Você está na página 1de 20

Chapter 3

Inclusion Formation in Steels


Henri R. Gaye, Research Manager, IRSID

Two main types of deoxidation practices are generally performed depending on the desired products:
Al-killed steels in which the indigenous inclusions are unfloated alumina
clusters formed during the ladle treatment and cooling of the steel between
tundish and mold. A calcium treatment is sometimes made to transform
these clusters into liquid aluminates and thus prevent nozzle clogging, and
in some cases to control the morphology of sulfides; or
semi-killed steels in which a large part of the indigenous inclusions are
formed during metal solidification. In these steel grades, inclusions that will
be easily deformed during hot rolling are generally desired. Depending on the
steel grade, a wide array of compositions can be obtained.
This chapter analyzes the conditions for the formation and chemical stability of oxide inclusions.
The first section summarizes thermochemical data and methods available for evaluating the equi-
librium between liquid metal and precipitates, the second section presents stability diagrams for
the most important systems and the third section discusses various treatments for inclusion con-
trol in different steel grades.

3.1 Selected Thermodynamic Data


3.1.1 Solute Activities in Steels
Throughout this chapter, the activity of solutes is defined with respect to Henrys law and mass
percent of the solute. This reference state is designated as the 1 mass % solution, and the symbol
of the solute is underlined to indicate that it is considered in this reference state.
In Table 3.1 are listed selected values of the free energy of solution of various elements in liquid
iron as a function of temperature, that is, the free energy change for the reaction:
i (pure, stable state, T) = i- (1 mass % solution, T) (Eq. 3.1)
The activity coefficient of solute i in multicomponent melts (fi = ai/%i) is expressed, as a function
of metal composition, with the formalism of interaction coefficients:

Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved. 1
Casting Volume

Table 3.1 Free Energies of Solution of Various Elements in Liquid Iron. From Ref. 1.
i (pure, stable state, T) = -i (1 mass % solution, T)

Element i Gdiss,i (J mol1)

Al (l) 63,178 27.91T


C (graphite) +22,594 42.26T
Ca (l) +121,000 35.5T
Cr (s) +16,736 45.6T
1
/2 H2 (g, 1 atm) +36,484 + 30.46T
Mg (g) 78,690 + 70.8T
Mn (l) +6,067 38.16T
1
/2 N2 (g, 1 atm) +3,598 + 23.89T
Ni (l) 7,531 38.7T
1
/2 O2 (g, 1 atm) 117,152 2.89T
1
/2 P2 (g, 1 atm) 122,173 19.25T
1
/2 S2 (g, 1 atm) 135,060 + 23.43T
Si (l) 131,378 15.06T
Ti (s) 25,104 44.98T

log10 f i = eij % j + rij,k % j%k (Eq. 3.2)


i j, k

where j and k represent the various solutes, and eij and rij,k are first- and second-order interaction
coefficients that describe the effect of solutes j and k on the activity of i. Selected values of first
order interaction coefficients at 1600C are given in Table 3.2. The impact of second-order terms
is in general negligible, except in the case of some high-alloy steels (e.g., N in stainless steels at
high Cr and Ni contents) and maybe for some highly reactive elements (see Table 3.6).
Computed values of activities and activity coefficients for typical compositions of liquid iron and
steels are given in Table 3.3. For low-alloy carbon steels, the activity coefficients of most solutes
differ little from 1, in contrast to stainless steels, for which very large deviations from unity are
observed for most solutes.

3.1.2 Solubility of Gases in Iron


Diatomic gases such as O2, N2 and H2 dissolve in liquid or solid iron alloys in the atomic form :
1
/2 X2 (g) = X (Eq. 3.3)
The activity of the dissolved element is directly proportional to the square root of the equilibrium
gas partial pressure (Sieverts law). The temperature dependence of the equilibrium constants for
H2, N2 and O2 solubilities in liquid and solid iron alloys, from the compiled data cited in Ref. 2,
are given in Table 3.4. In this table is also given the expression of the oxygen saturation as a func-
tion of temperature, that is, the dissolved oxygen content in the metal at equilibrium with the pure
stable iron oxide (liquid oxide, wstite or magnetite, depending on temperature).

3.1.3 Solubility Products of Oxides, Sulfides and Nitrides in Liquid


Iron
The solubility product of a compound XnYm is the equilibrium constant for the reaction between
the pure compound and the elements dissolved in liquid iron taken in the reference state 1 mass %
solution:

2 Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.
Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.

Table 3.2 Selected Values of Interaction Coefficients e j at 1600C Between Elements Dissolved in Liquid Iron. From Refs. 13.
i
j Al C Ca Cr H Mg Mn N Ni O P S Si Ti
i

Al 0.043 0.091 0.047 0.03 0.24 0.058 1.98 0.03 0.056

C 0.043 0.14 0.098 0.024 0.67 0.012 0.11 0.012 0.21 0.051 0.046 0.08

Ca 0.072 0.337 18 0.02 0.049 0.097

Cr 0.054 0.12 0.025 0.0003 0.33 0.017 0.19 0.0002 0.33 0.053 0.02 0.0043 0.059

H 0.013 0.06 0.0022 0.0 0.0014 0.0 0.19 0.011 0.008 0.027 0.019

Mg 0.01 0.012

Mn 0.07 0.31 0.003 0.091 0.10 0.0035 0.048 0.06

N 0.028 0.13 0.047 0.02 0.0 0.01 0.05 0.045 0.007 0.047 0.53

Ni 0.042 0.074 0.0003 0.24 0.036 0.028 0.0009 0.01 0.0035 0.0037 0.0057

O 1.17 0.44 0.037 3.1 0.03 0.057 0.006 0.0 0.07 0.133 0.14 0.37

P 0.13 0.03 0.21 0.0 0.094 0.0002 0.13 0.062 0.028 0.12

Inclusion Formation in Steels


S 0.035 0.113 0.011 0.12 0.026 0.01 0.0 0.27 0.029 0.028 0.063 0.072

Si 0.058 0.18 0.067 0.0003 0.64 0.033 0.09 0.005 0.25 0.11 0.056 0.11

Ti 0.055 1.1 0.04 1.8 1.1 0.11 0.056


3
Casting Volume
4

Table 3.3 Typical Values of Activities and Activity Coefficients in Liquid Iron and Steels (ai = fi %i 1 mass % solution).

Metal Al C Mn P S Si Ti H N O Cr Ni
Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.

Liquid %i 4.85 0.45 0.1 0.03 0.3


iron fi 5.8 0.3 9.8 6.0 10.5
1500C ai 28 0.135 0.98 0.18 3.15

Carbon %i 0.05 0.45 0.02 0.01 0.3 0.05


steel fi 1.05 1.06 1.0 1.1 1.0 1.15 0.93 1.0 0.97 0.85
1600C ai 0.053 0.45 0.022 0.01 0.345 0.046

Stainless %i 0.05 0.45 0.02 0.01 0.3 0.05 18 8


steel fi 3.6 0.49 1.0 0.32 0.66 1.24 9.4 0.93 0.17 0.21 0.97 1.0
1600C ai 0.025 0.45 0.0064 0.0066 0.372 0.47 17.5 8.0

Table 3.4 Solubility of Gases in Liquid and Solid Iron. From Ref. 2

log10 (a H /PH2 / ) log10 (aN /PN2 / ) log10 (aO /PO2 / )


1 1 1
Phase 2 2 2
log10 (%Osaturation)

Fe (l) 1820 / T 1.63 285 / T 1.21 6119 / T + 0.15 6320 / T + 2.73

Fe 1376 / T 2.405 1330 / T 1.180 8156 / T 2.26 3830 / T + 0.03

Fe i 1411 / T 2.208 420 / T 1.932 9106 / T 3.11 4410 / T + 0.107

Wstite ( > 570C) :


4515 / T + 0.47
Fe  1376 / T 2.405 1580 / T 1.009 9220 / T 2.85
Magnetite ( < 570C) :
5150 / T + 1.2
Inclusion Formation in Steels

XnYm = n X + m Y (X = Al, C, ...Y = O, S, N, ...), (Eq. 3.4)


that is,
n m
K = aX
aY
(Eq. 3.5)
Selected values of solubility products of oxides, sulfides and nitrides in liquid iron are listed in
Table 3.5.2,3 Fig. 3.1 indicates the relationship between oxygen activity and the activity of various
elements at equilibrium with the pure oxide at 1600C, and Fig. 3.2 shows the relationship between
Oxygen activity (1 mass % solution)

Activity of element (1 mass% solution)

Fig. 3.1 Activities of oxygen and various elements dissolved in liquid steel, at equilibrium with the pure oxide at 1600C.
Sulfur or Nitrogen activity (1 mass % solution)

Activity of element (1 mass % solution)

Fig. 3.2 Activities of sulfur (nitrogen) and various elements dissolved in liquid steel, at equilibrium with the pure sulfide
(nitride) at 1600C.

Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved. 5
Casting Volume
6

Table 3.5 Selected Values of Solubility Products of Oxides, Sulfides and Nitrides in Liquid Iron. From Refs. 2 and 3.

Element Oxide log10 K K1600C Sufide log10 K K1600C Nitride log10 K K1600C
Al FeAl2O4 70320/T+23.38 6.9 1015 Al2S3 103 AlN 12950/T+5.58 4.6 102
Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.

Al2O3 62680/T+20.54 1.2 1013

B B2O3 1.5 108 BN 10000/T+4.64 0.2

C CO (gas) 1168/T2.07 2.05 103

Ca CaO 9 107 CaS 1.7 105 Ca3N2 1.2 1011

Ce Ce2O3 68500/T+19.6 1017 CeS 20600/T+6.39 2.5 105

Cr FeCr2O4 50700/T+21.7 4 106


Cr2O3 40740/T+17.78 1.1 104

La La2O3 62050/T+14.1 9.3 1020 LaS 26000/T+8.98 1.3 105

Mg MgO 3 107 MgS 9 104

Mn (FeMn)Oliq 12760/T+5.57 5.8 102 MnS 2.7


(FeMn)Osol 15050/T+6.7 4.7 102

Nb FeNb2O6 88300/T+36.76 4.1 1011 NbN 11100/T+5.4 0.3


NbO2 32780/T+13.92 2.6 104

Si SiO2 31040/T+12.0 2.7 105

Ta FeTa2O6 79300/T+28.43 1.2 1014 TaN 15410/T+7.8 0.4


Ta2O5 63100/T+21.9 1.6 1012

Ti TiO2 5 107 TiS 8000/T+4.02 0.56 TiN 19755/T+7.78 1.7 103


Ti3O5 3.5 1018
Ti2O3 2.7 1011

U UO2 5.9 1011 US 1.6 104

V FeV2O4 8.3 108


V2O3 3.5 106
VO 15530/T+6.66 2.3 102

Zr ZrO2 40750/T+11.8 1.1 1010 ZrS 0.3 ZrN 13330/T+4.8 4.8 103
Inclusion Formation in Steels

Table 3.6 Solubility Products of CaO and MgO and Interaction Coefficients Between Ca and O
and Mg and O at 1600C. From Ref. 4

log10 KCa = 33696 / T + 7.77 eOCa = 3600 * / 990 ** rOCa = 5.7 105 * / 4.2 104 **

(K1600C = 6 1011) eOCa= 9000 * / 2500 ** rOCa,O = 2.9 106 * / 2.1 105 **
O,Ca
rCa = 2.9 106 * / 2.1 105 **
O
rCa = 3.6 106 * / 2.6 105 **

log10 KMg = 38063 / T + 12.45 eOMg = 300 +/ 10 rOMg = (1.6 +/ 0.04) 104

(K1600C = 1.3 108) O


eMg = 460 +/ 20 rOMg,O = (4.8 +/ 0.2) 104
O,Mg
rMg = (4.8 +/ 0.2) 104
O
rMg = (3.7 +/ 0.1) 104

* % Ca + 2.51% O < 0.005


** % Ca + 2.51% O = 0.005 to 0.018

sulfur or nitrogen activity and the activity of various elements at equilibrium with the pure sulfide
or nitride at 1600C.
The direct experimental values selected in Table 3.5 for the solubility products of CaO and MgO3
are several orders of magnitude larger than the thermodynamic value calculated from the free ener-
gies of reaction of the pure elements and free energy of solution of elements at infinite dilution in
both Ca or Mg and O. This discrepancy has been explained by very accurate measurements on
deoxidation equilibria in liquid iron,4 which show a very sharp variation of the activities of Ca, O
and Mg in the relevant metal composition range. Several regions are observed for the value of
apparent solubility product as the content in (%Ca + 2.51% O) is increased, with first a very sharp
increase for low values of this parameter, and then a leveling off at higher values. The proposed
coherent sets of first- and second-order interaction coefficients between Ca and O and between Mg
and O are listed in Table 3.6. Calculations made with the values selected in Table 3.5 are in rea-
sonable agreement with the values in the domain %Ca + 2.51% O = 0.005 to 0.018.

3.1.4 Slag Component Activities


Recent compilations of slag component activities are available.5,6 Slag models are, however, desir-
able in order to perform calculations in multicomponent systems necessary to represent industrial
applications. Among those, one of the most reliable is the cell model developed at IRSID.7-8 It is
presently used for calculations in the system SiO2-TiO2-Ti2O3-Cr2O3-Al2O3-Fe2O3-CrO-FeO-
MgO-MnO-CaO-CaF2-S.

3.1.5 Computational Thermodynamics Software/Database


Packages
The most widely used commercially available packages, Thermo-Calc,9 FACT,10 ChemSage,11
MTDATA12 and GEMINI213 combine computation codes for multiphase, multicomponent equilib-
ria calculation and databases and have found extensive applications in the metallurgical, materials
and chemical engineering industries. They can be applied for understanding inclusion stability in
steels, although some applications may require specific developments.

Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved. 7
Casting Volume

At Irsid, a package called CEQCSI,14 based in a large part on the Irsid slag model, has been devel-
oped for very specific applications in the field of iron and steelmaking. It is well adapted for the
calculation of transfer of elements between slag and metal, and for the calculation, from the over-
all steel composition, of the sequence of inclusion precipitation at equilibrium from liquid steel
treatment to subsequent temperature evolution during cooling and solidification of the steel. In this
last situation, an original method has been developed15 in which the microsegregation equations for
elements dissolved in liquid metal, with or without diffusion in solid metal, and the equilibrium
conditions between liquid steel and oxide, sulfide, nitride or carbide precipitates are treated simul-
taneously. The calculation also gives
the liquidus temperature of the oxide
inclusions and their crystallization path
during cooling.

3.2 Stability Domains Si activity (%)

of Oxide Inclusions in
Various Systems
These diagrams represent, as a function
of solute contents in liquid steel, the
nature of the stable oxide and the oxy-
gen activity in equilibrium with this
oxide. The metal composition has been
expressed with the solute activities, so
that the diagrams can be used regard-
less of the steel grade (carbon or stain-
less steels). They have been computed Al activity (%)
using the Irsid slag model.
Fig. 3.3 Domains of stability of oxides in the system Fe-Al-Si-Mn-O
at 1600C.
3.2.1 The Fe-Al-Si-Mn-O
System
This system is the basis for the under-
standing of oxide inclusion stability in
most Al-killed and semi-killed carbon
or stainless steels.
Si activity (%)

Fig 3.3 indicates the domains of stabil-


ity of oxides at 1600C, for three Mn
activities (0, 0.5 and 1.0). The stable
phases in the system without Mn are
solid oxides, Al2O3, mullite (3Al2O3
2SiO2) and SiO2. As the Mn content is
increased, a domain of stability of liq-
uid oxides replaces progressively the
stability domain of mullite, and a
domain of stability of galaxite (MnO
Al2O3) appears at low Si contents.
Al activity (%)
Fig. 3.4 indicates, for each one of the
three sections, the equilibrium oxygen Fig 3.5 Domains of stability of oxides in the system Fe-Al-Si-Mn-O at
activity, and Fig. 3.5 and Fig. 3.6 give 1480C.

8 Cop
right 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.
Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.

Si activity (%)

Si activity (%)

Si activity (%)
Al activity (%)
Al activity (%) Al activity (%)

(a) (b) (c)


Fig. 3.4 Oxygen saturation and domains of stability of oxides in the system Fe-Al-Si-Mn-O at 1600C: (a) section aMn = 0; (b) section aMn = 0.5; (c) section aMn = 1.0.
Si activity (%)

Si activity (%)
Si activity (%)

Inclusion Formation in Steels


Al activity (%) Al activity (%) Al activity (%)

(a) (b) (c)


9

Fig. 3.6 Oxygen saturation and domains of stability of oxides in the system Fe-Al-Si-Mn-O at 1480C: (a) section aMn = 0; (b) section aMn = 0.5; (c) section aMn = 1.0.
Casting Volume

Ti activity (%)

0.1

Ti2O3
IF Steels

0.01 Liquid
Al2O3-TiO2 TiOx Steels

Al2O3

O = 80 40 20 10 ppm
0.001
0.0001 0.001 0.01
Al activity (%)
Fig. 3.7 Oxygen saturation and domains of stability of oxides in the system Fe-Al-Ti-O at 1600C.

Ti activity (%)

0,1
Ti2O3

Al2O3-TiO2
0,01
Liquid Al2O3

O = 40 20 10 5 ppm
0,001
0.0001 0.001 0.01
Al activity (%)
Fig. 3.8 Oxygen saturation and domains of stability of oxides in the system Fe-Al-Ti-O at 1520C.

10 Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.
Inclusion Formation in Steels

the same information at 1480C, which is a typical steel liquidus temperature for high carbon
grades. From these diagrams, it is possible to estimate the amount of oxide inclusions that can pre-
cipitate during cooling of the steel between ladle and mold, and during steel solidification.
As shown in some of the examples treated in Section 3.3 of this chapter, some trace elements (Ca,
Mg) can also enter the oxide composition. However, the exact effect of these trace elements is eas-
ier to apprehend on a case-by-case basis rather than on general diagrams.

3.2.2 The Fe-Ti-Al-O System


This system is of interest for IF steels and Ti-deoxidized steels, developed under the name of Met-
allurgy of oxides in steels.16,17 It has indeed been shown that inclusions containing both Al2O3 and
TiO2 formed during solidification of steels for tubes or plates, and on which manganese sulfides
can precipitate, can serve as nucleation sites for acicular ferrite after welding or heat treatment.
The diagrams of Figs. 3.7 and 3.8 are the computed equilibrium diagrams for the system Fe-Ti-Al-
O at 1600C and at 1520C (liquidus temperature for the considered steel grades). In traditional
Al-killed IF steels stabilized with Ti, the equilibrium deoxidation product is alumina; the Ti-con-
taining oxides commonly found in SEN deposits are presumably the result of reoxidations in the
nozzle. In Ti-deoxidized steels, the small amount of dissolved oxygen will form the desirable
inclusions during steel solidification. It is apparent from Fig. 3.7 that very low Al contents (below
about 40 to 50 ppm) have to be reached in order to obtain indigenous oxide inclusions containing
titanium oxides in steels containing about 100 to 200 ppm Ti.

3.3 Treatments for Inclusion Control in Steels


3.3.1 Inclusions in Al-killed Steels
Indigenous alumina inclusions found in Al-killed steels are mostly residual deoxidation products
in the form of large clusters. The amount precipitated during steel cooling and solidification is
rather small, on account of the very small solubility product of Al2O3. As shown in Fig. 3.9, the
elementary alumina particles have a morphology that depends on their formation conditions (Al-
O supersaturation at the time of precipitation) and residence time in liquid metal at high tempera-
ture.18,19 In addition, for treatments conducted with a well-deoxidized basic ladle slag, transfer of

Fig. 3.9 Schematic representation of the morphology of Al2O3 during deoxidation and ladle treatment (from Ref. 19): 1.
Nuclei form small particles.; 2. Nuclei grow to dendrites.; 3a. Denditric growth continues in RH-OB because of high reaction
speed by large supersaturations of Al and O.; 3b. Agglomeration of dendrites gives clusters. At some dendrites spheres
have been formed. Big clusters float out of the steel.; 4. At the stirring station spheres form on the dendrite tips because of
a lower reaction speed by low supersaturation of Al or O. Agglomerates form and float out of the steel.; 5. The remaining
small clusters become compact and the particles become irregular.; 6. Small finger-like dendrites form in RH-OB when
adding aluminum in already Al-killed steel.

Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved. 11
Casting Volume

K CaS = 1.7 10-5 K CaO = 9 10-7

0.01
Ca activity (%)

0.001

0.0001

S activity (%) Al activity (%)

Fig. 3.10 Equilibrium diagram for the system Fe-Al-Ca-O-S at 1600C.

Mg and Ca from the slag may result in the transformation of alumina inclusions into spinels and
calcium aluminates.
Late reoxidations first generate large liquid manganese silico-aluminates that are thermodynami-
cally unstable. The final product, after reduction by Al from the steel, is a globular, Al2O3-saturated
inclusion. In Ti-containing IF steel grades, reoxidations can generate, after reduction of the transi-
tory TiOx-containing phase, alumina inclusions in the shape of veils that can aggravate nozzle clog-
ging. Similarly, mold slag entrapments result in the formation of complex inclusions partially
reduced by Al, easily recognizable by the presence of certain tracers (Al, K).

3.3.2 Calcium Treatment of Al-killed Steels


Calcium additions are widely used in steelmaking to transform alumina clusters into liquid calcium
aluminates, in order to prevent nozzle clogging during continuous casting, and to control the mor-
phology and composition of sulfide precipitates that remain in the solidified steel. During these
treatments, oxygen and sulfur in the steel compete to react with calcium. The stability diagram for
the system Fe-Al-Ca-O-S (Fig. 3.10) indicates that the latitude to form liquid aluminates is
strongly a function of steel S content. For instance, in a steel with 0.05 Al, calcium aluminates con-
taining 50% CaO can be obtained without CaS formation in a steel containing 0.006% S, whereas
the CaO content is limited to 40% in a steel containing 0.035% CaO.
Calcium is added to the metal as calcium alloys (Si-Ca, Si-Ca-Ba, ...) that can be injected pneu-
matically, or better by cored wire feeding, or added by immersion of small ingots (Ca-Fe log). Dur-
ing the treatment, only a small part (about 15 to 20%) of added calcium remains in the metal, the
rest being lost by vaporization. The calcium remaining in the metal at treatment temperature con-
sists of:
Ca contained in oxide inclusions. If this amount is too small, solid alumi-
nates are formed (CaO6Al2O3, CaO2Al2O3, CaOAl2O3), which are even
more detrimental than alumina inclusions with regard to nozzle clogging;20
Ca dissolved in liquid metal, this amount being larger at higher Al content.
Most of this calcium will form precipitates of CaS or (Ca,Mn)S during steel
solidification;

12 Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.
Inclusion Formation in Steels

Ca total (ppm)

O total (ppm)

(a)
Ca total (ppm)

O total (ppm)

(b)

Fig. 3.11 Nature of oxides formed and S content for CaS precipitation during Ca treatment, as a function of total Ca and
O contents: (a) steel Al content = 150 ppm; (b) steel Al content = 650 ppm.

Ca possibly contained in calcium sulfide in the liquid steel if the CaS satu-
ration limit is reached. In this case, the CaS precipitates formed in liquid
metal can also contribute to nozzle clogging, and the amount of residual cal-
cium may be too small to form liquid aluminates.
Fig. 3.11 summarizes, for a steel composition 0.8% C0.65% Mn0.2% Si0.13% Cr and Al con-
tents of 150 and 650 ppm, the calculated distribution of calcium among these various phases, as a
function of total Ca and O contents fixed in the metal at the end of the calcium treatment. On these
diagrams, the solid lines represent the limits of existence of the various oxides; that is, for increas-
ing calcium content at a given oxygen content, solid aluminates are in the domain labeled Solid

Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved. 13
Casting Volume

Fig. 3.12 Distribution of elements in a composite oxide (center) / sulfide (outer part) inclusion in a Ca-treated resulfurized
steel. From Refs. 21 and 22.

Oxides, mixture of CaOAl2O3 and liquid oxides (from 0 to 100% liquid) are in the domain
labeled CA + liq, and liquid oxides of increasing CaO content are in the domain labeled Liquid
Oxides. The dotted lines indexed with S contents indicate the calcium content for which CaS sat-
uration is reached for the given S content. In the domains CA + liq, the S content for CaS satu-
ration is practically constant, 90 ppm S for 650 ppm Al and 450 ppm S for 150 ppm Al.
These diagrams indicate the windows available for a proper transformation of alumina inclusions
into fully liquid aluminates without CaS precipitation in liquid steel. It is apparent that for the steel
grade with 650 ppm Al, this window is completely closed for an S content as low as 90 ppm,
whereas for the steel grade with 150 ppm Al, an adequate transformation of oxides can be obtained
without CaS precipitation for much higher S contents.
In Ca-treated resulfurized steels (for instance, free machining steel grades), part of the MnS pre-
cipitates on oxide inclusions, in the latest stages of solidification. Further transformation takes
place with transfer of Ca from the oxide core to the MnS and results in a composite inclusion with
globular Al2O3-rich core and outer layer of (Ca,Mn)S, as shown in Fig. 3.12.21,22

Fig. 3.13 Various origins of the oxide inclusions present in a semi-killed steel grade. From Ref. 23. Steel composition: 0.7%
C1% Mn0.35% Si8 ppm Al3 ppm Ca0.4 ppm Mg16 ppm O.

14 Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.
Inclusion Formation in Steels

3.3.3 Inclusions in Semi-killed Steels


Indigenous inclusions in semi-killed steels can have various origins, as shown in Fig 3.13, which
summarizes the results of a thorough investigation of inclusions in as-cast samples of such a steel
grade.23 The micro-analyses of about 50 inclusions were coherent with a calculation of the
sequence of equilibrium precipitation of the inclusions made with the CEQCSI software. They can
be classified into three categories:
Unfloated deoxidation products amount to about 5 ppm of oxygen out of the
total oxygen content of 16 ppm. They are composed essentially of SiO2,
Al2O3 and CaO and contain small amounts of MgO and MnO. They were
liquid when precipitated and, consequently, have spherical shapes in the as-
cast product. During cooling, a (Mg-Mn)O-Al2O3 spinel phase has crystal-
lized in some of them, the matrix remaining homogenous and glassy. Once
these inclusions have precipitated, the Ca and Mg contents remaining in
solution in the liquid steel are practically zero, and the inclusions formed
later on have a composition in the SiO2-MnO-Al2O3 system. They represent
nearly 70% of the residual inclusions and are globally distributed among
two large classes:
Inclusions formed during steel solidification. The ones precipitated at the
beginning of solidification are located in the alumina primary phase field,
and alumina crystals were formed in some of them during cooling. During
the later stages of solidification, as a result of segregations, their SiO2 con-
tents increase, and inclusions close to silica saturation may be formed.
Inclusions of manganese silicate with traces of Al2O3. From these originally
liquid inclusions, a large amount of silica crystals precipitate during cooling,
and the matrix has a composition close to the 39% SiO261% MnO eutectic.
The glassy matrix of these inclusions is usually easily deformed and elongated during hot rolling
(although its partial crystallization during the heat treatments prior to hot rolling can in some cases
create undeformable phases),22 whereas the hard crystallized phases (spinels and alumina) are not
deformed and can create defects in the final product. The potentially most harmful inclusions in
this respect are the residual deoxidation inclusions and the inclusions formed at the beginning of
solidification. One of the ways to avoid hard crystallized phases in those inclusions is to limit the
Al content of liquid metal to even lower values. Thus, a decrease of total Al content from 8 to 5
ppm would result in a large decrease of the Al2O3 content of the inclusions (from 25 to 15% for
the deoxidation products, and from 35 to 28% for the inclusions formed at the beginning of solid-
ification). This would also result in a switch from spinel and alumina primary phase fields to
melilite (solid solution of 2CaOAl2O3SiO2 and 2CaOMgO2SiO2) and spessartite
(3MnOAl2O33SiO2) crystallization domains.

3.3.4 Inclusion Control in Semi-killed Steels by Slag Treatment


The only practical method for decreasing the liquid metal contents in trace elements (Al, Ca, Mg)
below the levels that would result in the precipitation of deleterious inclusions consists in equili-
brating the metal with a ladle slag of adequate composition. In order not to exceed an aluminum
content of 5 ppm, while keeping the O content as low as possible, it is necessary to aim for a CaO-
Al2O3-SiO2 slag composition corresponding to a SiO2/CaO ratio of about 0.9 with Al2O3 contents
that do not exceed 10% (Fig. 3.14). Note that the Al2O3 content of the slag is much smaller than
the content that can be tolerated in the inclusions.
The diagrams in Fig. 3.15 show more accurately the residual aluminum contents at equilibrium
with industrial slags containing MgO, MnO and CaF2, in addition to CaO, Al2O3 and SiO2.22-24 On
this figure, it may be observed that an increase in slag basicity or MgO content of the lining greatly

Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved. 15
Casting Volume

Fig. 3.14 Principle of a ladle slag treatment of a semi-killed steel grade in order to limit the steel content in dissolved Al
and O.

increases the residual aluminum content. An increase in treatment temperature has a relatively
smaller effect on the aluminum content but results in a higher dissolved oxygen content and there-
fore in a higher inclusion content. It is therefore advisable to perform this ladle stirring at the low-
est temperature compatible with adequate casting conditions. These diagrams also show that a
decrease in slag basicity, either by an increase of the SiO2/CaO ratio or by a decrease in the MgO
content, leads to a perceptible increase of the MnO content at equilibrium.
Calcium and magnesium transfer from the slag to the metal can also strongly affect metal quality
when the basicity of the slag or its MgO content are too high. The risk is particularly acute for treat-
ments performed in ladles with magnesia or dolomite lining, or treatments made in ladles polluted
by ladle glazes from previous treatments of killed steels. The deleterious inclusions then formed
can contain calcium aluminates and spinels, and even 2CaO-SiO2 or 2MgO-SiO2 when the Al con-
tent is very low.

3.3.5 Inclusions in Free-Cutting Steels


In these steel grades with high S contents, two main objectives are sought:
obtain a good distribution of MnS precipitates to limit the chips length,
avoid the contact of the tool with hard abrasive oxides (Al2O3, spinels ...) by
avoiding their formation or embedding them in sulfides.
In addition, for high-speed machining, it is favorable to harden some of the sulfides that will then
form protective films on the tool. One way to do this is to add calcium to form (Ca,Mn)S; another
way is to obtain oxisulfides in which the hardening effect is created by the partial substitution of
S by O, as illustrated in Fig. 3.16.25 This last solution is applied for free-cutting steels.
Calculations, based on a thermodynamic description of the oxisulfide phase,26 were made to predict
the nature of oxide, oxisulfide and sulfide phases precipitating in liquid steel and during solidifi-
cation. This was in a base steel of composition 0.07% C1.35% Mn0.014% Si containing minute
amounts of tramp elements Al, Ca and Mg resulting from the ladle treatment, as a function of oxy-
gen content (20 to 120 ppm). The calculated sequence of inclusion precipitation for an oxygen con-
tent of 60 ppm is indicated in Fig. 3.17 (oxygen distribution in the various precipitates) and Fig.
3.18 (sulfur distribution in the various precipitates).
A summary of the effect of oxygen content on the amount of the various precipitates is presented
in Figs. 3.19 and 3.20. It appears that the optimal oxygen content is around 60 ppm, as the amount

16 Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.
Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.

Al (ppm)

Al (ppm)
% MnO
Al (ppm)

% MnO
% MnO
% Al2O3 % MgO
% SiO2 / % CaO

(a) (b) (c)

Fig. 3.15 Effect of treatment conditions on


equilibrium Al content and slag MnO con-
tent. Steel composition = 0.7% C0.6%
Mn0.25% Si: (a) effect of slag SiO2/CaO
ratio at 1550C, %Al2O3 = 10, %MgO = 6,
%CaF2 = 0; (b) effect of slag %Al2O3 content
Al (ppm)

% MnO
Al (ppm)

% MnO

at 1550C, CaO/SiO2 = 0.9, %MgO = 6,


%CaF2 = 0; (c) effect of slag MgO content at
1550C, CaO/SiO2 = 0.9, %Al2O3 = 10,

Inclusion Formation in Steels


%CaF2 = 0; (d) effect of slag CaF2 content at
1550C, CaO/SiO2 = 0.9, %Al2O3 = 10,
%MgO = 6; (e) effect of treatment tempera-
ture, CaO/SiO2 = 0.9, %Al2O3 = 10, %MgO =
6, %CaF2 = 0.
% CaF2 Temperature C

(d) (e)
17
Casting Volume

Fig. 3.16 Effect of sulfide composition on their plasticity at hot rolling temperatures. From Ref. 25.

Metal liquid fraction


O (ppm)

Temperature C

Fig. 3.17 Sequence of precipitation of oxide and oxisulfide inclusions during steel solidification, and amount of oxygen fixed
in these phases as a function of temperature, in a free-cutting steel grade.

18 Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.
Inclusion Formation in Steels

Metal liquid fraction


S (ppm)

Temperature C

Fig. 3.18 Sequence of precipitation of oxisulfide and sulfide inclusions during steel solidification, and amount of sulfur fixed
in these phases as a function of temperature, in a free-cutting steel grade.
O (ppm)

O total
Fig. 3.19 Distribution of oxygen among oxisulfides, liquid and solid oxides as a function of total oxygen content in a free-
cutting steel grade.

of oxisulfides is large and the amount of harmful oxides is minimized. This oxygen range has
indeed shown the best free-cutting properties in industrial casts.

References
1. A. Rist, M.F. Ancey-Moret, C. Gatellier and P.V. Riboud, Equilibres thermodynami-ques
dans llaboration de la fonte et de lacier, Techniques de lIngnieur, Form. M 1733 (1974).
2. C. Gatellier and H. Gaye, La Revue de Mtallurgie - CIT (Jan. 1986), pp. 2542.
3. M. Nadif and C. Gatellier, La Revue de Mtallurgie - CIT (May 1986), pp. 377394.
4. H. Ohta and H. Suito, Metall. Mater. Trans. B, 28B (1997), pp. 11311139.
5. Slag Atlas, 2nd edition, edited by VDEh, Verlag Stahleisen GmbH, D-Dsseldorf (1995).
6. The Making, Shaping and Treating of Steel, 11th Edition, Steelmaking and Refining Volume
(1998), edited by R.J. Fruehan, AISE Steel Foundation, chapter 2.

Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved. 19
Casting Volume

S (ppm)

O total
Fig. 3.20 Distribution of sulfur among sulfides, oxisulfides and oxides as a function of total oxygen content in a free-cutting
steel grade.

7. H. Gaye, J. Welfringer, Proc. 2nd International Symposium on Metallurgical Slags and


Fluxes (1984), edited by H.A. Fine and D.R. Gaskell, Warrendale, Pa., pp. 357375.
8. J. Lehmann and H. Gaye, 82nd Steelmaking Conference Proceedings (1999), pp. 463470.
9. B. Sundman, B. Jansson and J.-O. Andersson, Calphad 9 (1985), pp. 153190.
10. C.W. Bale, A.D. Pelton and W.T. Thompson, F*A*C*T, CRCT (1995), Ecole Polytechnique,
Montreal, Canada.
11. G. Eriksson and K. Hack, Metall. Trans. B, 21B (1990), pp. 10131023.
12. A.T. Dinsdale, S.M. Hodson, T.I. Barry and J.R. Taylor, in Computer Software in Chemical
and Extractive Metallurgy (1989), Pergamon Press, New York, pp. 5974.
13. B. Cheynet, in Computer Software in Chemical and Extractive Metallurgy (1989), Pergamon
Press, New York, pp. 3144.
14. C. Gatellier, H. Gaye, J. Lehmann and Y. Zbaczyniak, Revue de Metallurgie - CIT (Oct.
1992), pp. 887888.
15. M. Wintz, M. Bobadilla, J. Lehmann and H. Gaye, ISIJ Int. 35 (1995), pp. 715722.
16. J.-I. Takamura and S. Mizoguchi, Proc. of the 6th Iron and Steel Congress, ISIJ, 1 (1990),
pp. 591597.
17. S. Ogibayashi, K. Yamaguchi, M. Hirai, H. Goto, H. Yamaguchi and K. Tanaka, Proc. of the
6th Iron and Steel Congress, ISIJ, 1 (1990), pp. 612617.
18. M. Akiyoshi, Camp ISIJ, 4 (1991), p. 1235.
19. W. Tieking, J. Brockhoff and J. Van der Stel, Proc. 4th Intern. Conf. on Clean Steel (810
June 1992), Balatonszplak, Hungary, pp. 704717.
20. G.M. Faulring, J.W. Farrell and D.C. Hilty, Iron & Steelmaker, 7 (1980), pp. 1420.
21. C. Gatellier, H. Gaye, J. Lehmann, J.N. Pontoire and P.V. Riboud, Steel Research, 64 (1993),
pp. 8792.
22. H. Gaye, C. Gatellier and P.V. Riboud, Proceedings of the Ethem T. Turkdogan Symposium,
Iron & Steel Society, 1994, pp. 113124.
23. C. Gatellier, H. Gaye, J. Lehmann, J. Bellot and M. Moncel, Proc. 4th Intern. Conf. on Clean
Steel (810 June 1992), Balatonszplak, Hungary, pp. 638651.
24. F. Stouvenot, H. Gaye, C. Gatellier and J. Lehmann, Proc. 52nd Electric Furnace Confer-
ence (Nov. 1994), Nashville, pp. 423428
25. L. Luyckx, J.R. Bell, A. McLean and M. Korchynsky, Metall. Trans., 1 (1970), pp.
33413350.
26. H. Gaye and J. Lehmann, Proc. of the 5th International Conference on Molten Slags, Fluxes
and Salts (1997), Sydney, Australia, Iron & Steel Society, pp. 2734.

20 Copyright 2003, The AISE Steel Foundation, Pittsburgh, PA. All rights reserved.

Você também pode gostar