscosity simultaneously. The experimental data were first sorted into the same groups of similar XCaF2 values and these values were then plotted as function of Q (Fig. 4(a)) and equations were derived for ln η1573 K as a function of Q. A linear best fit of the data (ln η1573 K=3.073+1.93Q) was then derived since the relationship is close to linear in the range Q=1.5 to 3.Then the database was divided into groups of similar Q values and plotted against the mole fraction of CaF2 (Fig. 4(b)). The mean gradient of these plots (d ln η1573/dXCaF2)=−9 was determined from these plots. Values of ln η1573, CaF2=0 were then calculated using Eq. (33); the values of ln η1573, CaF2=0 were then plotted against Q (Fig. 5) and the best fit curve for ln η1573, CaF2=0 is given in Eq. (34). It can be seen that the mean scatter in ln η1573 values is about 0.3.which corresponds to an uncertainty of ca. ±25% in η. Attempts will be made in the future to reduce this uncertainty from a best fit regression of ln η1573 data as functions of Q and XCaF2. The calculated viscosity values are compared with values calculated by the Riboud26) and Iida models27) in the Viscosity worksheet and mean values for the three models were also calculated.   

ln η 1573, CaF2=0 =l n η1573 -9 X CaF2 (33)
  
ln  η 1573K, ref =0.1227exp ( 1.1454 Q ) (34)
Fig. 4.

The parameter, ln η1573 K as functions of (a) Q; XCaF2=(0.025±0.025)=O,+; black line (0.075±0.025)=♦ and red line,; (0.125±0.025)=▲ and green line;(0.175±0.025)=♦; (0.225 ±0.025)=■; >0.25=● and (b) mole fraction CaF2; O=(Q=1.5); ■=(Q=2); ▲=(Q=2.5); ♦=(Q=3).

Fig. 5.

Values of ln η1573, CaF2=0 for various mould slags as a function of the parameter Q; ◊=Lanyi;20) ■, x=McCauley21) ●=Persson22) ♦ Elahipanah.23)

Although viscosity measurements for supercooled liquids are available for minerals and glasses, measurements on slags are rare for metallurgical slags and the authors were unable to obtain any experimental data for the scl state of mould slags. However, it is possible to calculate values for these slags in the scl range since log10 η (d Pas) has values of 13.4 and 6, respectively, at Tg and Tcrit (=1040 ±10 K). These data can be coupled with calculated values for the liquid range at Tliq, 1573 and 1673 K. Attempts were made to obtain a best fit for these data using a polynomial relation (log10 η(d Pas)=a+bT+cT2+dT3 where a, b, c and d are constants). Unfortunately, the resulting value for log10 η at Tg departed significantly from log10 η (dPas)=13.4 in some cases. Consequently, values of log10 η (d Pas) were obtained from linear relationships between Tg and Tcrit and between Tcrit and Tliq; these values should not be regarded as accurate values for the scl but do provide the reader with approximate values which are unavailable in the literature.

The performance of the new model was tested by calculating the viscosity at 1573 K and comparing it with the experimental values for ca. 30 mould powders.20,21,22,23,24,25) The slags were randomly selected to give a range of η1573 values (0.5–12 dPas) but slags containing >1% B2O3, >1% ZrO2 and >3%FeO were excluded. The deviation (Δ=100 (ηcalc-ηexp)/ηexp) was calculated and the mean deviation determined (Δmean=[ Δ12+ Δ22+ Δ32+...]0.5). Then the Δmean values were compared with those for the Riboud26) and Iida27) models. It was found that performances of the new model, the Riboud and Iida models were similar (Δmean=35%; 35% and 43%, respectively). It was found that the mean η1573 value for the three studies was closer to the experimental values (Δmean=27%). It should be noted the experimental uncertainties associated with both the viscosity measurements and changes in chemical composition during the measurement sequence are ca. ±25% for most of the reported data.

2.6. Thermal Conductivity (k) and Thermal Diffusivity (a)

The thermal conductivity of the slag film is important in controlling the heat transfer from the shell. Thermal conductivities of liquid slags are difficult to measure since it is difficult to eliminate contributions from convection. This is usually achieved by using transient methods where the experiment is completed before convection is initiated. The two principal, transient techniques are the laser pulse (LP) and the Transient Hot wire (THW) methods.

However, glassy and liquid slags are also semi-transparent to infra-red radiation and heat transfer can occur simultaneously by both lattice conductivity (klat) and radiation conductivity (kR).Most investigators try to determine the values of klat and kR, individually and then combine them to give a total value. However, recent measurements of klat on mould slags using the LP and THW methods reveal that:

(i) Values of kLP16,18,28,29,30,31) and kTHW8,9,35,36,37) are in good agreement up to 1000 K (Fig. 6) but deviate markedly at higher temperatures.

(ii) The kTHW values for mould slags drop markedly for temperatures >1040 K, in contrast to kLP values which continue to rise (Fig. 6); this contradictory behaviour has been attributed variously, to electrical leakage in the THW measurements or to kR contributions to kLP in the LP measurements.

The source of the discrepancy between kLP and kTHW is unresolved at the present time. Here it has been assumed that kLP measurements contain contributions from kR and the kTHW measurements have been adopted.

The thermal conductivities of slags are also affected by (i) the degree of crystallinity developed in the slag (i.e. fcrys) since kcrys ≈ 2 kglass due to the higher packing density in the crystalline phase than in the glass and (ii) porosity which tends to reduce the thermal conductivity. Note crystallisation is accompanied by shrinkage and results in the formation of fine pores in slag films.

Plots of kTHW for glassy, M2O–CaO–SiO2 slags versus the calculated viscosities indicated that the sharp drop in kTHW occurred at the temperature where η=106 d Pas;7,13) thus Tcrit occurs midway between the softening temperature (where the sample can no longer support its own weight) and the flow temperature.7,13) The collapse in kTHW occurs for the scl phase (i.e. in glassy samples). This strongly suggests that the magnitude of the thermal conductivity is linked to the rigidity of the silicate lattice.7,13) Measured kTHW values ranged from 1.05–1.09 Wm−1K−1 at 298 K for glass samples;8,9) so a value of k298=1.07±0.02 Wm−1K−1 was adopted for all glassy mould fluxes.

The following values were also reported for the glassy phase, Tcrit=1040 ±10 K and kTcrit=1.65±0.05 Wm−1K−1.7) Measurements of kTHW for 15 mould slags indicated that k1040 K= 1.65±0.05 Wm−1K−1;8,9) thus kT values between 295 K and 1040 K can be calculated using Eq. (35). Note that if k295>1.65 Wm−1K−1 the temperature coefficient (dk/dT) will be negative and if k295<1.65 Wm−1K−1, (dk/dT) will be positive   

( 295-1040 K ) : k T ( W m -1 K -1 ) =1.07+0.00078 ( T-295K ) (35)

A large number of thermal diffusivity (aLP) values for casting slags in the glassy, crystalline and liquid states have been measured16,28,29,30,31) with the LP method; values of kLP were derived using Eq. (36).   

k LP = a LP ./ρ C p (36)

The model determines the thermal conductivity at three temperatures, namely, 298 K, Tcrit (=1040 K) and Tliq; the k-T curves between (298 K and Tcrit) and also (Tcrit and Tliq) are assumed to be linear. There are insufficient data available for the temperature coefficient (dk/dT) of the liquid slag at the present time to permit calculation of values for liquid mould slags.

The super-cooled liquid phase is formed at temperatures above Tg (> ca. 870 K) but the sharp drop in kTHW does not occur until the temperature exceeds the critical (or deformation) temperature, Tcrit. Values of kT between Tcrit and Tliq are calculated using Eq. (37).   

( 1040 K- T liq ) : k T ( W m -1 K -1 ) =1.65-{ ( 1.65- k Tliq ) ( T liq -T ) }/( T liq - 1040 K ) (37)

In partially-crystalline slags (or slag films) the thermal conductivity increases with increasing crystallinity8,28,30,33) since (kcryst≈ 2 kglass). Values of k295 for partially-crystalline slags are calculated from Eq. (38)8,9) using the fcryst derived with Eq. (7). It should be noted that Eq. (38) may produce a slightly low k295 value for a 100% crystalline slag (i.e. k≈ 2 Wm−1K−1) since crystallisation is accompanied by porosity, which lowers the conductivity of the sample. A fully crystalline sample would not be expected to show the collapse in kT above Tcrit exhibited by glasses (Fig. 6); thermal diffusivity values for a fully crystalline sample appear to remain reasonably constant with increasing temperature above 1040 K. Thus, kT has been assumed to remain constant for crystalline samples until the sample melts. However, slag films with a reasonable amount of glassy phase would be expected to show some collapse in kT above Tcrit.   

k 295 ( W m -1 K -1 ) =1.07+0.7  f cryst (38)

When the glass is heated above ca. 1000 K, crystallisation of the sample occurs; note crystallisation occurs roughly in the same temperature range as Tcrit; the model takes no account of any subsequent change in fcrys. Values of kT for partially-crystalline samples in the temperature range (1040 K -Tliq) are calculated using Eq (37).

There are few measurements of the thermal conductivity of liquid mould slags using the kTHW method,8,32,34) but several workers18,28,29,30,31) have reported values for the thermal diffusivity of liquid mould slags (aLP=4×10−7 m2s−1) using the LP method which yields values for kLP which are about 10 x higher than kTHW values (Fig. 6). The kTHW values have been tentatively, adopted. Eq. (39) was obtained from a relation reported for slags covering a wide compositional range.7)   

k m THW ( liq ) =0.139+3.65× 10 -5 exp ( Q/0.3421 ) (39)

Values of k298 have been reported for slag films taken from the mould.9,12) Approximate values for k295 of the slag film can be calculated by using Eq. (40) where fcryst is calculated from Eq. (7). Note k298=1.77 Wm−1K−1 is calculated with Eq. (38) for fcrys=1 and refers to a sample containing pores.   

k 298 ( slag film ) =1.77    f crys +1.07( 1- f crys ) W m -1 K -1 (40)

There are two outstanding problems to be resolved concerning thermal conductivity measurements, namely (i) the differences between kTHW and kLP for temperatures above 1040 K and (ii) the degree of porosity in slag film and partially-crystalline samples which has not yet been determined and which makes it difficult to determine the k value for a fully-dense slag. The calculated thermal conductivity values of the glass and partially-crystalline phases are probably prone to uncertainties of ca. ±5% and ±10%, respectively, for temperatures between 298 and 1040 K. However for temperatures >1040 K, it is difficult to attribute uncertainties until the differences between kTHW and kLP are fully explained.

2.7. Surface Tension (γ) and Interfacial Tension (γmsl)

Surface tension is a surface property and not a bulk property. The magnitude of the surface tension is determined by the composition of the surface. Components with low surface tension (denoted surfactants) tend to occupy the surface layer and hence largely determine the surface tension. Components such as CaO, MgO and Al2O3 have high surface tensions (Table 3) compared with the surface tension of SiO2 and the surfactants (B2O3, K2O, Na2O and CaF2).

Table 3. Values for slag components of surface tension (γ, mNm−1), temperature coefficient (dγ/dT, mNm−2)41) and Gibbs energy of formation for liquid oxides per O atom (−ΔGi)43) used to calculate interaction parameter, φ.42) The last row provides equations used to calculate depression of surface tension (X1γ1) by surfactants.41)
SiO2CaOAl2O3MgONa2OK2OLi2OFeOMnOCaF2ZrO2TiO2B2O3
γ1773260625655635297160[300]645530290400350110
dγ/dT0.031−0.094−0.177−0.13−0.11−0.11−0.11−0.1−0.1−0.07−0.15−0.15
−ΔGi kJmol−1292.2408.2354.6368.4195.2140357158240.5466365304
SurfactantNa2OK2OCaF2B2O3
Equation (Xiγi)surf=−0.8−1388 Xisurf
+6723 (Xisurf)2
=−0.8−1388 Xisurf
+6723 (Xisurf)2
=−2.934 Xisurf
+4769 (Xisurf)2
=−5.2−3.454 Xisurf
+22178(Xisurf)2

The temperature dependence of surface tension (dγ/dT) is usually negative for most components but as the concentration of surfactants increases (dγ/dT) becomes less negative and eventually changes to a positive value.

Few surface tension measurements have been reported for mould slags.17,18,35,36,37,38,39,40) The model is based on a previous model41) in which the following assumptions were made:

(i) The various slag components were divided into, surfactants and bulk components.

(ii) The surface tension of bulk components (Fig. 7(a)) at 1773 K were calculated using partial molar values for the various components (γ=X1γ1+X2γ2+X3γ3) bulk.

Fig. 7.

Schematic diagram showing compositional dependence (X) of surface tension (γ=solid line) and X1 γ1 and X2 γ2 (=dashed lines) for a binary slag system containing (a) no surface active constituents and (b) one bulk constituent (1) and one surfactant (2).41)

(iii) It was found that Xiγi contributions for surfactants resulted in a sharp drop in surface tension until a certain point was achieved (in the range, Xi=0.1 to 0.12) and the relation (up to this point) is expressed as, (X1γ1=a’ +b’X1+c’X12). Above this point, γ increased gradually (Fig. 7(b)) with increasing X (in range, 0.12 to 1.0 and is expressed as (X1γ1=d’+e’X1).

When this model was applied to mould slags (which contain a reasonably high concentration of surfactants) it was found that it tended to produce low values for the surface tension.41)

Consequently, the model was modified and the following assumptions were made:

(i) That the minimum in the surfactant curve in Fig. 7(b) corresponded to the point where the slag surface was saturated with surfactants (Xsurf=0.12).

(ii) That the surface is occupied preferentially by B2O3>K2O>Na2O>CaF2 until the surface becomes saturated with surfactant at ΣXsurf=0.12; (Xiγi)surf is calculated using the values for surfactants given in Table 3.

(iii) Any excess surfactant (left after ΣXsurf=0.12) is treated as a bulk component (γ=X1xsγ1+X2xsγ2).

(iv) The temperature dependence (dγ/dT) was calculated from Eq. (41).   

(dγ/dT)=[ X 1 (d γ 1 /dT)+  X 2 (d γ 2 /dT)] (41)

The surface tension is then calculated from Eqs. (42) and (43)   

γ 1773K = [ ( X 1 γ 1 + X 2 γ 2 + X 3 γ 3 ) conv +( X 1 xs γ 1 + X 2 xs γ 2 )+Σ ( X i γ i ) surf ] (42)
  
γ T = γ 1773K +(dγ/dT)( T-1773 ) (43)

There is a limited amount of data to test the predictions and the reported data show significant variation. Nevertheless, the calculated values are in good agreement with the measured values.1,18,35,36,37,38,39,40) The uncertainty associated with the estimations is probably ca. ±10%.

The interfacial tension (γmsl) can be calculated from Eq. (44) where γm=steel surface tension, γsl can be derived from Eqs. (42) and (43) and the interaction coefficient, φ, can be calculated using Eqs. (45) and (46), where ΔGi is the free energy of formation of the individual, liquid oxides42,43) with the values being given in Table 3.43) Thus, φ is treated as a measure of the oxide components in the slag to supply soluble oxygen in the metal and thereby, decrease the interfacial tension, γmsl. The value of γmsl is largely determined by γm since γm≈ (4 to 5) γsl and the value of γm, in turn, is largely determined by the S content of the steel. The calculated values of γmsl obtained with this procedure were reported to lie within ±100 mNm−1 of the experimental values.42) Values of γmsl are calculated in steel property estimation software, worksheet “surface tension”.   

γ msl = γ m + γ sl -2φ ( γ m γ sl ) 0.5 . (44)
  
φ=0.89-1.5× 10 -3 Δ G i D . (45)
  
Δ G i D .==Δ G FeO- Σ  X i Δ G i . (46)

2.8. Emissivity (εTN)

Total normal emissivity measurements have been reported for mould slags;6) values were recorded for the solid (at T>1030 K) εTN=0.92±0.05 and liquid, εTN=0.91±0.03.

3. Results and Discussion

Routines have been developed in this study to calculate the following properties of mould slags from their chemical composition for subsequent use as input data in mathematical models of the continuous casting process: Tliq, Tg, Tbr; Tcrit; Cp (HT−H298); density and α; viscosity, thermal conductivity, surface and interfacial tension and emissivity.

Although mould slags have relatively low melting points and are benign in their behaviour, there are few measurements available for some properties (e.g. Cp, enthalpy, density and surface tension). By contrast, there are a large number of reported data for viscosity and for thermal conductivity/diffusivity. However, there is high level of experimental uncertainty with viscosity measurements (±10–25%) and there is an unresolved dispute regarding the thermal conductivity values at temperatures at >1040 K derived using the THW and LP methods. It is important that this dispute be settled quickly. One further problem is that many workers fail to report post-measurement analysis of their samples and F and Na losses can be significant during a measurement campaign and affect the property values. This study aims to provide reliable property data for the mould slag for use in mathematical models of continuous casting. Consequently, it is intended to update the existing software as new data become available. The following improvements in the software are planned:

(i) The present software does not include B2O3 additions since various workers report B2O3 causes both increases and decreases in slag viscosity.

(ii) To include calcium-aluminate- based slags used for casting high Al steels.

(iii) To improve the viscosity model for the super-cooled liquid phase.

4. Conclusions

(1) Routines have been developed to calculate the thermo-physical properties of mould slags from their chemical composition for their subsequent use in mathematical models of heat and fluid flow in the continuous casting mould.

(2) The accuracies of estimated values are affected by the limitations in the experimental data.

(3) The cause of the huge discrepancies in thermal conductivities for T>1050 K must be resolved.

Acknowledgements

This work was made possible by the facilities and support provided by the EU (RFSR-PR-10005 DDT), the Research Complex at Harwell, and the EPSRC (EP/I02249X/1).

Symbols Abbreviations

a =Thermal diffusivity (m2s−1)

Cp =Heat capacity (JK−1mol−1 or JK−1kg−1)

fcryst =Fraction crystalline phase

f* =Fraction of powder forming slag

(HT−H298) =Enthalpy relative to 298 K (Jmol−1 or Jkg−1)

k =Thermal conductivity (Wm−1K−1)

S =Entropy (JK−1kg−1)

T =temperature (K)

X =Mole fraction

α =Thermal expansion coefficient(K−1)

γ =Surface tension (mNm−1)

γms =Interfacial tension (mNm−1)

η =Viscosity (dPas)

ρ =Density (kgm−3)

BO =Bridging Oxygens

FO =Free Oxygens

LP =Laser pulse method

NBO =Non-bridging Oxygens

NBO/T =Measure of de-polymerisation

THW =transient hot wire method

Q =measure of polymerisation

scl =super-cooled liquid

Subscripts and Supercripts

m =Value for liquid at Tliq.

References
 
© 2016 by The Iron and Steel Institute of Japan
feedback
Top