2 Physico-Chemical Principles of Steelmaking Processes: Heterogeneous Vs Homogeneous Systems

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

2

Physico-Chemical Principles of Steelmaking Processes

2.1 INTRODUCTION

Ironmaking and steelmaking involve


• a variety of chemical reactions, and
• other physicochemical processes, such as viscous flow, interfacial phenomena, mass transfer,
etc. at high temperatures.

Physical chemistry constitutes the most important scientific fundamental for understanding the
metallurgical aspects of ironmaking and steelmaking.
• of which chemical thermodynamics, consisting of reaction free energies and enthalpies,
chemical equilibria, has found the most widespread application.

2.2 THERMODYNAMICS OF STEELMAKING PROCESS

2.2.1 Principal Concepts of Physical Chemistry Useful in Steelmaking

A system is a group of bodies which interact with one another and can be isolated, either mentally or
physically from the surroundings
• metal-slag systems, refractory lining-metal systems

In practice, one most often must deal with complex systems


• ‘lining-metal-slag-furnace atmosphere’ system

Real systems are too complicated to be analysed easily


• usually subdivided into several simpler systems, e.g. ‘metal-slag’, slag-furnace atmosphere’ etc.

Heterogeneous vs homogeneous systems


• ‘metal’, ‘slag’, ‘furnace atmosphere’, etc.
In practice, they are far from homogeneous
• metal may contain non-metallic inclusions and gas bubbles
• slag may contain undissolved slag formers, metal beads, and gas bubbles
• furnace atmosphere may carry suspended slag particles (melting dust) and sometimes, metal
particles.

Uniform vs non-uniform system


• identical composition (a chemical uniform system) or identical properties (a physically uniform
system)
• in the real steelmaking processes, the composition and the properties of various portions of a
system are mostly not uniform
Rashid . Ferrous Production Metallurgy

The state of a system is characterised by the parameters of state (P, T, X, etc.)


A process is characterised by process parameters (Q and W)
Steelmaking processes are essentially isobaric (H = QP)
Three common parameters to consider G, H and Q

For enthalpy change, absolute value of enthalpy H are still non-existent, and instead, H, is considered
• enthalpy of simple substances (elements) at 298 K is zero (𝐻298 𝑜
= 0)
• entropies of all substances at absolute zero are equal to zero, i.e., 𝑆𝑜 = 0 at T = 0 K.

The relations between the principal thermodynamic functions:


𝐻 = 𝑈 + 𝑃𝑉
𝐺 = 𝐻 − 𝑇𝑆
Free energy change at temperature T is determined by the equation
∆𝐺𝑇 = ∆𝐻𝑇 − 𝑇∆𝑆𝑇
For practical calculations, changes under standard conditions (1 atm P and 298 K T) are considered:
𝑜 𝑜 𝑜
∆𝐺298 , ∆𝐻298 , ∆𝑆298

Steelmaking processes take place high temperatures and change of H with T is calculated as
𝑇
𝑜
∆𝐻𝑇 = ∆𝐻298 + ∫ 𝑛∆𝑐𝑃 𝑑𝑇
298
n = number of moles, cP = = specific heat capacity
𝑐𝑃 = 𝑎 + 𝑏𝑇 + 𝑐𝑇 2 + 𝑑𝑇 −1 + 𝑒𝑇 −2
𝑟𝑝
∆𝑐𝑃 = 𝑐𝑃 − 𝑐𝑃𝑠𝑟
The coefficients a, b, c, etc. can be found in tables.

When state of a component changes (an allotropic transformation, melting, vaporisation):


𝑇𝑡𝑟 𝑇
𝑜
∆𝐻𝑇 = ∆𝐻298 + ∫ 𝑛∆𝑐𝑃′ 𝑑𝑇 ± 𝐿𝑡𝑟 + ∫ 𝑛∆𝑐𝑃′′ 𝑑𝑇
298 𝑇𝑡𝑟
Ltr, = latent heat’ of the transformation

Similar treatment can be used to determine entropy change of reaction:


𝑇𝑡𝑟 𝑇
𝑜
𝑛∆𝑐𝑃′ 𝐿𝑡𝑟 𝑛∆𝑐𝑃′′
∆𝑆𝑇 = ∆𝑆298 + ∫ 𝑑𝑇 ± + ∫ 𝑑𝑇
𝑇 𝑇𝑡𝑟 𝑇
298 𝑇𝑡𝑟

If there are no transformations in the given temperature range, G can be found as follows:
𝑜 𝑜 𝑜
∆𝐺298 = ∆𝐻298 − 𝑇∆𝑆298
If the specific heat changes only insignificantly with temperature (cP = 0):
∆𝐺𝑇𝑜 = ∆𝐻298
𝑜 𝑜
− 𝑇∆𝑆298

Considering the effect of temperature on H and S, the effect of temperature on G is:
∆𝐺 𝑜 = 𝑎 + 𝑏𝑇 + 𝑐𝑇 log 𝑇
𝑜
In most cases, the effect of T on ∆𝐺 can be approximated, with a good accuracy, by a straight line:
∆𝐺 𝑜 = ∆𝐻𝑜 − 𝑇∆𝑆 𝑜 = 𝐴 + 𝐵𝑇
𝑜 𝑜
where ∆𝐻 and ∆𝑆 are taken as the average values in a given temperature interval.

20
2. Physico-Chemical Principles of Steelmaking Processes

Based on nature of entropy change, the following generalisation of G with T can be made:

1. For volume expansion accompanying a reaction (gas evolution), S at constant P and T is positive,
hence, G decreases with an increasing T.
C + CO2 = 2CO; G° = 166,560 – 171.0 T Joules

2. For volume contraction, i.e. gas consumed in the reaction, S at constant P and T is negative, hence,
G increases with an increasing T.
H2 + 1⁄2S2 = H2S; G° = –91,600 + 50.6 T Joules

3. For little or no volume change, S  0, hence temperature has little effect on G.
C + O2 = CO2; G° = 395,300 – 0.5 T Joules

Most ironmaking and steelmaking systems are chemically reactive systems

The standard free energy change is directly associated with the equilibrium constant, Ke of the reaction:
∆𝐺 0 = −𝑅𝑇 ln 𝐾𝑒
where, for the chemical reaction
𝑏𝐵 + 𝑑𝐷 = 𝑚𝑀 + 𝑛𝑁
the equilibrium constant can be expressed as
(𝑎𝑀 )𝑚 (𝑎𝑁 )𝑛
𝐾𝑒 = [ ]
(𝑎𝐵 )𝑏 (𝑎𝐷 )𝑑 𝑒

The free energy change of any chemical reaction at any given condition:
∆𝐺 = ∆𝐺 𝑜 + 𝑅𝑇 ln 𝑄
here Q = activity quotient and is defined as
(𝑎𝑀 )𝑚 (𝑎𝑁 )𝑛
𝑄= [ ]
(𝑎𝐵 )𝑏 (𝑎𝐷 )𝑑 𝑛𝑒

Replacement of the value of ∆𝐺 𝑜 yields the value of ∆𝐺, the molar Gibbs free energy change of a reaction
in the given temperature and pressure
𝑄
∆𝐺 = −𝑅𝑇 ln 𝐾𝑒 + 𝑅𝑇 ln 𝑄 = 𝑅𝑇 ln ( )
𝐾𝑒
This relation is known as the van’t Hoff Isotherm, and can be used to determine thermodynamic
possibility of a reaction.
• If ∆𝐺 is positive, the reaction would not occur.

21
Rashid . Ferrous Production Metallurgy

Example 2.1
Calculate the ratio of partial pressures of CO2 and CO gas in a CO-CO2 gas mixture at equilibrium
with pO2 = 10–6 atm at 1600 °C. Calculate the chemical potential of oxygen and predict whether solid
FeO can be reduced to metallic iron by this gas mixture at a temperature of 1000 °C.

Solution:
For the chemical reaction
(𝐶𝑂) + 1/2(𝑂2 ) = (𝐶𝑂2 )
the free energy of formation
𝑝𝐶𝑂 1
∆𝐺𝑓𝑜 = −𝑅𝑇 ln 𝐾𝑒 = −𝑅𝑇 ln [ 2 x ]
𝑝𝐶𝑂 (𝑝 )1/2
𝑂2 𝑒

Now, combining the free energy of formations of CO and CO 2, we get the free energy of the above
reaction at 1873 K is -119,805 cal/mol. Then
𝑝𝐶𝑂 1
−119805 = −(8.314)(1873) ln [ 2 𝑥 ]
𝑝𝐶𝑂 (10 )1/2 𝑒
−6
𝑝𝐶𝑂2
= 2.194
𝑝𝐶𝑂
Now the chemical potential of oxygen is
𝜇𝑂2 = 𝑅𝑇 ln 𝑝𝑂2 = (8.314)(1873) ln(10−6 ) = −215.1 kJ/mol
Now the reduction reaction of FeO with the CO-CO2 gas mixture can be written as
< 𝐹𝑒𝑂 > +(𝐶𝑂) =< 𝐹𝑒 > +(𝐶𝑂2 ); ∆𝐺 𝑜 = ∆𝐺𝑓𝑜 (𝐶𝑂2 ) − ∆𝐺𝑓𝑜 (𝐹𝑒𝑂) − ∆𝐺𝑓𝑜 (𝐶𝑂)
Using data from Table 2.1, ∆𝐺 0 for the reaction at 1273 K is
J 𝑝𝐶𝑂
∆𝐺 𝑜 = 8081.41 = −𝑅𝑇 ln ( 2 )
mol 𝑝𝐶𝑂 𝑒
This gives (𝑝𝐶𝑂2 /𝑝𝐶𝑂 )e = 0.466, which is lower than the (𝑝𝐶𝑂2 /𝑝𝐶𝑂 )a of 2.194. Thus, the reaction is not
possible.

Reactions in the steelmaking processes can be generalised, using 1 mol of oxygen, as


2𝑥 2
𝑀 + 𝑂2 = 𝑀𝑥 𝑂𝑦
𝑦 𝑦
Examples are:
<C> + (O2 ) = (CO2 )
<Si> + (O2 ) = <SiO2 >
4 2
{Al} + (O2 ) = <Al2 O3 >
3 3

For determining equilibrium and stability of the resultant oxide formed by these reactions, the standard
free energy of formation ∆𝐺𝑓𝑜 of oxides per mole of oxygen can be calculated.
∆𝐺𝑓𝑜 = ∆𝐻𝑜 − 𝑇∆𝑆 𝑜 = 𝐴 + 𝐵𝑇

Figure 2.1 (Ellingham diagram)

Table 2.1 (Data for standard free energies of reactions encountered in ferrous metallurgical processes.

22
2. Physico-Chemical Principles of Steelmaking Processes

2.2.2 Thermodynamics of Solution

Steelmaking process consists of one or more homogeneous parts or phases and frequently each phase is
not a pure substance
• solution – perfectly homogeneous mixture of two or more substances whose concentrations in
the mixture may vary continuously within definite ‘solubility’ limits.

All practical steel-making processes involve the interaction between solutions either chemically or
physically with one another
• reaction between liquid metal and liquid slag, solid additions, and gas bubbles

Formation of a solution
• Some cases, involve no chemical reactions and one component of a solution is simply ‘diluted’
by another
• In other cases, the solvent and solute may react chemically with each other with the evolution
or absorption of heat. THIS IS IMPORTANT.

Example: much heat is liberated on dissolution of silicon in iron


• practically utilised in the manufacture of steel grades high in silicon (dynamo, silicon and other
high alloy steels).
• high-silicon ferro-alloys charged in the ladle in the solid state without running the risk of cooling
or even ‘chilling’ the tapped metal

2.2.2.1 Ideal and Non-Ideal Behaviour

Ideal solution
• forces of interaction between like and unlike atoms are equal
• chemical or physical behaviour of the solute i is proportional to its atomic or molecular
concentration in the solution.

At steelmaking temperatures, ideal behaviour is followed by


• solutions of gases
• metallic solutions such as Fe-Ni, Fe-Mn, Fe-Co
• many ionic melts containing a common ion such FeO-MnO, FeS-FeO, Fe2SiO4-Mg2SiO4.

If the solution were ideal, the vapour pressure would be proportional to the concentration of i, i.e.
𝑝𝑖 = 𝑝𝑖0 𝑋𝑖

If the solution were not ideal,


𝑝𝑖 = 𝑝𝑖0 𝑎𝑖

𝑎𝑖 = Raoultian activity
𝑝𝑖0 = proportionality constant = the vapour pressure of pure i at the same T , 𝑋𝑖 = the atom fraction of i

Standard State (where activity is unit)


• Raoultian standard state of i is that of the pure element at the same temperature and in the same
physical condition – solid, liquid or gaseous – as in the solution.

For ideal solution, 𝑎𝑖 = 𝑋𝑖 ( Raoult’s Law)


• Solutions showing ideal behaviour are often referred to as Raoultian
• Solutions of nickel, cobalt, manganese and chromium in iron are almost ideal.

23
Rashid . Ferrous Production Metallurgy

Deviation from ideality

Positive deviation from ideality (activities greater than those indicated by Raoultian behaviour)
• the system is observed to expand (V > 0) and to absorb heat (H > 0)
• copper in iron (strong Fe-Fe and Cu-Cu bonding and a weak Fe-Cu attraction)
• the tendency is so powerful that a separation into an iron-rich and a copper-rich phase occurs

Negative deviation from ideality (activities lower than those indicated by Raoultian behaviour)
• the evolution of heat (H < 0) with a decrease of its volume (V < 0)
• strong Fe-Si attractions, giving a tendency for Si atoms to have Fe atoms for nearest neighbours
(‘ordering’ behaviour)
• ultimately results in the formation of the solid-state compound FeSi when the liquid cools.

Figure 2.2: Activities in liquid Fe–Si and Fe–Cu alloys at 1600°C showing strong negative
and positive departures from Raoult’s law.

A measure of the deviation from ideality of non-ideal solutions;


𝑎𝑖
𝛾𝑖 =
𝑋𝑖
𝛾𝑖 = Raoultian Activity Coefficient ( = 1 for ideal solutions, ( < 1 for negative deviations, ( > 1 for
positive deviations)

Analytical correlations between activity and compositions for non-ideal solutions

The simplest is the regular solution model.

For a binary solution with components A and B, it predicts


𝑅𝑇 ln 𝛾𝐴 = 𝛼 𝑋𝐵2 ; 𝑅𝑇 ln 𝛾𝐵 = 𝛼 𝑋𝐴2
 = constant at constant temperature, 𝑋𝐴 + 𝑋𝐵 = 1.

Molten slags are concentrated solutions


• minimum concentration of any component oxide is more than a few weight percent
• activity-composition relation in binary slag systems can be represented as shown in Fig. 2.2 for
binary alloy systems
• Mole fraction is the composition scale, activity and activity coefficients defined with Raoult’s
law as the reference state (i.e. pure components)

24
2. Physico-Chemical Principles of Steelmaking Processes

Industrial slags are, however, multi-component solutions


• cannot be represented in the same manner as in Fig. 2.2
• ternary CaO-SiO2-Al2O3 constitutes the basis for blast furnace slags, and some slags
encountered in steelmaking processes

Figure 2.3 (activity of SiO2 in the ternary CaO-SiO2-Al2O3 system at 1550 °C)
• iso-activity lines for SiO2, CaO and Al2O3 (for other T, separate diagrams are required)
• liquid field (i.e., the molten slag field) is bounded by the liquidus lines
• Al2O3 is written as AlO1.5, because the molecular masses of CaO, SiO2, and AlO1.5 are close,
being equal to 56, 60 and 51, respectively. Therefore, the mole fraction scale becomes
approximately the same as the weight fraction scale.

Example 2.2
Determine the activity of SiO2 in the CaO-SiO2-AlO1.5 system at 1823 K at XCaO = 0.5 and XSiO2 = 0.4.
Consider the equilibrium of this slag with an Fe-Si liquid alloy and a gas mixture of CO-CO2 at this
temperature, Calculate (𝑝𝐶𝑂2 /𝑝𝐶𝑂 ) ratio in the gas mixture at equilibrium.
Given data: In the alloy, XSi = 0.02, Si = 1.6*10–3.

Solution:

< 𝑆𝑖 > + 2(𝐶𝑂2 ) = {𝑆𝑖𝑂2 } + 2(𝐶𝑂); ∆𝐺 𝑜 = ∆𝐺𝑓𝑜 {SiO2 } + 2∆𝐺𝑓𝑜 (CO) − 2∆𝐺𝑓𝑜 (CO2 )
From data of Table 2.1, at 1823 K, ∆𝐺 𝑜 = -337,179 J/mol. Hence
2
9
𝑝𝐶𝑂 𝑎𝑆𝑖𝑂2
𝐾 = 4.6 ∗ 10 = [( ) x ]
𝑝𝐶𝑂2 𝑎𝑆𝑖
𝑒
From Fig. 2.3, at XCaO = 0.5 and XSiO2 = 0.4, 𝑎𝑆𝑖𝑂2 in slag = 0.1.
Now, the activity of silicon in the alloy
𝑎𝑆𝑖 = 𝑋𝑆𝑖 𝛾𝑆𝑖 = 0.02 (1.6 ∗ 10−3 ) = 3.2 ∗ 10−5
Substituting the values, 𝑝𝐶𝑂2 /𝑝𝐶𝑂 = 8.24*10-4 .

25
Rashid . Ferrous Production Metallurgy

2.2.2.2 Dilute Solutions

Liquid steels fall in the category of dilute solutions


• concentrations of solutes (C, O, Si, Mn, S, P, etc.) are mostly below 1 wt% or so
• except for high-alloy steels

Solvent (Fe) atoms


• XFe  1
• show approximately ideal behaviour
• 𝑎𝑖 = 𝑋𝑖

Solute atoms
• mole fraction approaches zero (Xk ➔ 0)
• activity varies linearly with mole fraction (Henry’s Law)
𝑎𝑖 = 𝛾𝑖0 𝑋𝑖
0
𝛾𝑖 = constant of proportionality (Henry’s Law coefficient), independent of composition for low
concentration of solute.

Deviation from Henry’s law occurs when the solute concentration increases (Fig. 2.4)

Figure 2.4: Activity of carbon in austenite (relative to graphite) at l000 °C,


demonstrating deviation from Henry's law.

Activities in liquid steel are expressed with reference to Henry’s law, and not Raoult’s law
• the composition scale is weight percent, not mole fraction.

With this modification, if the solute i, in solution with Fe as solvent in the binary Fe-i solution,
1. obeys Henry’s law, then
ℎ𝑖 = 𝑤𝑡. %𝑖
2. does not obey Henry’s law, then
ℎ𝑖 = 𝑓𝑖 x (𝑤𝑡. %𝑖)
ℎ𝑖 = Henrian activity, 𝑓𝑖 = Henrian Activity Coefficient of solute i (in one weight percent standard state)

For 1wt% i, ℎ𝑖 = 1 (Henry’s law is obeyed)

26
2. Physico-Chemical Principles of Steelmaking Processes

Conversion of Raoultian activities to the Henrian form (Fig. 2.5)

Figure 2.5: Diagrammatic illustration of the relationship between Henrian and Raoultian activities
in a binary system with moderate negative deviations from ideal Raoultian behaviour.

ℎ𝑖 𝑎𝑖 𝛾𝑖
𝑓𝑖 = = 0 = 0
𝑤𝑡%𝑖 𝛾𝑖 𝑋𝑖 𝛾𝑖
Rearrangement of this equation gives:
ℎ𝑖 𝛾𝑖0 𝑋𝑖
𝑎𝑖 =
𝑤𝑡%𝑖
But
𝑤𝑡%𝑖
𝑀𝑖
𝑋𝑖 =
𝑤𝑡%𝑖 𝑤𝑡%𝐹𝑒
+
𝑀𝑖 𝑀𝐹𝑒
which for wt%i →0 (and hence wt%Fe → 100) simplifies to:
𝑤𝑡%𝑖 𝑀𝐹𝑒
𝑋𝑖 =
100 𝑀𝑖
Then, by inserting the molecular weight of iron, 𝑀𝐹𝑒 = 56, the Henrian activity simplifies to:
0.56 ℎ𝑖 𝛾𝑖0
𝑎𝑖 =
𝑀𝑖

By using this relation, any thermodynamic expression initially derived in terms of Raoultian activities,
can easily be converted for use with the more convenient Henrian activities
• For example, the free energy change of a pure substance on solution at an activity 𝑎𝑖 is:
0.56 𝑓𝑖 𝑤𝑡%𝑖 𝛾𝑖0
∆𝐺 = 𝑅𝑇 ln 𝑎𝑖 = 4.575 𝑇 log ( )
𝑀𝑖
e.g. ∆𝐺 for the formation of the Henrian reference solution (1 wt% solution, 𝑓𝑖 = 1, is:) is:
∆𝐺 = 4.575 𝑇 log(0.56 𝛾𝑖0 /𝑀𝑖 )

27
Rashid . Ferrous Production Metallurgy

2.2.2.3 Interaction Between Solutes in Dilute Solutions

Molten steel and iron have several dissolved solutes in dilute state (e.g., C, Mn, Si, S, P, etc)
• solute in a multi-component solution interact with one another and thus influence activities of
other solutes.
• activity coefficient of a solute B in solution is altered by the addition of other solutes, even in
dilute solutions

The Henrian activity coefficient of solute B, fB, in solution containing other solutes C, D, ..., i, etc. is give
approximately as
𝑓𝐵 = 𝑓𝐵𝐵 𝑓𝐵𝐶 𝑓𝐵𝐷 … . 𝑓𝐵𝑖
𝐵
𝑓𝐵 = the activity coefficient expressing the effect of the concentration of B on the activity of B in the
simple binary A-B solution and defined as
ℎ𝐵𝐴−𝐵
𝑓𝐵𝐵 =
𝑤𝑡%𝐵
𝑓𝐵𝑖 = the activity coefficient expressing the effect of i on the activity of B in the A-B-i ternary solutions
and defined as
ℎ𝐵𝐴−𝐵−𝑖
𝑓𝐵𝑖 =
𝑤𝑡%𝐵

• equation is an approximate one as it does not consider the interactions amongst C, D, ..., i in
solution

Activity of B in an A-B-C-D-...-i multi-component dilute solution can be expressed as


ℎ𝐵 = 𝑓𝐵 𝑤𝑡%𝐵 = (𝑓𝐵𝐵 𝑓𝐵𝐶 𝑓𝐵𝐷 … 𝑓𝐵𝑖 ) 𝑤𝑡%𝐵

In ternary solutions, the effect of a second solute element C on the first solute B usually follows a
logarithmic law for low concentrations of C, and the behaviour can be summarised by equations of the
type
log 𝑓𝐵𝐶 = 𝑒𝐵𝐶 𝑤𝑡%𝐶
𝐶
𝑒𝐵 = constant and known as the interaction coefficient. It expresses the influence of solute C on the
activity coefficient of solute B in the ternary solution.

Logarithmic form of the activity equation


log ℎ𝐵 = log 𝑓𝐵𝐵 + (𝑒𝐵𝐶 𝑤𝑡%𝐶) + (𝑒𝐵𝐷 𝑤𝑡%𝐷) + ⋯ + (𝑒𝐵𝑖 𝑤𝑡%𝑖) + log(𝑤𝑡%𝐵)

If one interaction coefficient is known, the other can be obtained by the following relation:
𝑀𝐵 𝑀𝐵
𝑒𝐵𝐶 = 𝑒𝐶𝐵 ( ) + 434 ∗ 10−4 (1 − )
𝑀𝐶 𝑀𝐶
where M is the molecular weight.

This equation is strictly correct for infinitely dilute solutions.

Table 2.2 (interaction coefficients of solute in liquid iron at 1600 °C)

28
2. Physico-Chemical Principles of Steelmaking Processes

Example 2.3
An Fe-Mn solid solution having 0.001 mole fraction of Mn, is in equilibrium with an FeO-MnO solid
solution in an oxygen containing atmosphere at 1000 K. Calculate the composition of the oxide
solution. Both metallic and oxide solutions are ideal.
Given data: <Mn> + 1/2(O2) = <MnO> ; ∆𝐺𝑓𝑜 = -384700 + 72.8 T J/mol

Solution:
<Fe> + <MnO> = <Mn> + <FeO>; ∆𝐺 𝑜 = ∆𝐺𝑓𝑜 (FeO) – ∆𝐺𝑓𝑜 (MnO)

Using the given data and Table 2.1, at 1000 K


∆𝐺 𝑜 = 114850 J/mol
𝑎𝑀𝑛 𝑎𝐹𝑒𝑂 𝑋𝑀𝑛 𝑋𝐹𝑒𝑂 (0.001)𝑋𝐹𝑒𝑂 −114850
𝐾= = = = exp ( )
𝑎𝐹𝑒 𝑎𝑀𝑛𝑂 𝑋𝐹𝑒 𝑋𝑀𝑛𝑂 (1.0)𝑋𝑀𝑛𝑂 8.314 x 1000
(0.001)𝑋𝐹𝑒𝑂
= 10−6
(1.0)𝑋𝑀𝑛𝑂
𝑋𝐹𝑒𝑂 = 103 𝑋𝑀𝑛𝑂

Noting that 𝑎𝐹𝑒𝑂 = 𝑋𝐹𝑒𝑂 , 𝑎𝑀𝑛𝑂 = 𝑋𝑀𝑛𝑂 , and 𝑋𝐹𝑒𝑂 + 𝑋𝑀𝑛𝑂 = 1, calculation yields

𝑋𝐹𝑒𝑂 = 0.001 and 𝑋𝑀𝑛𝑂 = 0.999

Example 2.4
Calculate the chemical potential of nitrogen gas at equilibrium with liquid steel at 1600 °C. The steel
contains 0.01% N, 0.5% C, 0.5%Mn.
1/2 581
Given data: [𝑁]𝑝𝑝𝑚 𝑠𝑡𝑑 𝑠𝑡𝑎𝑡𝑒 𝑖𝑛 𝑖𝑟𝑜𝑛 = 𝐾𝑁 x 𝑝𝑁2 where log 𝐾𝑁 = − + 2.937
𝑇
4
(Note: 1 wt% = 10 ppm).

Solution
At equilibrium, 𝜇𝑁2 (gas) = 𝜇𝑁2 (steel) = 𝑅𝑇 ln 𝑝𝑁2
From the given data
log 𝑝𝑁2 = 2 [log ℎ𝑁 − log 𝐾𝑁 ], where ℎ𝑁 is in ppm
Now at 1873 K, log 𝐾𝑁 = 2.6268. Again,
ℎ𝑁 = 𝑓𝑁 (𝑤𝑡. %𝑀𝑛)
log ℎ𝑁 = log[𝑁]𝑝𝑝𝑚 + 𝑒𝑁𝑁 (𝑤𝑡. % 𝑁) + 𝑒𝑁𝐶 (𝑤𝑡. % 𝐶) + 𝑒𝑁𝑀𝑛 (𝑤𝑡. %𝑀𝑛)
Substituting values of interaction coefficients from Table 2.2,
log ℎ𝑁 = log(0.01 x 104 ) + (0)(0.01) + (0.13)(0.5) + (−0.02)(0.5)
log ℎ𝑁 = 2.055
Thus,
log 𝑝𝑁2 = −3.1436
𝜇𝑁2 = 2.303 (8.314)(1873)(−3.1436)(103 ) = −112.7 kJ/mol 𝑁2

29

You might also like