You are on page 1of 6

Journal of

AND COMPOUNDS
ELSEVIER
Journal of Alloys and Compounds 220 (1995) 126-131

Calculation of the effect of alloying elements on the Ms temperature in steels


K. Ishida
Department of Materials Science, Faculty of Engineering, Tohoku University, Sendai 980, Japan

Abstract

A thermodynamic calculation was carried out to predict the effect of alloying elements on the Ms temperature in iron-base alloys. It is shown that changes in Ms temperature by alloy addition are due to both chemical and mechanical effects. Magnetic and non-magnetic contributions to the Gibbs energy constitute the chemical part, while the mechanical effect arises mainly as a result of the change in friction stress required to move transformation dislocations, which in turn is related to solid solution hardening in the parent austenite. In this paper the chemical contributions to Gibbs free energy are evaluated quantitatively from thermodynamic data. These data are used in combination with mechanical data to calculate the Ms temperature. Calculated M~ temperatures are compared with results obtained using previously reported empirical formulae.
Keywords: Martensitic transformation; Ms temperature; Thermodynamics; Steel; Free energy

1. Introduction
The effect of alloying elements on the Ms temperature of steels has been investigated extensively [1-8], and it has been well established that most elements except for Co and AI decrease the Ms temperature. For practical purposes, several empirical formulae have been proposed for estimating the Ms temperature from the alloy composition [1-7]. It has been noticed that several ferrite stabilizers such as Mo, Si and V have opposite effects on the Ms and To temperatures, where To is the temperature at which the Gibbs energies of ferrite (a) and austenite (3') are equal. Normally one would expect Ms and To to be affected in the same way. To explain the anomaly in the case of the above additions, Zener [9] suggested that the magnetic free energy plays an important role in determining the direction of change in Ms temperature with alloying addition. As reviewed by Kaufman and Cohen [10], however, Zener's treatment that relates the effect of alloying elements on Curie temperature directly to their effect on the Ms temperature has only limited applicability. Hillert et al. [11] developed Zener's model in a more quantitative way and have shown that the temperature dependence of the stabilities of the ferrite and austenite phases can be affected significantlyby alloying elements through their effect on magnetic contributions to the Gibbs free energy. A more extended treatment on similar lines

has also confirmed the significant effect on the od7 equilibria of many elements through their effect on the magnetic contribution to the Gibbs free energy [12]. The complex temperature dependence of the distribution coefficients of alloying elements between o: and 7 phases has also been analyzed quantitatively by taking magnetic contributions into account over a wide range of temperatures [13]. The availability of such thermodynamic data pertaining to the stable a/7 equilibrium at low temperatures, where martensitic transformation often occurs, makes it possible to estimate the driving force accurately for the martensitic transformation and thereby to calculate the effect of alloying elements on the Ms temperatures in steel. The present study describes in detail the thermodynamic treatment involved in calculation of the Ms temperature as a function of alloying elements.

2. Thermodynamic calculation of Ms temperature


As mentioned earlier, calculation of the Ms temperature in iron-base alloys requires that the Gibbs free energy of the stable phases, namely ferrite and austenite, be calculated first. This enables calculation of the driving force for martensitic transformation, which can then be combined with the estimated contribution

0925-8388/95/$09.50 1995 Elsevier Science S.A. All rights reserved SSDI 0925-8388(94)06002-9

K Ishida / Journal of Alloys and Compounds 220 (1995) 126-131

127

from mechanical effects, leading to a prediction of M~ temperatures as a function of alloying elements.


2.1. Estimation of driving force for martensific transformation

The driving force for martensitic transformation is defined as the difference in Gibbs energies between parent (austenite) and product (martensite) phases, and is .generally proportional to the degree of supercooling between To and Ms. It has been recognized that the critical driving force for martensitic transformation in iron alloys at Ms is about 900-1300 J moland the supercooling is greater in comparison with nonferrous alloys [10]. If the free energy of martensite is assumed to be the sarae as that of ferrite, the difference in Gibbs energy between martensite and austenite in the Fe-M binary system can be described by the regular solution approximation as AG~'~ = AG~'~(1 - x ) + A G ~ ~x+ AD~'~x(1 - x ) (1) where A GF. and AG~~ are respectively the differences in Gibbs energy between b.c.c, and f.c.c, forms of iron and the component AI2~'~ is the difference in interaction parameters of both phases and x is the mole fraction. Eq. (1) can be simplified in the dilute solution case as AG~.2~ = A G ~ " ( 1 - x ) - AG~vF~x (2)

where the terms labeled p and f denote the values in fully paramagnetic and ferromagnetic states respectively, T~ and T~ are the Curie temperatures of pure iron and solid solution respectively, ATM(ATM=~Tdrx) is the rate of change of Curie temperature with alloying addition, and T* (T* = T(TdT~)) is the converted temperature. The constant m which represents the effect of alloying elements on the magnitude of the ferromagnetic term is assumed to be unity for all the nonmagnetic alloying elements, and is assumed to be zero for Ni and Co. The ferromagnetic parts of the thermodynamic function of a-Fe, [AG~] f and [hHv~], ~ f have been evaluated from the data of specific heat of pure Fe [15] and are shown in Fig. 1 together with the difference in Gibbs energy between y and a Fe [13,14]. These values in ferromagnetic parts are essentially the same as those presented by Hillert and Jarl [16] and Chuang et al. [17]. trm V M ] is the ~o~F~lp paramagnetic term of AG~ vF~ and has also been evaluated as listed in Table 1 [13].
2.2. Thermodynamic criterion for the onset of martensite transformation

The thermodynamic criterion for a martensitic transformation to occur in a system can be written [10] as AGe +AGnc ~<0 (7)

where AG~ ~F~= - (AGG~" + AD{~"). The value of AG~ ,v': is the parameter which represents quantitatively the effect of alloying elements on the relative stability between a and 3' phases and is defined as a change in partial molar Gibbs energy accompanied by transfer of 1 mol of each alloying element from dilute ferrite to austenite [11,12]. This parameter can also be estimated from the od3" phase equilibria: AG~ ~F'~= R T In x~/x ~ (3)

where AG~ ~" and AG,Vg are the chemical and non"" chemical free energy changes respectively. The Ms point is defined as the temperature at which the chemical driving force for the transformation exceeds the nonchemical free energy. Although the non-chemical free energy has been considered mainly to be a combination

-2000

where x" and x" are the equilibrium compositions of a and 3" phases respectively. The parameter AG~ ~F* has generally a large temperature dependence, mainly owing ~o the magnetic effect, and can be expressed by the following equation [13,14]:

/ [A G Fe ] ~
/ / / / / /

~-4000
w

//

a G ~ : ' - = laG ~:,"M ~'+ [AG~"] ' ]


where

(4)

~-6000
w
:/
/ /

,/

/ [A H~Fe ] f

[AG~,"o]. = [acT~- , ] . + [AO~2U]"


and

(5)

-8000

400

[AG~V~] f = - (1 - m )( T/T~)[AG~( T* ) ]f
- (ATM/T~)[AH~] f

600 80'0 1000 Temperature/K

1200

(6)

Fig. 1. Ferromagnetic terms of thermodynamic functions for pure ot-Fe.

128

K Ishida / Journal of Alloys and Compounds 220 (1995) 126-131

Table 1
V a l u e s o f [AG~Y] P Element [ A G ~ r ] p (J m o l - z ) - 54390 + 24.334T 4480 - 0.725T - 3440 + 1.325T - 1522 + - 1790 + -5200 + 1860 + 5996 - 7820 + 7870 7716 41005439 1.728T 1.050T 3.110T 1.175T 0.444 T 3.825T 3.820T 1.324T 0.460T 1.255T

C
AI Co Cr Cu Mn Mo Nb Ni Si Ti V W

where d is the interatomic spacing of the shear plane on which the transformation takes place, b the Burgers vector of the transformation dislocation and z the transformation stress. The magnitude of transformation strain (b/d) is evaluated to be approximately 0.24 [22].

3. Calculation of the change in Ms temperature with alloying element


As mentioned earlier, Fig. 2 shows schematically the disposition of the Ms temperature with respect to transformation energy arising from the effects on Ms which are now evaluated quantitatively as follows.

3.1. Chemical effect


Fig. 3 shows the schematic representation of the change in Ms temperature due only to the chemical effect. Let the chemical driving force be changed from [ A G ~ ] F e to [AG~-~]F~M by alloy addition:
rILl

A(AG)-[AG
--xSAa~ (x << 1)

ko
(10)

coO --

,p.o
03 t-

C
Ms Temperature To

where 3 A G e = - AG~ ~v~- AG~U". As discussed in the foregoing section, the chemical part of the Gibbs energy can be divided into paramagnetic and ferromagnetic terms, obtained from Eqs. (4)--(6) as

~AG~= [SAG~].+ [~AQ]'


where [~AG]" = - [aCr2F] ~- [aC{~;" F and

(11)

(12)

Fig. 2. S c h e m a t i c i l l u s t r a t i o n o f t h e r e l a t i o n b e t w e e n M~ t e m p e r a t u r e and transformation energy.

of the interracial energy AG~, and the elastic strain energy AGst, it has been shown that allowance should be made for an irreversible expenditure of mechanical energy connected with the shear mechanism of martensitic transformation in the overall free energy balance [18-21]. This shear energy AGsh is consumed by the frictional stresses involved in moving the transformation dislocations. In view of this, therefore, the non-chemical energy can now be expressed as AG,~ = AG~, + AGs~+ AGsh

\
~) ~

\C
~ ~a 1

(8)

Fig. 2 is a plot of transformation energy vs. temperature, which illustrates schematically the variation of the different energy contributions with temperature and the resulting position of the M~ temperature. The shear energy component AG,h of Eq. (8) is given by the following equation [18]: AG~h = (b/d) ~" (9)

[Ms'] P [Ms] P

Ms"

Ms

To' To

Temperature
Fig. 3. S c h e m a t i c i l l u s t r a t i o n o f t h e c h a n g e in Ms t e m p e r a t u r e d u e to t h e c h e m i c a l effect.

K. Ishida I Journal of Alloys and Compounds 220 (1995) 126-131

129

[3AQ] t = _ rn(T~/T~)[AG~.(T*)] ~
I~ II II I ~

+ ATM[AH{=~''~(T*)]f/T,

(13)
I
I I I

Accordingly, the change in Ms temperature due to the chemical effect alone is calculated as
AM, = A (AG~)/AS ~-~"~

= [AM[] p + [AM~]r where [AM~]P= x[ $AG~]PlAS~E"" and

(14)
I

(15)

[AMZ]'= x [ 6 A G , ] f / A S ~ " "

(16)

I I I

where AS~C" is the entropy change of the 3,~ ot transformation of Fe.


3.2. Mechanical effect

II

II

II

I1~1

I I I

The change in Ms temperature due to the mechanical effect alone is shown in Fig. 4, where the non-chemical energy is changed from [AG,~E'"]F~ to [AG,~ ]r~M by the addition of the alloying element. The effect of alloying elements on the non-chemical energy has not been analyzed quantitatively. One can expect, however, that the shear energy AG,h would be changed by alloying because the stress necessary for moving transformation dislocations is related to that needed for moving lattice dislocations which in turn changes on solid solution hardening. This change in shear strength due to solid solution hardening and its effect on the flow stress of austenite and the 3,/, temperature have been observed by Breinan and Ansell [20]. In a previous paper [21],

7
~ ~ i ~ ~

-..,.,
O

IFeM I ~
e~ o

IU
(3

c3

!-

I--

mi
e,

7,?-71

Ms'

Ms

To

Temperature
Fig. 4. Schematic illustration of the change in M, temperature due to the mechanical effect.
[-- ~.1

130

I~ lshida / Journal of Alloys and Compounds 220 (1995) 126-131


3 and observed changes in M~ temperature in literature Ms (C/wt.%) in this work [1] [21 - 361 [31 - 300 [41 - 324 [5] - 361 +30 + 15 - 28 - 39 - 22 - 27 - 20 - 10 - 33 - 11 - 17 -11 - 39 -28 - 19 - 33 - 11 - 17 -11 - 32 - 11 - 16 -11 -35 -11 -11 -5 - 39 -5 - 17 - 33 -21 - 17 - 30.4 -7.5 - 17.7 -7.5 - 17 - 12.1 [61 - 474 [7] - 423 - 330 +2 + 7 - 14 - 13 - 23 -5 -4 - 13 -7 +3 +4 0

Table

Calculated Element

Ms (C/wt.%)

C AI Co Cr Cu Mn Mo Nb Ni Si Ti V W

- 317

the present author has also shown that the critical driving force of martensitic transformation in Fe-Ni base alloys is not constant but changes by alloy addition and is strongly related to the solid solution hardening effect. AGin and AGst, the interfacial and strain energy parts of the non-chemical energy in Eq. (8), may also be affected by alloy addition. However, this has not been analyzed quantitatively. Since it has also been suggested by Magee [19] that the shear energy AGsh forms the dominant part of the non-chemical energy term, it seems reasonable to assume that the effect of alloying elements on (AGin + AGst) is negligible. Then, the change in Ms temperature due to the mechanical effect per 1 at.% alloying element can be expressed by
= tSAG.c /ASFe

in transformation shear strength AT is assumed to be one-half the change in the yield strength.

4. Results of calculation

The total change in Ms temperature due to chemical and mechanical effect is given by AMs = AM~+ AM~ (19)

= (b/d)AT/AS~'"

(17)

where AT is the change in shear strength of austenite on adding 1 at.% solute. Although the effect of alloying elements on the solid solution hardening of f.c.c. Fe is not well established, AT can be estimated using the following equation, which gives the yield strength as a function of alloying elements in austenitic stainless steels and has been proposed by Irvine et al. [23]: 0.2PS (N mm -2) = 68 + 494(wt.% N) + 355(wt.% C) + 20(wt.% Si) + 4(wt.% Cr) + 15(wt.% Mo) + 19(wt.% V) + 5(wt.% W) + 40(wt.% Nb) +26(wt.% Ti)+13(wt.% A l ) + 2 ( 6 % ) + 7 d -1/2 (18) where ~ refers to the 6 ferrite and d is the average grain size. The difference in yield strength between alloyed and unalloyed austenite can be calculated using the above formula. In the present calculation, the change

where AM~ and ~ are calculated from Eqs. (14) and (17) respectively. Since the martensitic transformation in dilute Fe alloys takes place in the range 500-700 K and the paramagnetic and ferromagnetic parts of the Gibbs energy in Eqs. (15) and (16) are temperature dependent, their values in that specific range only are considered in the calculation. The results of this analysis for the effect of alloying elements on the Ms temperature of Fe are shown in Table 2. They can also be expressed by Ms (C, a t . % ) = 5 4 5 - 7 1 C + A 1 + 7 C o - 14Cr - 15Cu - 23Mn - 8Mo - 6Nb - 13Ni - 4Si + 3Ti - 4V + 0W Ms (C, w t . % ) = 5 4 5 - 3 3 0 C + 2AI + 7Co - 14Cr- 13Cu - 23Mn - 5Mo - 4Nb - 13Ni - 7Si + 3Ti + 4V + 0W (21) (20)

where the Ms temperature of pure Fe (545 C) is taken from the value reported by Gilbert and Owen [24]. Table 3 shows a comparison of the calculated and observed changes [1-7] in Ms temperature. It can be seen that the agreement is quite satisfactory and confirms the validity of the thermodynamic approach to predicting Ms temperatures in steels.

If. lshida / Journal o f Alloys and Compounds 220 (1995) 126-131

131

Acknowledgements The author would like to thank Professor A.P. Miodownik of the University of Surrey for helpful discussions. He would also like to thank Dr. L. Chandrasekaren of the University of Surrey for help in preparation of this manuscript.

References
[1] P. Payson and C.H. Savage, Trans. Am. Soc. Met., 33 (1944) 261. [2] R.A. Grange and H.M. Stevert, Trans. AIME, 167 (1946) 467. [3] A.E. Nehrenberg, Trans. AIME, 167 (1946) 494. [4] E.S. Rowland and S.R. Lyle, Trans. Am. Soc. Met., 37 (1946) 27. [5] J.H. Hollomon and L.D. Jaffe, Ferrous Metallurgical Design, Wiley, New York, 1947, p. 47. [6] W. ;~teven and A.G. Haynes, J. Iron Steel Inst., 183 (1956) 349. [7] ICW. Andrews, J. Iron Steel Inst., 203 (1965) 721. [8] C.Y. Kung and J.J. Payment, Metall. Trans. A, 13 (1982) 328. [9] C. ;Sener, J. Met., 7 (1955) 619.

[10] L. Kaufman and M. Cohen, Prog. Met. Phys., 7 (1958) 165. [11] M. Hillert, T. Wada and H. Wada, J. Iron Steel Inst., 205 (1967) 539. [12] K. Ishida and T. Nishizawa, Trans. Jpn. Inst. Met., 15 (1974) 217. [13] K. lshida, H. Wakakuwa and T. Nishizawa, in J.M. Gray, T. Ko, Z.S. Wu B. and Xie X. (eds.), HSLA Steels; Metallurgy and Applications, ASM International, Metals Park, OH, 1987, p. 851. [14] T. Nishizawa, S.M. Hao, M. Hasebe and K. Ishida, Acta Metall., 31 (1983) 1403. [15] R.L. Orr and J. Chipman, Trans. Met. Soc. AIME, 239 (1967) 630. [16] M. Hillert and M. Jarl, Calphad, 3 (1978) 227. [17] Y.Y. Chuang, R. Schmid and Y.A. Chang, Metall. Trans. A, 16 (1985) 153. [18] J. Friedel, Dislocations, Addison-Wesley, Reading, MA, 1964, p. 189. [19] C.L. Magee, Ph.D. Thesis, Carnegie Institute of Technology, Pittsburgh, PA, 1966. [20] E.M. Breinan and G.S. Ansell, Metall. Trans., 1 (1970) 1513. [21] K. Ishida, Scr. Metall., 11 (1977) 237. [22] A.J. Bogers and W.G. Burgres, Acta Metall., 12 (1964) 255. [23] K.J. Irvine, T. Gladman and F.B. Pickering, J. Iron Steel Inst., 203 (1969) 1017. [24] A. Gilbert and W.S. Owen, Acta Metall., 10 (1962) 45.

You might also like