You are on page 1of 7
Basicity of metallurgical slags* By 1D Sommerville! and Yindong Yang’ ABSTRACT Despite the imponance and usefulness of basicity in metallurgical sags, the optimum quantitative expression of i has mained eusve arly doe to the fact thatthe precise concept of basicity a not bee propely clarified. The two basic approaches 10 the concept of bisety are the Lux-Flood, which perceives basicity in terms of the activity ofthe fee ‘oxide fon, 4,» andthe Lewis, which perceives basiciy in terms of the ability to doBate negative charge: The former has been anes! universally fdopied in the area of malurgial slags wile te Tater has boon ‘creasingly accepted inthe fil of glass chemist, this paper, some dificlies ofthe Lux-Flod approach are pointd out, inluing the need for inducct measures ofa, which have proliferated in recent year. Accordingly, is suggest thatthe Lewis approach has good deal to commend it and that the quantification of In terms of optics basicity ts the beet measite of Bascty availble, which should therefore be more widely applied in metallurgical slags than it curently is The fact ha thre ae ntinsic problems in evaluating the optieal ase forthe anstion metal aides acknowledged, and ‘new sees recommended. The pattems of iso-opticl asic Tine in Ternary systems are shown 1 spre well withthe patern of i-acvity and iso-capacity ines under appropiate conditions INTRODUCTION Despite its undoubted importance in pyrometallurgical, processing, the concept of basicity of the slags and fluxes employed has remained somewhat ambiguous and ill-defined. AS will be reviewed briefly, a large number of ways of expressing basicity have been suggested over the years, but no completely satisfactory definition has emerged ‘The most common approach is sil the use of a basicity ratio for basicity index in which the basic oxides are placed on the numerator and acid oxides on the denominator. As indicated in the compilation of Mills (1995), several different ratios have been devised for various metallurgical processes, and in particular the diferent stages ofthe iron and stee!-making cycle. In addition to including different components as appropriate, these ratios have. incorporated weighting factors in an attempt to take account of, the relative basicities or acidities of the various components, While some of these have been quite successful when applied t0 the specific conditions typical of a single process, none has achieved any wide applicability. ‘Another measure which fas been employed is Excess Base ‘expressions, of the form E basic oxides ~ ¥ acid oxides, where the concentration units can be weight percent, molar per cent or ‘mole fraction. The main problem with these expressions for both Iasicity ratio and excess base is that they iavolve an arbiteary 4ecision as to whether a component isa basic or an acidic oxide. For most components this poses no real problem, However, for the group of intermediate oxides, which includes such commonly ‘occurring oxides as Al,Os and Fe;Os, and also TiO>, this ean indeed pose some cificuty. The Jon-Onygen Atuaction, defined by the expression 221d, where "Z" isthe charge on the cation and ‘dis the sum of the ionic radii of the cation and the oxygen anion, has never achieved any widespread use as a measure of basicity, but in fact, it does have considerable meri, 2s will be discussed later. Other {quantities which have been suggested as indirect measures of basicity include the relative bond strength, the relative stabilities, ‘of compounds suc as the carbonate at 1873°K and the sulphate at 298°K, values of pO, analogous to pH in aqueous systems and s0 defined as pO =logN,.. , EMF measurements relative © & 1. Department of Metallurgy and Materials Sconce, Univesity of ‘Tonto, Toronto, Ontario MSS 3, Canad. + Orignialy published in the Proceedings of MINPREX 2000, published by The AusIMM. ‘The AuIB Precis standard slag and optical basicity. This last one willbe discussed in later Secuons, ater some necessary background has. been sven. In recent years, the CuICu* redox couple (Nakamura and Sano, 1991) and solubilies of metals such as plainum (Nakamura and Sano, 1997; 1998) and sier (Park und Moon, 1999) have also. been suggested 28 indicators of basicity. However, Yang and Belton (1998), who investigated the Fe'"fe™ redox couple. in. CaO-ALOySiOz and 20-MgO-AI,0;-SiO; slags found that a linear correlation with optical basicity was only obtained when the Pep, Bo ratio Was heli constant, This would seem o indicate tit the parameter chosen as a measure of basicity is also dependent on oxygen Povental the variables have tbe separated, and the tre Alependence on one can only be obtained when the other is held constant I seems clear ht basicity should be regarded as being intrinsically independent of ‘oxygen potential, except. when transition meal ions are present. When they are they wil be present as ions of more thn one veleney, and changes i he ratio Of the two ions eaused by variation in oxygen potential will have repercussions on the basicity of the sag, since the higher valency jon is normally significantly less baste than the lower valency joa, Thus, redox couples, or other parameters which also depend 6n the oxygen potential are only reliable measures of basicity ‘when the oxygen potential s constant Various capacities of components in slags, pariculaly the sulphide, phosphate and carbonate capacities, have been widely correlate with other measures of basicity, an hence have come to be regarded, at least by ome, a inicators of basicity in their ovum righ. Twenty-five years ago, Wagner (1995) suggested that the carhonate capacity in particular could prove to bean excellent measure of basicity, and in the years since this has been proved to be correct. The carbonate capacity is defined by the expression Ces = (wt por cent CO Yb, and was prefered to the sulphige capacity since it was fet tha ibe carbonate fon as Tes likely to become invelved in specific interactions with other species, particularly cations, in the slag, and there is now evidence that this predition ssl comet. Carbonate capacities have now been measured in a wide range of systems, including some containing halides, which are dificult or impossible o deal with by most ofthe eter approaches, and have proved tobe very Yalueble'indoed in permiting comparison of the. relative basicities of widely different types of slags However, ss note in an earioe paper (Sommerville, MeLean and. Yang, 1997), capacities ean only be regarded 43 reliable measures of basicity when certain limitations are recognised and respected, Thus, the concept of capacity should be used only when 1. the species in solution s Known, andi single species over te entie composition range, 2 the concentration of the solute is sufiieny low that it is within the range of obedience to Henry's La, so that is activity cocficent is constant, since this assumption is Jiplicitin he denitons of capi, 3. the ‘compouna ofthe element being dissolved is less sable than it oxide at the temperature concemed, because oaly thea isthe ‘compound! being dissolved rather than simply precipitated, and 4. where the capacity iscalelated using a relationship with a basicity expression rather than determined experimentally, the slag concerned must be know tobe completely iq Limitation (1) means, for example, that the phosphate capacity is not suitable asa measure of basicity asthe PO,” ion dimenses at contents 28 low as about three per cent. With regard to limitation (2), itis known that complexes ofthe Sion with the Vol 3065No #201 n TD SOMMERVILLE and YINDONG YANG Fe" cation occur as the amber chromophore in commercial alasses (Douglas and Zaman, 1969), and it seems highly probable that similar complexes can be formed in slags, although this has, not so far been demonstrated experimentally. I? S* ean also form complexes with other transition metal cations this would help to ‘explain why the values of optical basicity calculated from the correlation with the sulphide capacity (Sommerville and Sosinsky, 1984) would be erroneously high, ‘since such complexing would lower the activity coefficient of the sulphide jon, thus rasing its content and hence the value of Cs ealeulated using this content. [ti also entirely concsivable that atthe very high sulphur contents reached in highly basic slags, the content hhas moved out of the range of obedience to Henry's Law, so that its activity coefficient begins to increase, leading 10 a non-linear dependence of Cs on basicity, as observed by Young et a, 1992) ‘Thus, the carbonate capacity emerges as the best measure of basicity on the basis of limitations (1) and (2). However, itis ironic that i is the one, which has been most seriously misused by disregarding timitation (3). Very high carbonate capacity values have been reported in some systems, particularly several involving Na:O, implying very high basiities, but a simple thermodynamic “analysis indicates that since Na;COs is more stable than Na,O under the conditions employed, much of the carbonate is not being dissolved at all, but simply precipitated as, a separate phase a unit activity. Since, however, all the carbonate present is regarded as being in solution, the calculated carbonate ‘capacity is erroneously high. A similar situation may pertain to some carbonate capacities for slags containing BaO, and possibly also for water or hydroxyl capacities in some systems. Since NasS is more stable than NaO at steelmaking temperatures, ‘some sulphide capacity values for systems containing NaxO are possibly also erroneously high. However, this situation is most prevalent for carbonate capacities, and also most serious for these, because of the fact that carbonate capacity is so widely held tobe the best measure of basicity available, It isa relatively simple matter o caleulate the critical value of pn, which cannot be exceeded at any given temperature if this siation is to be avoided, and this was done in the previous paper (Sommerville, MeLean and Yang. 1997). There is also the point that capacity is 2 temperature-dependent parameter, and it can not be simply calculated using slag composition like V-ratio or optical basicity ‘THE CONCEPT OF BASICITY IN SLAGS FFor slags and glasses there have been two main approaches to this problem (Dufly and Ingram, 1976): 1. the Lux-Flood approach in which the, characteristic ‘acid-base reaction involves exchange of O° ions, so that basicity is perceived in terms of an oxide ion activity, a. and 2. the Lewis approach in which basicity is perceived in terms of the ability to donate negative change, so that anions such as O° behave as bases in the Lewis sense while the cations behave as Lewis acids tis true to say that the former approsch has won general ‘acceptance inthe treatment of acid-base equilibria in molten salts and slags, as well as in glasses, so that, following the classic paper of Flood and Forland (1947), it'has become virally ‘axiomatic thatthe correct expression for basicity in an oxyanion system is in terms of a,,, However, it should be clearly understood that there are to major difficulties associated with this approach. The frst is that Guggenheim (1929) showed that single 1on activities such as a,.. are indefinable because of the absence of a suitable standard 6r reference state, and attempts to tse the concept of oxide ion activides, or to set up Seales ‘analogous to the pH scale applicable in aqueous solutins, have Jed 10 several controversies in both the metallurgical and glass literature in recent yeats, although the possibility of measuring single ion activities has beon resurrected again fairly recently (Ma and Lu, 1993) ‘The second dlificuly is perhaps even more serious fundamentally. It the fact that the O% ion is a hypothetical entity, which has no independent existence in real systems. [ts formation would require completely ionic bonding which is never atiained in practice, and Jorgensen (1966) has shown that if it existed alone it would have infinite polarsablity and zero jonisation energy. By the same token, O” ‘ions’ would require ‘completely covalent bonding, which is never achieved in real systems. so that the caditional view of slags as consisting of three types of oxygen, OF, O° and O°, is eeally a mathematical convenience rather than a physical reality. The three types of ‘oxygens, namely bridging (doubly bonded), non bridging (singly bonded) and ‘free’ (unbonded) do exist, of course, but the assignment of integral charges to each of these types is almost certainly an unwarranted oversimpification, as shown by work in related fields Binks and Duffy, 1980; Kohan and Ceder, 1998; ‘Ceder eral, 1998). In spite of allthis, the hankering of metallurgists for some means of measuring a... or some quantity closely related (0 it, hhas continued. In the°absence of the ability to measure a, ively, recourse has beea taken to one or other of the indirset ‘measures discussed earlier, and in view of all this ti suggested that the application of the Lewis approach to basicity in slags and lasses has a good deal to commend it In the Lewis approach, basicity is perceived and expressed in terms of the state of polarisation, or more simply the “state’ of foxygen ions in the slag. This quantity must be clearly distinguished from both their concentration and their thermodynamic activity, though clearly a system where the ‘oxygen ions are only weakly polarised and therefore have ‘considerable residual negative charge will be equivalent 0 one ‘with a high oxide ion activity, and Vice versa. The statement of Flood and Forland (1947) that an acid-base reaction is one in ‘hich ‘an oxide ioa changes from one state of polarisation to another’ is entirely correct, but, 2s pointed out by Duffy and Ingram (1976), their assumption that this state of polarisation can be expressed in term of axide ion activity is not valid for glasses, and may well nt be ue in slags either. ‘Average or bulk values of this state are quantified by employing the concept of optical basicity, which was pioneered and developed in the field of glass chemistry by Duffy and Ingram (1971. 1976), Kohan and Ceder (1998), and Ceder et a (1998). In this approach, basicities are regarded in terms of the electron donor power of the oxygen ions present, as indicated by 2 probe ion dissolved in the slag undergoing the nephelauxetic effect (Duffy and Ingram, 1971; 1976) which can be measured spectsoscopically. The electron donor power of the oxygens is related t0 the average residual negative charge left-on the oxygens, which is inversely related to the extent to which they hhave been polarised by the surrounding eations. Inthe presence of eations such as silicon or phosphorus, with their high charge nd small radius, the oxygens are highly polarised, whereas in the presence of ions such as calcium or magnesium, the degree of polarisation is much les. Thus, in ‘acid’ slags where the oxygens are highly polarised, their residual negative charge will be low and therefore the extent of electron donation experienced by the probe ion will be low, so that the extent of the nephelauxetic effect and thus the optical basicity will below. In “basi” slags, the reverse will be the ease Before moving onto the evaluation of optical basicity, wo ‘other points in favour ofthe Lewis approach may be mentioned, Firsily, it removes the idea that the O* ion is wnigue, and therefore opens the way for other anions such asthe halides to be incorporated into the scheme of basicity. This is particularly useful for the Muoride ion, which is isoelectronic with O® and very similar in many ways, because ofthe presence of fluorides, in important metallurgical slags and fluxes. Secondly, it avoids the need for discussion as to whether or not the ayy in a ‘multicomponent system is an accurate measure of 8,5 ‘THE EVALUATION OF OPTICAL BASICITY ‘The average charge donating power of the oxygens can be determined experimentally by introducing trace quantities of ‘metal ions such as Pb or TI* and measuring the ed shift in the UY spectrum brought about by the expansion ofthe cuter sand p electronic orbitals caused by the charge donated from the oxide ‘The AmIMM Proceedings environment. In addition to the normal spectroscopy techniques, (Dufiy and Ingram, 1971; 197S2; 1976), photoacoustic spectroscopy has also been applied recently to the measurement ‘of optical basicites of slags (Nakamura, Ucda and Noguchi, 1984; Nakamura, Ueda and Toguri, 1986) On the basis ofthe experimentally measured values of optical basicity, correlations have been developed which allow extension ‘of this approach to other components. and. systems. For non-trnsition metal oxides Dufly and Ingram (1975p) deduced, ‘an excellent correlation between the optical basicity of the oxide, ‘A, and the Pauling electronegativity, x, of the cation. This has proved an extremely useful conelation, which has allowed pplication of optical basicities to many more systems than have been studied spectroscopically, but the fact remains that the Pauling electronegativty is a rather coarse measure in the sense that it is normally only quoted to the first decimal place so that fine differences hetween cations cannat be taken into account. Similarly, Nakamura, Ueda and Noguchi (1984) and Nakamura, ‘Ueda and Toguri (1986) developed an alternative scale based on average electron density which provides values for the transition metal oxides and also fair number of halides as ‘well as forthe main series oxides. Values calculated from Pauling clecironcgativities and from average electron densities for oxides fof interest in iron and steel-making are compared in Table 1, ‘where the agreement between the two is reasonable for the main seties oxides, but not for the transition metal oxides. Cleary, the ‘ales for these later oxides calculated from the average electron densities are more reasonable, and therefore this scale may wel be the prefered one for metallurgical slags which tend to contain, at Jeast small quantities of one or more ofthese transition meta oxides, ‘On the optical basicity scale, the state of polarisation of the ‘oxygens can be both defined and quantified since a standard state has been designated, namely the “state” of oxygen in calcium ‘oxide. Thus, although the optical basicity scale is constructed on the basis that A (CaO) = 1, and the A values for all other components are measures’ of ‘lime character’ or ‘lime equivalence’, the values can be used in systems which do not Contain lime. It should also be noted that this approach allows realistic basicity values to be ascribed to slags which contain no recognised ‘acid’ constituents. BASICITY OF METALLURGICAL SLAGS (One of the most useful aspects of the treatment developed by Dafly and Ingram isthe ability to calculate average values of A {ora slag of any composition by means of the expression Aa Kegan Hayle + ow where X i the equivalent cation fraction of each oxide. The most convenient method of calculating values of X for components in slags has been outlined elsewhere (Sommerville and Sosinsky, 1984; Sonsinsky and Sommerville, 1986). One serious problem in the aplication of optical basicity 10 slags isin the assignment of A values to the transition metal oxides. It appears from a number of statis (Sommerville and Sosinsky, 1984; Bergman and Gustafsson, 1988; Bergman and Gustafsson, 1989; Carey era, 1989; He, Serge and Carey, 1991) that the effective” optical basicity of a wansiion metal oxide vaties depending on the specific chemical reaction in which the slag is participating, For example, itis larger for the sulphar equilibrium than forthe phosphores equilibrium. This may be ‘ue to diferences in the extent of interaction with the sposife anion concemed, or to the fat that te oxypen potential and therefore the redox equilibrium between the valeny states ofthe transition metl cation is very diferent inthe two esse, © 10 & combination of both these efecs. This problem has been well Known for over 20 years Also included in Table | is list of recommended values for the optical bascites ofthe oxides of interest in metal extraction. In genera, this ist follows that of Nakamura, Ueda and Noguchi (1984) and Nakamura, Ueda and Togu (1986) very closely, while assigning sighly higher values to Na,O and K:0. The ‘values assigned 10 FeO and MnO are in fly good agreement With those esigned by Dully and Tagram (1992) onthe basis of the elatonship between optical besiity and the polrisability of the O* species for silicates or oer sets" of analogous compounds [Mills (1995) has suggested that when SiO: is replaced by ‘AhOn, cations are required to provide electical neutrality, and ‘te therefore not available for network breaking, so thatthe elective compositions and therefore the optical basicity of the tmelt must be adjusted to take this into account. The corected ‘optical basicity, Aan, bas 80 far only been colated with Physical. properties such as viscosity, clecrical conductivity, {ermal conductivity and the thermal expansion coefficient Tame 1 Yates of eprica basic of oxides from diferent scales. ovite Tec bona faction | Opal basis, | Opal ass, A | Optic bas, A, derive am Pt | derived fem aneage | "recommended ats letroncgtviy = | Seton demi Ko oat ry uo Lis 10 Nao 036 08s Ls 0 | io aso os Ln Los as ass os us Los Lia ass st 10 Loe as : 070 1 1 Lon 1 083 a1 059 a9s__| ass os 038 ast on 093 9s oss ar an | oss Las oat ass ase as C La 036 as ase 09 Ls ow | oat ass ass [ss oat ast ase ss | 26 | 03 0a 042 Hi Dat 036 as ou? Fm a2 a0 038 Te AuiMA Poses 06 1200 0 1D SOMMERVILLE and YINDONG YANG THE CORRELATION OF OPTICAL BASICITY WITH OTHER BASICITY INDICATORS Figure 1 shows the correlation between optical basicity, A, and both the CaO/SiO, and the CaQ/AL:Os ratios, each in the binary systems, For each curve, the solid portion is the part of the system Which is liquid at 1600°C, and i i obvious immediately that the whole of the liquid field in the CaO-ALO, system is, more basic than that in the CaO-SiO: system, which tes in with ‘common experience in Iadle slags. However, it should be noted that very few other basicity indicat would allow this simple clear comparison. I is also clear that within the limited range of ‘compositions liquid at 1873°K, there is an almost linear relationship between A and both ratios, so that these ratios are fine within these limitations. ol 040080060070 O80 iG 1 - Relationship between optical basicity ad basicity ratios ‘A indicated eater, although the ion-oxygen attraction bas not boon widely used, it does have intuitive appeal as a measure of the relative basctes of different oxides. Thus, in Figure 2 the Togarthm of iis ploued agains he optical basicity values of a number of oxides: When the Duly stale is used, a very good comelation (? = (965) is obained when te four transition metal ‘oxides, FeO, MnO, Fe,O; and CrOy ae omitted. When these fare included. the coreaton isnot nearly so stong (= 0826), indicaing again that de values of derived {rom elecso- negativities are not generally applicable. Using the values of Nakamura, Ueda and Noguchi (1984) and Nakamura, Ueda and ‘Togui (1986), a very good corelation (7° = 0.940) is obuined, cven when the four vansition metal oxides are included, From ‘ther work it seems that the values derived from average electron density are rather Tow for NasO and KO, and wien the values recommended in this paper are employed, an_ excellent conelation (# = 0982) is obtained, as shown in Figure 2. It Scems surprising tat such a good corcation is obained between quantity which implies the existence of O ions. and the opial basicity. which takes into account that OF ions cannot exist independanty as such In fact, the charges on both the oxygen anions, go, and the cation presen, gy, have Been shown 10 be linearly dependent on the optical basicity ofthe melt. On the basis of wave mechanical calculations forthe iso-electtonie sores AIO,", Si0.*, PO.", 04? and C10<, Binks and Duffy (1980) have established that go = LISA. Calculations on qu have sinee been caried out by Duly, Chriss and Kamitos (1995) for binary borate systems, and it would seem reasonable to presume that qualitatively m ‘ol306No 12501 10 - y=0.81 - 1.06x (20982) 0s + Ss g Q 004 2 08 4 10+} 7 1 0.0 02 04 06 08 10 12 14 16 OptB(R) ic 2 - Relationship between ion-oaygen atracton and optical basicity of various oxides. Similar considerations apply to silicates and silico-aluminates. ‘The relationship between qo and A. also allows the calculation of ‘microscopic and group optical bascities for complex silicate anions, Which have been discussed elsewhere (Sommerville and ‘Masson, 1992), but are really only of academic interest. Relative bond strengths or relative stabilities can be quantified as the Gibbs free energy of an exchange reaction such as, for example: C204) 41/2825) = CaS + 12025) 2 which is sometimes expressed as A(AG*). The values of [A(AG®) for oxide-sulphide equilibria are summarised in Figure 3, where they are shown to display a strong dependence on oxide basicity, and almost no temperature dependence. A similar plot for oxide — carbonate equilibria shows considerably more ‘caters and a much greater temperature dependence. With regard to redox equilibria, the work of Yang and Belton (1998) mentioned earlier is entirely consistent with studies of redox equilibria in glasses (Baucke and Duffy, 1993; 1994; DDufly, 1996). In these studies, the temperature was constant at 1673°K, and the partial prossure of oxygen was fixed because the melts were in equilibrium with air, and therefore the oxygen Potential was constant. Linear correlations were obtained between the logarithm of the ratio of the concentrations of the ‘oxidised and reduced species and optical basicity for a very wide range of redox couples (Duffy, 1996) as shown in Figure 4 It has also been shown that equilibria between metal ions and the retallic sate in melts can be tated in a similar fashion, and again linear correlations are obtained between the logarithm of the ratio ofthe concentrations of the species and optical basicity (Duly and Baucke, 1997). Correlations of optical basicity with various capacities have ‘been reviewed in a number of papers over the years, and there is, no need to review these further in this paper. Excellent correlations have been obtained with sulphide, phosphate, water and carbonate capacities, and the last have proved particularly ‘useful in oxide-Mluoride systems. The fat that these capacities ail, correlate with A. also means, of course, that they correlate with each other, and ean in fact be calculated from each other, so ‘greatly reducing the amount of experimental work required to establish these for any one system. ‘Tae AusIMM Poss @ r=1273k) —O t773k) 08 09 10 14 12 13 14 15 ‘Opt. BIR) ‘T= 1273 K, Yo606:38 -$1430X, ( =0.996) T= 177K, Y=618A0 528.24K, (= 0.997) G3 - Relationship between the fee energy for oxide-sulphide ula andthe optical basicity ofthe oxide oxidised tog ( ———_) 0.50 055 060 0.65 Optical basicity 0.70 lo 4 Plo of logarithm of (xidsedreduced) for meal ion ouples| versus optical basicity for silicate glass systems a 1673” K and at -aosphere Iso-optical basicity lines in ternary systems {can readily be shown that iso-optical basicity lines in temary systems are straight lines when the concentrations of the components are expressed either as mole fractions or in weight per cent. If equivalent cation fractions are employed, these lines will also be parallel and equally spaced, but tis more convenient touse the more conventional units for concentration, ‘The Aus roses BASICITY OF METALLURGICAL SLAGS, Since the lines are straight, the ine for any value of A can be ‘obtained simply by joining the points on the two binaries where that value of A is attained. Insertion of the values for the two equivalent cation fractions in terms of the mole fractions of the two oxides in Equation 1, and recalling thet for any binary system Ny, = (No, Yields the expression: Ayo —¥A Iw, “oe ° Nth, PB, ‘This expression was used to calculate the iso-optical basicity lines in Figure Sa-Sd, which illustrate the effect of varying the third component ina temary system where the other (wo ‘components are Time and silica. As can be seen, completely different pattems are obtained depending on whether the basicity ‘of the third component is between that for lime and silica, as in the case for alumina and magnesia, greater than that for lime, as in the case of barium oxide, or less than that of silica, asin the ‘ease of phosphorus pentoxide. By comparison of these diagrams, it is possible wo formulate the general rule that the iso-optical basicity lines always ‘fan out’ from the corner occupied by the ‘oxide with the intermediate value of basicity. Also, with SiO at the apex ofthe triangle, the closer the A values of the two oxides ‘onthe base of the triangle, the lower the slope ofthe iso-A lines will be. This is clearly shown by comparison of the patterns for the Ca0-Al,0s-Si0; and CaO-MgO-SiO> systems. In View ofthe cortelations that have been developed over the years, these iso-A. lines ae als0 i80-Cs, 1S0-C,,., 18Cy,o and is0-C.., ines, ‘which makes them extremely'uieful. A Knowledge ofthe way in which basicity, and consequently refining capebiliy, varies across any given temary system is surely very valuable and powerful information, ‘As an example of this, Figure 6 shows the comparison of the is0-C,, lines in the CaO-AL,O»-SiO sysiem produced by IRSID (Dor @f al, 1988; Ribaud and Gatellier, 1985) with the iso-A lines calculated on 2 weight percent basis using the values being recommended in this paper, and as can be seen the degree of parallelism is excellent Furthermore, in the case where there is only one clearly basic oxide in the system, and therefore one oxide is making the major contribution to basicity, these iso-A. lines come close to being Iso-actvity lines for that basi oxide. This has been demonstrated by Gaskell (1989), as shown in Figure 7, forthe iso-activity ines, for CaO in the CaO-AlOy-SiO; system, and less convincingly for activities of Nas in the Na0-P,0;-SiO, system. In Principle, this would allow at least first. approximation calculations of the iso-ay lines in a whole series of systems MO.ALO+-Si0: and MO'SiOrP.O., where MO could be any ‘oxide which is atleast as basic as CaO. Itis also of interest to note that in the everse situation, where there is only one clearly acidic oxide, a similar effect is observable. An example of this is the FsO-MnO-TiOs system, where the iso-sctivty lines for TiO; are almost parallel to is, iso-content lines, which is exactly what would be predicted from the fact thatthe optical basicity values for FeO and MnO are very close together. A qualitatively similar effect is also seen in the FeO-Mn0-SiO; system, but the degree of parallelism is not so impressive. Systems involving FeO, CaO and an acid oxide ‘would seem to fall into this category, but the fact that CaO stabilises the feric ion complicates the picture, since it means that these are not truly ternary systems. CONCLUDING COMMENTS ‘As has been demonstrated, the essential approach of metallurgists to the concept of basicity’ has been to favour the activity of the fice oxide ion, a, ws the real measure, and, since this is not attainable, to adopt a plethora of indirect measures which approximate to it. In view ofthis, it seems appropriate to ask two ‘questions: or 306 1200 8 1 SOMMERVILLE and YINDONG YANO Maa (8) 30-A10,Si0, (©) C20.MgO-SI0, 310, eae ()C20-B20.Si0, 0, (C107: Si0, [Fic 5 -Calulte iso-opticalbasiciy tines in various ternary systems 8. How many more indirect measures are we going to adopt? b. What criterion do we have for assessing their relative validity when they differ? Since the answer to the second question appears to be none, ppethaps the answer to the first question should also be none ‘Overall, it really does seem thatthe Lewis approach to basicity in ‘oxide melts is fundamentally more satisfactory. The paper of Flood and Forland (1947) was a magnificent contibution st the time, and served us well for many years. But advances since then in related fields have clearly shown that oxygen ions in solid oxide mixtures, glasses and slags do not generally have exactly integral charges, and that the assumption that they do leads 10 erroneous values for various quantities. AS far as the present authors are aware, the Lewis approach is the oaly ane’ which ‘makes allowance for non-integral charges, and therefore it is recommended that it should be adopted for expressing basicity, at least in fundamental studies, 6 oI 306 No 2901 REFERENCES Baucke, FG K and Duffy. JA, 1993, Redox actions between cations of ‘ferent polyvalent elements in glass mols: an opal basicity study, Phye Chem Glasses, MA) 158 163, Baucke, FG K and Duffy, JA, 1994, Oxidation states of metal ions in alas melts, Phys: Chem, Glasses, 35 (11721 Binks, JH and Duffy, JA, 1980. A molecular orbital treatment of the bsety of oxyanion mats, J Non-Cryst Seis, 37387400, Bergman, A and Gastson, A, 1988, On the relation between optical Tasty ‘and phosphorus capacity of complex slags, See Res, 5981-288, ‘Bergman, A and Gustafson, A, 1989 Use of optical basicity to cafelate Dosphorus and oxygen contents in molten ina, in Proceedings Third nernatonal Confrence on Molten Slags ane Fase, Institate of Metals, pp 80-153 (TMS. AIME: London), (Carey, CL, Serge, RJ, Gregor, K and He, QL, 1989, Optical basicity: a ‘exible bass for Tux contol in stelmaking, in Proceedings Third Iniersational Conference on Molten Slags and Fluxes, Insite of Metals, p 157-162 (TMS-AIME: London). BASK ITY OF METALLURGICAL SLAGS. cao Pr rr rr ro Weiss ALO, iG 6- Water capacity in CaO-ALO, Si, lags at 1823°K, (Ceder,G, Kohan, AF. Aydinol, M K, Tepesch, PD and Van der Ven, A, 1998. hemodynamics of oxides with substtuional disorder a Imirescopie mode and evaluation of import enerty contusions, ‘Amer Ceram Soe, 813)S17-52, Dor, PH, Carer, B, Naif, M and Gateier,C. 1988, Inlusace of s ‘on sel dehydrogenation under vacuum, IRSID Report No 13 19. Maire les- Met, France. Douglas, R W and Zaman, M S, 1969. The chromophore inion sulphur "amber glasses, Phys Chem Glasses, 103): 125.182 Dally, J A, 1996, Redox equilibria in glass, J Non-Cryst Solids, 1964550, Duly, J A and Baucke, FG K, 1997. Efe of glass composition and ‘set on reduction of metal ions othe metallic state in melts, Phys Chem Glasses, 950) 25-2, Duly, JA, Chrssikos, GD and Kamit, E 1, 1995. Chemical bonding ‘alysis of alkali oxide glass systems: charges on metal ions, Pye Chem Gases, 360)53.58 Duly, and Ingram, MD, 1971. Batalishment of an optical sale for Lewis basiiy in ‘inpanic oxyacids, moltn salts and. lasses, 7 Amer Chom Sec, 916418.6458 Daffy, J A and Ingram, M D, 1975a, The behaviou of inicator ion ia Felaton to the ideal optical basicity of glass, Phys Chem Glasses, e119-133, Dull. 1 A and Ingram, M D, 1975b. Opical asi, TV influence of. clecuoseyaivity on the Lewis busty and solvent properties of ‘molten oxyanion sas and glasses, J) Jnorg Nuclear Chem, 37:1203-1206, Duly, J and Ingram, M D, 1976, An interpretation of glass chemistry In terms of the optical basicity concept, J Nor-Cryst Solids, 2137-40. Daly, 1A and Ingram, MD, 1992. Comments on the application of ‘ical basicity to glss, J Non-Cryst Solids, 1476-40 Flood H and Foran, T, 1947. The acidic and basic properties of oxides, ‘Acta Chem Scand, 1392-604 Guggenim, EA, 1929. On acivtes of single ions, J Phys Chem, 35.8125, Gaskell, DR, 1989. Optical basicity andthe thermedynamie propentin of slags, Metall Trans B, 20B:1. 3118, Jorgensen, C K, 1966. Structre and bonding, Vol 1, pp 234-248 "Springer. Vetag: Bess), Kohan, AP and Celer, G, 1998. Charge tansfer in multicomponent oxides, Pls Rev B, 310) 3838-3882. He, QL, Serge, R J and Carey, C L, 1991. Application of the optical basicity concept to fx conta nthe RUF steelmaking process ron and Steetmaker, 18(7) 31-33. ‘The Au Prose F107 -Io-a, contours and iso-optical basicity ines in the sytem C20-ALO,SiO, a1 879° Ma, D and Lu, W-K, 1993. A study of ion activites in ionic solutions, Metall Trans B,248:317-323 Mills, K C, 1995. Te bsiiy and optical basicies of slags, Slag Aas, pp 919 Veriag Suaeisen GmbH: Duselder. "Nakamura. S and Sano, N, 1991, Th redox rquibris of copper tons in the molten silicate les a8 8 measure of basicity, Metall Tans B, 223.829, Nakamara Sand Seno, N, 1997, Solubility of platinum in molten fixes ‘6 a measure of basicity, Metall Trans, 288:103-108. ‘Nakamura, $ and Sano, N, 1998, The influence of basicity om the ‘solubility of platinum in oxide mel, Metall Trans B, 29B:411-814 [Nalamura,T, Ueda, ¥ and Togus, 1M, 1946, new development of the ‘optical basic, J Japan fast Metals, 80456 401 Nakamura, 7, Ueda, ¥ and Noguchi, F, 1984. The characterization of| slags ‘using photcaroustic spectroscopy, in Proceedings Second International Symposium Metall Slags and’ Fuses, pp 1005-1013 (TMS-AIME: Lake Tahoe, Novad). Park, JH and Moon, D J, 1999. Quanittive analysis of the elaive basicity of CaO and BaO by silver solubility in slags, Metall Trans B, 3069-694, Ribaud, PV and GatellerC, 1985, New products: what should be done in” secondary steelmaking? formating ond Steelmaking, 12e)79-86- ‘Sommerville, 1D, MeLean, A and Yang, YD, 1997. The use and misuse ‘of capacities in sigs, in Proceedings Fh Intemational Symposium ‘on Molten Slag, Faxes and Salts, pp 375-383 (ISS-AIME:Syaes, Ausaia, ‘Sommerville, and Sosinsky,D J, 1984 The application ofthe optical basicity “concept to metallurgical slags, in Proceedings. Second International Sympesiam Metall Slage and’ Flutes, pp 1015-1026 (CTMS-AIME: Late Taos, Nevada), Sommerville, I D and Masson, C R, 1992. Group optical basics of polymerized anioas in sags Metal Trans B,25B:227 229 Sosinsky, DJ and Sommerville, I'D, 1986. The composition and temperature dependence of the sulphide capacity of metallurgical slags, Metall Trans B, 17B:331-397 Wagner, C, 1975. The concept of basicity in slags, Mull Tens 8, (8405-400, ‘Yang, 2nd Bellon, GR, 1998, lon redox equi in CaO ALO S10; ‘nd MgO-CiO-AlsOy-SiO3 sags, Metall Trans B,29B:837 845. Young, R-W, Duy, J A, Hasal, GJ and Xu, Z, 1992. Use of optical icky concept for determining phosphors aad sulphur slag-netal partons mating and Stelmaking,12:201-219 Vo! 206 1 201 n

You might also like