17.07.2013 Views

Modeling and Simulation for Material Selection and Mechanical ...

Modeling and Simulation for Material Selection and Mechanical ...

Modeling and Simulation for Material Selection and Mechanical ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

2<br />

Design <strong>Simulation</strong> of Kinetics<br />

of Multicomponent Grain<br />

Boundary Segregations in the<br />

Engineering Steels Under<br />

Quenching <strong>and</strong> Tempering<br />

Anatoli Kovalev <strong>and</strong> Dmitry L. Wainstein<br />

Physical Metallurgy Institute, Moscow, Russia<br />

I. INTRODUCTION<br />

The basic factors controlling grain boundary segregations (GBS) in engineering<br />

steels are discussed. In contrast to single-phase alloys, in engineering<br />

steels, the multicomponent segregation is developed simultaneously<br />

with undercooled austenite trans<strong>for</strong>mations <strong>and</strong> martensite decomposition.<br />

Based on these reasons, the influence of steel phase composition <strong>and</strong><br />

kinetics on concurrent segregations is discussed. It is established that grain<br />

boundary enrichment by harmful impurities (S <strong>and</strong> P) is possible after carbon<br />

<strong>and</strong> nitrogen segregation dissolution. Two models of GBS are<br />

described. The dynamic model of segregation during quenching is based<br />

on the solution of independent diffusion <strong>and</strong> adsorption–desorption equations<br />

<strong>for</strong> various impurities in steel. The model of multicomponent segregation<br />

under tempering considers the influence of alloying <strong>and</strong> tempering<br />

parameters on concentration <strong>and</strong> thermodynamic activity of carbon in<br />

the a-solid solution.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


II. GRAIN BOUNDARY SEGREGATION AND PROPERTIES<br />

OF ENGINEERING STEELS<br />

Chemical composition <strong>and</strong> structure of the grain boundary influences various<br />

properties of engineering steels. The following are some of these properties:<br />

inclination to temper <strong>and</strong> heat brittleness, resistance to hydrogen<br />

embrittlement, corrosion, delayed fracture, <strong>and</strong> creep. The intergranular<br />

fracture is the main reason <strong>for</strong> decrease of many steel exploitation properties.<br />

Application of modern physical experimental or calculation methods<br />

has successfully helped in solving the old metallurgical problem of intergranular<br />

fracture.<br />

The affinity of various kinds of intergranular brittleness is associated<br />

with two main unfavorable factors that decrease intergranular bonds. The<br />

impurity segregation to grain boundaries (GBS) <strong>and</strong> localization of internal<br />

microstresses are necessary <strong>and</strong> sufficient conditions that could initiate<br />

embrittlement [1].<br />

Despite the common features, certain kinds of steel brittleness are distinguishable<br />

from each other <strong>and</strong> are stipulated by complex interaction of<br />

these factors. The concentration of internal stresses on grain boundaries<br />

could be an effect of martensite trans<strong>for</strong>mation, hydrogen accumulation,<br />

or carbide precipitation; <strong>and</strong> grain boundary segregations could appear during<br />

the equilibrium or non-equilibrium processes of element redistribution in<br />

steel.<br />

The concentration of microstresses on grain boundaries cause the<br />

initiation of cracking <strong>and</strong> acts as the primary reason <strong>for</strong> brittleness. The<br />

enrichment of grain boundaries by harmful impurities is a major <strong>and</strong> common<br />

condition <strong>for</strong> development of various intercrystalline brittleness phenomena<br />

<strong>and</strong> it specifies crack propagation entirely along grain boundaries<br />

at low stresses.<br />

The concept of intercrystalline internal adsorption [2] that was confirmed<br />

by theoretical [3], <strong>and</strong> experimental work [4], the thermodynamic analysis<br />

of chemical element interaction during equilibrium grain boundary<br />

segregation [5], <strong>and</strong> investigations of quenched <strong>and</strong> tempered steel [6,7]. This<br />

made it possible to interpret the tempering embrittlement phenomenon.<br />

Elemental impurities enrich grain boundaries in thin layers up to<br />

several atoms <strong>and</strong> change the type <strong>and</strong> value of interatomic bonds that<br />

lead to intercrystalline fracture. Embrittlement power is commonly<br />

attributed to the elements of the 3rd to 5th periods of groups IV to<br />

VI in the periodic system [1]. Sulfur, phosphorus, arsenic, selenium, tellurium,<br />

antimony, bismuth, <strong>and</strong> oxygen are the most harmful impurities<br />

that segregate in grain boundaries. The concentration in grain boundaries<br />

could reach several atomic percentages exceeding the volume one<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


y several hundreds. Intercrystalline brittleness, as caused by GBS, due<br />

to harmful impurities is observed, as a rule, in BCC metals <strong>and</strong><br />

alloys. The austenite alloys are significantly more resistant to this kind<br />

of fracture.<br />

The motonic increase of plasticity that is expected after martensite<br />

decomposition due to tempering of engineering steels is disturbed by two<br />

anomalies resulting in a relative decrease of impact strength. These anomalies<br />

are accompanied by intergranular fracture. Steels may be susceptible to<br />

embrittlement when they are heated <strong>for</strong> prolonged period in the temperature<br />

range 350–5508C, or when slowly cooled through it. Depending on the heattreatment<br />

cycle, the phenomenon is called either temper embrittlement (3508<br />

embrittlement) or reversible temper embrittlement (5508 embrittlement).<br />

Common indications of embrittlement are a loss of toughness, segregation<br />

of harmful impurities to grain boundaries <strong>and</strong> the fracture path usually<br />

along prior austenite grain boundaries, <strong>and</strong> the impact transition temperature<br />

(FATT—the fracture appearance transition temperature) which is displaced<br />

towards higher values.<br />

The irreversible temper embrittlement of low-alloyed steels (


Figure 1 Impact strength at room temperature of samples with V-shape cut <strong>for</strong><br />

several melts of industrial steel 4340 after quenching (8508C, 1 hr) in oil <strong>and</strong> 1 hr<br />

tempering at various temperatures. (From Ref. 8.)<br />

Figure 2 (a) Lamellar parts of cementite Fe 3C on primary austenite grains <strong>and</strong> (b)<br />

martensite packs boundaries. Steel 0.35C–1.5Mn–0.1P. Tempering at 3508C, 1 hr<br />

(TEM, replicas).<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 3 (a) Brittle fracture through primary austenite grains <strong>and</strong> (b) martensite<br />

pack boundaries in steel 0.35C–1.5Mn–0.1P. Tempering at 3508C, 1 hr (SEM).<br />

de<strong>for</strong>mation <strong>and</strong>, consequently, <strong>for</strong>mation of grain boundary cracks. The<br />

brittle crack develops, in this case, within boundaries of primary austenite<br />

grains <strong>and</strong> martensite packs (Fig. 3a, b). The significant enrichment of grain<br />

boundaries by phosphorus is confirmed by Auger spectroscopy. The decisive<br />

role of austenitization when compared with tempering in achieving equilibrium<br />

GBS is shown by low diffusion mobility of harmful impurities’ atoms<br />

(P, As, S, <strong>and</strong> Sb) <strong>for</strong> medium tempered steel. Calculations show that only<br />

nitrogen has significant diffusion mobility in this temperature range which is<br />

sufficient <strong>for</strong> diffusion at a distance of 10 mm <strong>for</strong> 1 hr at 3508C [1]. The<br />

temperature of steel <strong>for</strong> quenching is sufficiently high <strong>for</strong> intensive diffusion<br />

of P in austenite with the <strong>for</strong>mation of equilibrium GBS [9], <strong>and</strong> quenching<br />

fixes this enrichment. The level of P segregation depends on the austenitization<br />

temperature <strong>and</strong> increases when the temperature decreases below<br />

10508C. This is related to the decrease of P solubility in austenite. It is confirmed<br />

by a significant decrease of the quenched steel delay fracture resistance<br />

with respect to temperature in the austenite region (Fig. 4).<br />

Phosphorous content in steel influences its embrittlement at 3508C tempering.<br />

Its concentration in GBS is several hundred times higher than in the<br />

volume, <strong>and</strong> this harmful element segregates in austenite even at low concentrations,<br />

about 0.01% mass, leading to a significant decrease in the<br />

impact strength (Fig. 5).<br />

Change of the relative energy of grain boundaries results in segregation<br />

enrichment by impurities during austenitization. The interferometry is<br />

one such direct technique <strong>for</strong> grain boundary energy determination. For this<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 4 Delayed fracture diagram of quenched steel 0.35C; 1.5Mn; 0.1P after<br />

austenitization at 1423K, 3 hr <strong>and</strong> 30 min interim cooling to temperatures: 1, 1123K;<br />

2, 1223K; 3, 1273K <strong>and</strong> cooling in water. (From Ref. 10.)<br />

purpose, the opening angle of GB slot (y) is determined on metallographic<br />

grinds after vacuum etching at various temperatures. The grinds are austenitized<br />

in vacuum at high temperature, then subjected to interim cooling to<br />

the desired temperature <strong>and</strong> then quenched at a high cooling rate. The relative<br />

energy of GB is calculated by the equation<br />

gb ¼ 2cos<br />

gs y<br />

ð1Þ<br />

2<br />

where symbols b <strong>and</strong> s correspond to boundary <strong>and</strong> bulk, respectively.<br />

Figure 6 shows the change of relative surface energy with respect to the<br />

interim cooling temperature during austenitization <strong>and</strong> P content in the<br />

0.35% C, 1.5% Mn. For the steel with low phosphorus content, the GB<br />

energy regularly decreases slowly as the temperature increases within the<br />

austenite region. Significant segregation enrichment of grain boundaries<br />

by phosphorus in the steel with 0.1% P does not show this dependence<br />

<strong>and</strong> decreases the surface energy of GB with decreasing of the temperature.<br />

The decrease of P solubility in austenite with decreasing temperature assists<br />

its adsorption on the GB <strong>and</strong> decreases significantly its surface tension<br />

energy. At the same time, one can observe another process: P enriches<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 5 Temper embrittlement of steel Fe–0.3C–3.5N–1.7Cr: 1, high purity steel;<br />

2, 0.01 mass% P; 3, 0.03 mass% P; 4, 0.06 mass% P. (From Ref. 8.)<br />

non-metallic inclusions. The dependence of P content on Mn sulfides in steel<br />

0.35% C, 1.5% Mn, 0.1% P from austenitization temperature is shown in<br />

Fig. 7. When annealing temperature decreases, P redistributes between the<br />

bulk <strong>and</strong> grain boundaries enriching them <strong>and</strong> the non-metallic inclusions.<br />

Steel grain size reduction can decrease significantly its tendency to temper<br />

embrittlement. Increase of the specific surface of the GB at dispersion of<br />

the steel structure decreases the concentration of the impurity in the grain<br />

boundary that leads to growth of steel brittle fracture resistance (Fig. 8) [3].<br />

Phosphorous segregations decrease the surface energy of intergranular cohesion.<br />

Using the approach proposed in Ref. [11], one can estimate the role of<br />

phosphorus in change of surface energy of intergranular cohesion <strong>for</strong> development<br />

of this kind of embrittlement.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 6 Change of relative surface energy of grain boundaries in dependence of<br />

temperature of austenitization. Fe–0.35C–1.5Mn steel: 1, 0.03% P; 2, 0.1% P.<br />

Figure 7 Phosphorus concentration in manganese sulfides in 0.35C–1.5Mn–0.1P<br />

steel vs. austenitization temperature.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 8 FATT (T50) of steel vs. grain size d 1=2 <strong>and</strong> specific grain boundary surface<br />

P S. Phosphorus content: 1, 0.03%; 2, 0.1%. (DT50 ¼ T50(0.1% P) T50(0.03% P)).<br />

Influence of grain size on ductile–brittle transition temperature FATT<br />

is determined from the well known Petch–Hall equation<br />

sf ¼ so þ Ky d 1=2<br />

where Ky ffi aðGgÞ 1=2 ; a is a constant (1–3); g is the surface energy of intergranular<br />

cohesion <strong>for</strong> generation of cracks on grain boundaries; G is the<br />

shear modulus.<br />

It is described by the equation<br />

dT50<br />

d d 1=2<br />

1<br />

¼<br />

ð Þ b<br />

where b is determined as a tangent of inclination angle of straight lines in<br />

Fig. 8. Taking into account that<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.<br />

ð2Þ<br />

ð3Þ


dT50<br />

d d 1=2<br />

ð Þ<br />

dT<br />

¼ ð K<br />

dsf<br />

0 yÞ; ð4Þ<br />

it is possible to use the change of the inclination angle of lines 1 <strong>and</strong> 2 in<br />

Fig. 8 to determine the influence of P on GBS by the coefficient b <strong>and</strong><br />

the effective intergranular cohesion surface energy g that is proportional<br />

to b 2 . The values of b <strong>for</strong> steel samples containing 0.03% <strong>and</strong> 0.1% P at<br />

the temper embrittlement state are equal to 0.54 <strong>and</strong> 0.22 sec, respectively.<br />

Taking into account dependencies (2) <strong>and</strong> (3), one can conclude that effec-<br />

tive intergranular cohesion surface energy in steel with higher P concentra-<br />

¼ 6:04 times.<br />

tion decreases in g 0:03=g 0:1 ¼ b 2<br />

0:03 =b2<br />

0=1<br />

B. Ductile Intergranular Fracture of Overheated Steel<br />

The ductile fracture of steel as well as brittle fracture could be characterized<br />

by the lowest energy capacity. Such a fracture occurs when the inclusions are<br />

located along grain boundaries occupying a very large volume near the<br />

boundary. The intergranular microvoid fracture is observed in this case<br />

due to overheating of the steel. The samples are exposed to high temperature<br />

heating to dissolve inclusions. As a rule, the large <strong>and</strong> lamellar oxysulfides<br />

that did not embrittle steel are dissolved. After their dissolution, segregation<br />

of O, S, P, <strong>and</strong> precipitation, disperse particles during steel cooling occurs.<br />

In low-alloyed steels, precipitates could be sulfides of chromium <strong>and</strong> manganese<br />

(MnS, CrS), <strong>and</strong> aluminum nitride, AlN. These particles build a<br />

dense network on grain boundaries. The fracture occurs at higher or room<br />

temperatures by intergranular microvoid coalescence at low stress intensity.<br />

The micrometer scale cavities nucleate on the intergranular fine dispersion<br />

of sulfide or nitride particles (Fig. 9). The segregation of harmful impurities<br />

is observed on grain boundaries in this case.<br />

C. Reversible Temper Embrittlement<br />

The reversible temper embrittlement (RTE) is observed in engineering steel<br />

alloyed by carbide-<strong>for</strong>ming elements after quenching <strong>and</strong> high tempering<br />

(500–6008C). This phenomenon is developed in steels of industrial purity.<br />

It consists of a large decrease in the steel impact strength after slow cooling,<br />

but after rapid cooling at 6508C, the steel has a st<strong>and</strong>ard impact strength.<br />

The RTE phenomenon was identified <strong>for</strong> the first time in 1883 [12], when<br />

blacksmiths observed that some steels had to be water quenched after tempering,<br />

to avoid embrittlement. This decrease of impact strength is not<br />

accompanied by a change of physical or other mechanical properties of<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 9 Ductile fracture of steel through grain boundaries (SEM).<br />

steel. The duration of the exposure time within a definite range of tempering<br />

temperatures (brittleness zone) plays a decisive role in the development of<br />

RTE. The brittle fracture goes through the primary austenite grains<br />

(Fig. 10). The duration of the exposure time of the normalized or annealed<br />

steel within the dangerous temperature range also leads to this kind of<br />

embrittlement. Steels sensitive to RTE are subjected to rapid cooling from<br />

6508C during different heat treatments.<br />

The GBS of phosphorus is the main reason <strong>for</strong> this kind of brittleness.<br />

The RTE of alloyed steels is mainly sensitive to two factors: the chemical<br />

composition of the grain boundaries, <strong>and</strong> the mechanical <strong>and</strong> microstructural<br />

parameters of the alloy. The direct correspondence of embrittlement<br />

kinetic features <strong>and</strong> GBS of P at steel tempering has been established. Figure 11<br />

shows the iso-FATT curves <strong>for</strong> temper embrittlement of Ni–Cr steel [13].<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 10 SEM image of intergranular fracture of Fe–0.35C–1.5Mn–0.1P steel<br />

after quenching with temperature of 9508C, 1 hr <strong>and</strong> tempering at 5508C, 2 hr.<br />

At a constant temperature of embrittlement tempering, the FATT increases<br />

with time. The embrittlement is reversible. It can be rejected by short heating<br />

above the nose of the C-curve in the ferrite range. Renewed aging <strong>and</strong> slow<br />

cooling of a de-embrittled steel in the critical temperature region gives reembrittlement.<br />

These processes are accompanied by a redistribution of<br />

impurities on the grain boundaries.<br />

III. FACTORS DETERMINING MULTICOMPONENT<br />

INTERFACE ADSORPTION IN ENGINEERING STEELS,<br />

AND THE METHODS OF ITS CALCULATION<br />

The intercrystalline internal adsorption, or grain boundary segregation<br />

phenomenon, means the increased concentration of small impurities on<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 11 Diagram of time–temperature FATT, 8C: (1) 60; (2) 55; (3) 50; (4)<br />

45; (5) 40; (6) 35; (7) 30; (8) 25; (9) 20; (10) 15; (11) 10; (12) 5; (13) 0;<br />

(14) þ5; (15) þ10; (16) þ15; (17) þ20; (18) þ25; (19) þ30.<br />

the grain boundary (compared to bulk) is caused by decreasing boundary<br />

energy. The energy of impurity segregation corresponds to energy gain in<br />

the ‘‘bulk-boundary’’ system that accompanies the transition of one impurity<br />

atom from bulk to boundary. Thermodynamic description of this phenomenon<br />

is based on the Gibbs theory of the equilibrium segregation on<br />

free surface. The decrease of redundant energy of GB is the thermodynamic<br />

stimulus to change its chemical composition compared to bulk. The impurities<br />

that decrease the energy of interfaces are surface-active. These elements<br />

could <strong>for</strong>m equilibrium segregations under favorable conditions.<br />

The impurities that increase surface tension escape from surface. Gibbs’<br />

adsorption isotherm <strong>for</strong> grain boundary segregation in a solid binary system,<br />

where the matrix obeys Raoult’s law <strong>and</strong> Henry’s law of diluted solutions,<br />

may be expressed<br />

Gb ¼ Xc dgb kT dXc<br />

where Gb is the surplus concentration on the GB, mol=m 2 ; Xc is bulk mole<br />

part of impurity; dg b=dX c is the tendency to adsorb.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.<br />

ð5Þ


McLean [14] has developed a statistical model of intercrystalline internal<br />

adsorption. The main points of this theory are the following:<br />

1. A definite number of adsorption centers (one monolayer) are<br />

present on the GB. They have equal adsorption potential.<br />

2. The impurity atoms are adsorbed independently on each of the<br />

centers.<br />

3. The adsorption decreases at temperature growth.<br />

This model adequately describes the GBS process <strong>for</strong> two-component<br />

system:<br />

X 0 b<br />

Xb<br />

Xb<br />

¼ Xc<br />

exp<br />

1 Xc<br />

DG<br />

RT<br />

where Xb <strong>and</strong> Xc are the impurity concentrations on boundary <strong>and</strong> in bulk,<br />

respectively; X 0 b is the ultimate equilibrium concentration of the impurity<br />

on boundary; DG is the segregation energy.<br />

Hondros <strong>and</strong> Seah [15] have established the interrelation of GB enrichment<br />

<strong>and</strong> solubility limit X 0 c<br />

X 0 b<br />

Xb<br />

Xb<br />

¼ Xc<br />

X0 exp<br />

c<br />

DG<br />

RT<br />

The enrichment factor, determined as the GB content=bulk concentration<br />

ratio, is of the order of magnitude 10 4 <strong>and</strong> 10 1 <strong>for</strong> impurity <strong>and</strong> alloying elements,<br />

respectively (see Fig. 12).<br />

A. Binding Energy of Impurities with Grain Boundaries<br />

One can determine the adsorption energy as an alteration of the system free<br />

energy during transition of the dissolved atom from the grain bulk to boundary<br />

E ¼ðEb E 0 b Þ ðEc E 0 c Þ ð8Þ<br />

where Eb <strong>and</strong> Ec are the free energy of the system with impurity atoms on the<br />

grain boundary or in the bulk, respectively; Eb 0 <strong>and</strong> Ec 0 are the free energies<br />

of boundary <strong>and</strong> bulk in the pure solvent.<br />

The value of energy E is tied with the elastic (dimensional) <strong>and</strong> chemical<br />

interaction of impurity with boundary.<br />

The influence of dimensional discordance of atoms is accounted by the<br />

Eshelby equation [16]<br />

E ¼ 16<br />

3 pGr3 ðs 1Þ 2<br />

where G is the shear modulus of the solvent; r is the solvent atom radius; s is<br />

the ratio of impurity <strong>and</strong> solvent atoms radii.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.<br />

ð6Þ<br />

ð7Þ<br />

ð9Þ


Figure 12 Calculated <strong>and</strong> experimental enrichment coefficients <strong>for</strong> surface<br />

segregations. (From Ref. 15.)<br />

The decrease of elastic distortion energy during the transition of the<br />

impurity atom from ideal bulk lattice to distorted boundary lattice is the<br />

driving <strong>for</strong>ce adsorption.<br />

According to Ref. [17], one can describe the chemical part of adsorption<br />

Ech ¼ DZðe12 e11Þ ð10Þ<br />

where DZ is the difference of the atom coordination numbers in bulk <strong>and</strong><br />

on boundary; eij is the energy of interaction between the nearest neighbors<br />

(1 <strong>for</strong> the solvent, 2 <strong>for</strong> impurity).<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


The Fowler–Guggenheim theory [18] accounts <strong>for</strong> the interaction of<br />

atoms adsorbed on GB:<br />

X 0 b<br />

Xb<br />

Xb<br />

¼ Xc<br />

X0 exp<br />

c<br />

E1 ZoXb=X0 b<br />

KT<br />

ð11Þ<br />

where Z is the coordination number <strong>for</strong> impurity atoms; o is the interaction<br />

energy of nearest atoms.<br />

The portion of segregation energy ZoXb=X 0 b corresponds to interaction<br />

of the same <strong>and</strong> different atoms on boundary, <strong>and</strong> depends on the<br />

degree of neighbor centers filled. At o > 0, the atoms repulse mutually,<br />

<strong>and</strong> they attract at o < 0. The attraction increases the adsorption energy<br />

substantially. Interaction of atoms on interface influences segregation<br />

enrichment <strong>and</strong> promotes <strong>for</strong>mation of 2D phases with ordered atomic<br />

structure.<br />

B. Impurities’ Concurrence During Adsorption<br />

None of the existing adsorption theories adequately describe the micromechanism<br />

of impurities’ concurrence on the adsorption centers. This is related<br />

to the peculiarities of adsorption from gaseous phase to the free surface to<br />

describe the grain boundary segregation mechanism [19–21]. According to<br />

this point of view, all segregating elements (<strong>for</strong> example N, C, S, <strong>and</strong> P)<br />

occupy equal positions on GB, described by the ‘‘site competition’’ term.<br />

The peculiarity of GBS <strong>for</strong>mation consists of diffusion of alloying elements<br />

<strong>and</strong> impurities from bulk to interface. The migration mechanisms <strong>for</strong> substitial<br />

<strong>and</strong> interstitial impurities are different. The reason <strong>for</strong> this is that the<br />

adsorption centers on interface are different <strong>for</strong> these two kinds of impurities.<br />

There<strong>for</strong>e, the interstitial impurities (C, N) are located in interstices, but<br />

S, P, Sb, Bi, etc. occupy substitution position on the GB. Adsorption of any<br />

surface-active impurity on GB decreases its free energy <strong>and</strong> lowers the thermodynamic<br />

stimulus <strong>for</strong> adsorption of other impurity in a similar way. This<br />

causes a site competition between atoms on the GB [22].<br />

The concurrence of impurities A <strong>and</strong> B segregating at the same temperature<br />

<strong>and</strong> occupying the same positions in crystalline lattice (lattice<br />

points or interstices) with different binding energies to GB was investigated<br />

in Ref. [23].<br />

The equilibrium concentration of competing impurities A <strong>and</strong> B could<br />

be calculated using equations:<br />

X A b ¼<br />

XA C expðEAseg =kTÞ<br />

1 þ XA C expðEAseg =kTÞþXB C expðEBseg =kTÞ<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.<br />

ð12Þ


X B b ¼<br />

XB C expðEBseg =kTÞ<br />

1 þ XA C expðEAseg =kTÞþXB C expðEBseg =kTÞ<br />

ð13Þ<br />

where Eseg is the segregation energy; Xb I is the GB concentration of element I<br />

(atomic fraction). As one can see from these equations, at EA seg > EB seg , the<br />

concentration of element A decreases with increasing temperature.<br />

In this case, the adsorption level of the element B reaches its maximum<br />

at critical temperature:<br />

E A seg<br />

Tcr ¼<br />

k lnðEA seg =ðEBseg EA seg ÞXA b Þ<br />

ð14Þ<br />

The grain boundaries are enriched in this case by element B at low<br />

temperatures, <strong>and</strong> by element A at high temperatures.<br />

C. Thermodynamic Calculations of the<br />

Segregation Energy<br />

The segregation energy calculations are based on various models of solid<br />

solution electronic structure or quasi-liquid model of grain boundary.<br />

Thermodynamic properties of the solid solution determine firstly the<br />

surface activity of small impurities in the <strong>for</strong>mation of equilibrium GBS.<br />

The components of the alloy influence the energy of interaction with grain<br />

boundaries significantly. The parameters of such interaction are determined<br />

by thermodynamic calculations, phase equilibrium diagram analysis, computer<br />

modeling of GBS. The heat of solution of different atoms in solid<br />

solution is the thermodynamic measure of their interaction.<br />

The model establishing interrelationship segregation energy <strong>and</strong> heat<br />

of solutions is proposed in Ref. [24]<br />

Eseg ¼ FHsol Pðg A g BÞV 2=3<br />

A<br />

ð15Þ<br />

where F <strong>and</strong> P are empirical coefficients; H sol is the heat of solution of A in<br />

B [25–30]; g A, g B is the surface enthalpy of elements A <strong>and</strong> B [31,32]; V A is<br />

the molar volume of A.<br />

Using the liquid grain boundary model approximation, the segregation<br />

energy of impurities (Eseg) could be determined from an analysis of<br />

the solidus <strong>and</strong> liquidus curves on phase equilibrium diagrams:<br />

Eseg ¼ KT LnðK0Þ ¼ KT LnðCL CSÞ ð16Þ<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


where K0 is the coefficient of equilibrium distribution of the element<br />

between solid <strong>and</strong> liquid phases [33]; T is the melting temperature of the<br />

pure solvent; CL <strong>and</strong> CS are concentrations of impurity in the liquid <strong>and</strong><br />

solid solutions, respectively.<br />

An experimental method <strong>for</strong> determination of binding energy of<br />

impurity atoms to grain boundary is used.<br />

The analysis of large number of phase equilibrium diagrams has led<br />

to the establishment of the basic property of two-component solid solutions<br />

consisting of periodic variation of the segregation <strong>for</strong>mation energy<br />

of an element as a function of its location in the periodic table (atomic<br />

number). As seen in Fig. 13, the impurities could have positive or negative<br />

surface activity, or be neutral. The elements So, Mo, Ni, <strong>and</strong> Co are<br />

neutral in two-component alloys with Fe. At the same time, it is well<br />

known that molybdenum is the surface-active element in steels <strong>and</strong> it<br />

reduces the tendency of steel to the reversible temper embrittlement. This<br />

change of surface activity is observed only in multicomponent alloys, <strong>and</strong><br />

it is due to the mutual influence of elements on its thermodynamic activity.<br />

The binding energy of an impurity to the GB depends significantly on<br />

boundary structure. The wide spectrum of Eseg exists <strong>for</strong> the given substance<br />

analogous to the spectrum of the GB energy. This circumstance<br />

explains the wide dispersion of the segregation energy <strong>for</strong> various impurities<br />

that are listed in literature sources. Based on this reasoning, it is<br />

useful, <strong>for</strong> segregation modeling, to apply the unified approach <strong>for</strong> determination<br />

of the generalized characteristic of definite impurity segregation<br />

in definite solvent. The thermodynamic calculations of segregation energy<br />

are the most suitable way <strong>for</strong> its estimation. Auger electron spectroscopy<br />

(AES) <strong>for</strong> investigation of segregation kinetics on the free surface of polycrystalline<br />

foils is a reliable experimental technique <strong>for</strong> the averaged E seg<br />

determination.<br />

The part of elastic <strong>and</strong> chemical interaction in GBS process could be<br />

estimated experimentally based on concentration dependencies of segregation<br />

energy. These dependencies were determined <strong>for</strong> alloys whose compositions<br />

are listed in Table 1.<br />

The segregation energy was determined based on AES of equilibrium<br />

free surface segregations of phosphorus. The polycrystalline foil samples<br />

were tempered at 823K <strong>for</strong> 4 hr in a work chamber of electron spectrometer<br />

ESCALAB MK2 after quenching from austenitization temperature of<br />

1323K. The segregation energy of P was determined using Eq. (6) based<br />

on surface <strong>and</strong> bulk impurity concentration. Figure 14 shows the dependence<br />

of the P segregation energy <strong>and</strong> its bulk content in alloy. For the<br />

diluted solid solutions, Eseg is independent of concentration or temperature.<br />

It is caused only by elastic distortions that are <strong>for</strong>med by impurity atoms in<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 13 Change of calculated E seg of impurities in Fe-base alloys in accordance<br />

to its number in periodic system. Calculations were based on Hsol in the following<br />

publications: (a) Refs. 25 <strong>and</strong> 26; (b) Refs. 27 <strong>and</strong> 28; (c) Ref. 33.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Table 1 Composition (at.%) of the Fe–P <strong>and</strong> Fe–P–Mo Alloys<br />

Chemical composition, at.%<br />

C S P Mo<br />

0.01 0.002 0.017 0<br />

0.01 0.003 0.10 0<br />

0.01 0.002 0.15 0<br />

0.01 0.003 0.093 3.1<br />

0.01 0.002 0.033 3.1<br />

0.01 0.002 0.14 3.1<br />

0.01 0.005 0.09 0.3<br />

0.01 0.014 0.074 0.02<br />

bulk <strong>and</strong> on the interface. As seen in Fig. 14, the elastic interaction energy of<br />

the P atoms with grain boundaries in iron is equal to 0.53 eV=at <strong>and</strong><br />

decreases significantly at molybdenum alloying to 0.24 eV=at in the alloy<br />

Fe–3.1at.% Mo. Decrease of segregation energy of the impurity at its<br />

Figure 14 Change of E seg of phosphorus with its volume concentration in Fe (1)<br />

<strong>and</strong> Fe–3.1 at.% Mo–P alloys (2). Auger electron spectroscopy of free surface<br />

segregations at 823K.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


volume concentration growth is caused by chemical pair interaction of the<br />

atoms in alloy.<br />

Using the example of the Fe–P system, we could determine chemical<br />

interaction of elements by applying the approach proposed in Ref. [34].<br />

Analyzing the solidus <strong>and</strong> liquidus equilibrium (volume <strong>and</strong> GB) on the<br />

equilibrium phase diagram at three temperatures permits the construction<br />

of a system of three equations that describe this equilibrium<br />

100 Xs<br />

kT qa ln<br />

100 Xl<br />

¼ X 2 s W0 X 2 l W00 þ kqaTa ð17Þ<br />

where k is the Boltzmann constant; T a is the melting temperature of<br />

Fe; qa is melting entropy per atom divided by Boltzmann constant;<br />

W 0 <strong>and</strong> W 00 are the mixing energies in solid <strong>and</strong> liquid states; Xs <strong>and</strong><br />

Xl are the impurity concentration in solid <strong>and</strong> liquid phases at the temperature<br />

T.<br />

Solving these equations <strong>for</strong> the phase diagram of Fe–P binary system<br />

[35], the sign <strong>and</strong> value of mixing energy in liquid phase equal<br />

0.425 eV=at were determined. The positive value (in accordance with physical<br />

sense) means that binding <strong>for</strong>ce of P–P <strong>and</strong> Fe–Fe atoms is higher<br />

than <strong>for</strong> Fe–P atoms:<br />

W ¼ WFe P<br />

1<br />

2 ðWFe Fe þ WP PÞ ð18Þ<br />

emphasizing the tendency <strong>for</strong> solid solution tendency <strong>for</strong> stratification or<br />

intercrystalline internal adsorption.<br />

D. Effect of Solute Interaction in Multicomponent System<br />

on the Grain Boundary Segregation<br />

Guttman has exp<strong>and</strong>ed the concept <strong>for</strong> synergistic co-segregation of alloying<br />

elements <strong>and</strong> harmful impurities at the grain boundaries. His theory is<br />

very important <strong>for</strong> analysis of steels <strong>and</strong> alloys that contain many impurities<br />

<strong>and</strong> alloying elements. In accordance with the theory, the interaction<br />

between alloying elements <strong>and</strong> the impurity atoms could be estimated from<br />

enthalpy of <strong>for</strong>mation of the intermetallic compounds (NiSb, Mn2Sb, Cr3P,<br />

etc.). The alloying elements could influence on the solubility of impurities in<br />

the solid solution. Only the dissolved fraction of the impurity takes part in<br />

the segregation [36]. When preferential chemical interaction exists between<br />

M (metal) <strong>and</strong> I (impurity) atoms with respect to solvent, the energy of<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


segregation becomes functions of the intergranular concentrations of I<br />

<strong>and</strong> M:<br />

DGI ¼ DG 0 I<br />

DGM ¼ DG 0 M<br />

þ bb<br />

MI<br />

C b Yb M<br />

þ bb<br />

MI<br />

a b Yb I<br />

b a<br />

MI<br />

C a Xa M<br />

ð19Þ<br />

b a<br />

MI<br />

a a Xa I ð20Þ<br />

where C b <strong>and</strong> a b are the fractions of sites available in the interface <strong>for</strong> I <strong>and</strong><br />

M atoms, respectively ðab þ Cb ¼ 1Þ; Y b is the partial coverage in the interface;<br />

X a is the concentration in the solid solution a; bMI is the interaction<br />

coefficient of M <strong>and</strong> I atoms in a-solid solution (a) or on the grain boundary<br />

(b). For a preferentially attractive M–I interaction, the bMIare positive <strong>and</strong><br />

the segregation of each element enhances that of the other. If the interaction<br />

is repulsive, the bMI are negative <strong>and</strong> the segregations of both elements will<br />

be reduced. For a high attractive M–I interaction in the a-solid solution, the<br />

impurity can be partially precipitated in the matrix into a carbide, or intermetallic<br />

compound. The interface is then in equilibrium with an a-solution<br />

where the amount of dissolved I, XI a , may become considerably smaller than<br />

its nominal content.<br />

In the ternary solid solutions, the segregation of impurity (I) could be<br />

lowered or neglected at several critical concentrations of the alloying<br />

element (M) whose value (C M a) depends on surface activity of each compo-<br />

I,M<br />

nent (ESeg ) <strong>and</strong> interaction features of the dissolved atoms (bMI):<br />

C M a ¼<br />

E I Seg<br />

bMIðexpðEM Seg =RTÞ 1Þ<br />

ð21Þ<br />

The critical concentration of alloying element is accessible <strong>for</strong> segregation of<br />

impurity <strong>and</strong> alloying element E I;M<br />

Seg > 0 <strong>and</strong> repulsion of different atoms<br />

bMI > 0; or without segregation of alloying element EM Seg < 0 <strong>and</strong> with<br />

attraction of different atoms bMI < 0.<br />

I,M<br />

In this case, the dependence of ESeg on the dissolved element concentration<br />

is not taken into account. Indeed, <strong>for</strong> systems with limited solubility, the<br />

alteration of value <strong>and</strong> sign of segregation energy is possible at a definite content<br />

of alloying element. The phase equilibrium diagram analysis allows the<br />

determination of mutual influence of components on their surface activity.<br />

The equilibrium distribution of solute elements between solid <strong>and</strong><br />

liquid phases in iron-base ternary system (distribution interaction coefficient<br />

K0) is known to be an important factor in relation to microsegregation during<br />

the solidification of steels. As it was shown above, these analogies<br />

are useful <strong>for</strong> the prediction of GBS <strong>and</strong> <strong>for</strong> impurity segregation energy<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 15 Change of the equilibrium distribution coefficient of some elements with<br />

carbon concentration in Fe–C-based ternary systems. (From Ref. 37.)<br />

determination in the given solvent. The K0 of some elements, especially in<br />

multicomponent systems, is considered to be different from those in binary<br />

systems because of the possible existence of solute interactions, but the<br />

mechanisms are so complicated that detailed in<strong>for</strong>mation has not yet been<br />

obtained. There<strong>for</strong>e, it would be very useful if the effect of an addition of<br />

an alloying element on the distribution could be determined by the use of<br />

a simple parameter.<br />

Equilibrium distribution coefficient K0 1 of various elements in Fe–C<br />

base ternary system is calculated from equilibrium distribution coefficient in<br />

iron-base binary systems [40–43]. In Fig. 15, the calculated results are compared<br />

with the measured values by various investigators. The changes of the<br />

K 0 1 of P <strong>and</strong> S with various alloying elements are shown in Fig. 16(a, b) in<br />

Fe–P <strong>and</strong> Fe–S base ternary system, respectively.<br />

These data could be applied <strong>for</strong> calculation of phosphorus segregation<br />

energy change under the alloying element influence in Fe–Me–0.1at.% P<br />

alloys (Fig. 17) or <strong>for</strong> calculation of the segregation energy change of<br />

alloying elements with concentration of carbon in Fe–0.1Me–C alloys<br />

(Fig. 18). For the growth of carbon volume content, the segregation energy<br />

of C <strong>and</strong> P decreases which means lowering of the segregation stimulus <strong>for</strong><br />

these elements.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 16 (a) Change of the equilibrium distribution coefficient of phosphorus<br />

with the concentration of alloying elements; <strong>and</strong> (b) change of the equilibrium<br />

distribution coefficient of sulfur with the concentration of alloying elements. Solid<br />

line: a-phase. Chain line: g-phase. (From Ref. 38.)<br />

E. Kinetics of Segregation<br />

The existing models of multicomponent adsorption do not analyze in detail<br />

the kinetics of the process. But in reality, GBS <strong>for</strong>ms only during a limited<br />

time of the heat treatment process. The difference of segregation level from<br />

the equilibrium one depends on temperature <strong>and</strong> time. At low temperatures<br />

<strong>and</strong> limited time of heat treatment, segregation is controlled by diffusion. As<br />

the temperature increases, segregations with lower equilibrium concentra-<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 16 (Continued)<br />

tion are developed, but rich segregations dissolve. Distinguishing diffusion<br />

mobility <strong>and</strong> mutual influence of elements on their diffusion coefficients<br />

determines much of their segregation ability. Amplification or suppression<br />

of adsorption could be due to a kinetic factor. This peculiarity determines<br />

the fundamental factor of distinguishing adsorption from gas phase to free<br />

surface when comparing it to intercrystalline internal adsorption: GBS is<br />

controlled by diffusion during heat treatment of steels <strong>and</strong> alloys.<br />

Many GBS features in multicomponent systems cannot be predicted<br />

adequately using the equilibrium segregation thermodynamic accounting<br />

basis. Particularly, the thermodynamic concept of the cooperative (synergis-<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 17 Change of E seg of phosphorus with the concentration of alloying<br />

elements in Fe–Me–0.1% P alloys.<br />

tic) adsorption of elements is disturbed when they do not segregate at the<br />

same temperature. The concurrence of impurities at GBS could be tied not<br />

only due to their attractive or repulsive interaction, but also with higher diffusion<br />

mobility of some impurities. In many cases, the determination interatomic<br />

interaction on grain boundary that is proposed in Guttmann’s<br />

theory has a significantly lower effect <strong>for</strong> segregation prediction than accounting<br />

of mutual influence of elements on their thermodynamic activity in the<br />

grain bulk.<br />

Mutual influence of the alloy components on their surface activity is<br />

caused by their interaction in solid solution in the bulk. The interaction<br />

on grain boundaries could be analyzed only <strong>for</strong> those elements that<br />

segregate in near temperature ranges. Many postulates of the thermodynamic<br />

theory of equilibrium grain boundary segregation could not be<br />

applied simply <strong>for</strong> heat treatment of multicomponent alloys. This is especially<br />

important <strong>for</strong> steels, which have complex phase trans<strong>for</strong>mations during<br />

treatment that accompany change of the solid solution composition.<br />

Auger electron spectroscopy permits the investigation of multicomponent<br />

adsorption kinetics. The composition of grain boundaries on the intercrystalline<br />

fracture surface made under high vacuum is analyzed <strong>for</strong> this purpose.<br />

In these cases, the experimental modeling of GBS is widely used.<br />

The chemical composition of free surface of thin poly-crystalline foils that<br />

are heated in situ is investigated using Auger electron spectrometers. The<br />

P grain boundary adsorption isotherms <strong>for</strong> samples of three Fe–Cr–Mn<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 18 Change of Eseg of alloying elements (Me) with the concentration of<br />

carbon in Fe–0.1% Me–C alloys.<br />

steels after quenching from 1273K <strong>and</strong> tempering at 923K <strong>for</strong> 25 min, 1 <strong>and</strong><br />

2 hr with air cooling are presented in Fig. 19. Dissolution of Ti <strong>and</strong> V carbonitrides<br />

after steel quenching promotes enrichment of the solid solution<br />

by these elements. They have high values of Gibbs energy <strong>for</strong> phosphide <strong>for</strong>mation,<br />

decrease the thermodynamic activity of phosphorus in solid solution<br />

<strong>and</strong> reduce its GBS.<br />

Most models of kinetics are classically analyzed in terms of the law<br />

derived by McLean [14] <strong>for</strong> binary alloys<br />

XbðtÞ Xbð0Þ<br />

Xb Xbð0Þ<br />

4Dit<br />

¼ 1 exp<br />

ðXb=Xa i Þ2d2 " # pffiffiffiffiffiffiffi<br />

2 Dit<br />

erfc<br />

ðXb=Xa i Þ<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.<br />

ð22Þ


Figure 19 Kinetics of P GBS in steel 0.3C–1.6Mn–0.8Cr–008P (1) with adds of<br />

0.047Ti (2) or (0.07Ti <strong>and</strong> 0.026V) (3), quenched from 1273K <strong>and</strong> tempered at 923K.<br />

where Xb(t) is the interfacial coverage of element, at time t; Xb(0) — is its<br />

initial value <strong>and</strong> Xb its equilibrium value as defined by Eq. (7); Xi a —is<br />

its volume concentration; Di is the bulk diffusivity of i <strong>and</strong> d is the interface<br />

thickness.<br />

a<br />

Assuming Xb=Xi ¼ const, using Laplace trans<strong>for</strong>mation <strong>for</strong> (22), one<br />

can obtain the approximate expression<br />

XbðtÞ Xbð0Þ<br />

Xb Xbð0Þ ¼ 2Xa i<br />

Xbd<br />

rffiffiffiffiffiffiffiffiffi<br />

FDti<br />

p<br />

ð23Þ<br />

where F ¼ 4 <strong>for</strong> grain boundaries <strong>and</strong> F ¼ 1 <strong>for</strong> free surface.<br />

The kinetics of segregation dissolution could be described by these<br />

equations (22) <strong>and</strong> (23). But, in this case, the variables Xb(0) <strong>and</strong> Xb<br />

exchange places. The influence of Mo, Cr, <strong>and</strong> Ni additions on kinetics of<br />

P segregation has been studied in six Fe–Me–P alloys, whose base compositions<br />

are listed in Table 1. These materials were austenitized <strong>for</strong> 1 hr at<br />

1323K <strong>and</strong> quenched in water. The tempering of foils at 773K was carried<br />

out in a work chamber of an electron spectrometer ESCALAB MK2 (VG).<br />

The kinetics of P segregation studied <strong>for</strong> Fe–Me–P alloys (Figs. 20–22) show<br />

that equilibrium is reached within several hours. Based on the starting position<br />

of adsorption isotherms, the phosphorus diffusion coefficients in these<br />

alloys were calculated using Eq. (22). The data are presented in Table 2.<br />

Molybdenum reduces significantly P surface activity <strong>and</strong> decelerates its<br />

diffusion. Nickel is not a surface-active element in carbonless alloys, Fe–<br />

P–Ni. It increases sharply P thermodynamic activity <strong>and</strong> equilibrium GB<br />

concentration, <strong>and</strong> accelerates its diffusion. Chromium segregates poorly<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 20 Kinetics of P segregation on free surface in Fe–P–Mo alloys with<br />

different relative concentration of Mo=P at 773K.<br />

in these alloys. It also, as Ni, increases diffusion mobility of P <strong>and</strong> its grain<br />

boundary adsorption.<br />

The adsorption isotherms of various elements have non-monotonous<br />

shape in multicomponent alloys. The isodose thermokinetic diagrams<br />

present the averaged in<strong>for</strong>mation on segregation of all components. Such<br />

Figure 21 Kinetics of P <strong>and</strong> Cr segregation on free surface in Fe–0.04P–2.3 Cr<br />

alloy at 773K.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 22 Kinetics of P segregation on free surface in Fe–P–Ni alloys with<br />

different relative concentration of Ni=P at 773K.<br />

diagrams <strong>for</strong> steel with 0.2–0.3% C alloyed by Cr, Mo, Ni, Mn, V, <strong>and</strong> Nb<br />

are presented in Figs. 23–27.<br />

Chemical composition of steel is listed in Table 3. The T–t diagrams<br />

are obtained based on adsorption isotherms on free surface of foils that were<br />

heat treated in Auger spectrometer ESCALAB MK2 at vacuum about<br />

10 10 Torr. The isodose curves characterizing time of definite segregation<br />

level access depending on temperature are shown in these diagrams. At elevation<br />

of an isothermal exposition temperature, the mobility of impurities<br />

increases, <strong>and</strong> time <strong>for</strong> reaching of definite segregation level decreases.<br />

The lower branch of isodose curve means decrease of segregation <strong>for</strong>mation<br />

Table 2 Composition (at.%) of the Fe–Me–P Alloys <strong>and</strong> Kinetics Characteristics<br />

of P Free Surface Segregation, Deduced from the Segregation Kinetics<br />

Chemical Composition, at.%<br />

P Mo Ni Cr<br />

Surface<br />

activity X b=X i a<br />

Bulk diffusivity of<br />

P DP 10 18 (m 2 =sec)<br />

0.07 0 0.9 0 285 5.34<br />

0.03 0 3.1 0 2530 1048<br />

0.04 0 0 2.3 1450 160<br />

0.16 1.0 0 0 380 27<br />

0.21 2.1 0 0 250 2.28<br />

0.15 3.1 0 0 80 0.3<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 23 The isodose C-curves of multicomponent interface segregation in 0.3C–<br />

Cr–Mo steel (see Table 3). Auger electron spectroscopy of free surface segregations.<br />

time at increasing temperature. With temperature increase, the solubility of<br />

impurity in solid solution increases, <strong>and</strong> its GB concentration reduces. It follows<br />

that the probability to <strong>for</strong>m the segregation with high impurity content<br />

reduces, <strong>and</strong> time <strong>for</strong> such segregation increases extensively. The upper<br />

branch of isodose curves corresponds to dissolution of rich segregations<br />

<strong>and</strong> access to new equilibrium with lower impurity concentration. The<br />

Figure 24 The isodose C-curves of multicomponent interface segregation in 0.2C–<br />

Cr–Mn–Ni–Si steel (see Table 3) under its tempering. Auger electron spectroscopy of<br />

free surface segregations.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 25 The isodose C-curves of multicomponent interface segregation in 0.3C–<br />

Cr–Mn–Nb steel (see Table 3) under its tempering. Auger electron spectroscopy of<br />

free surface segregations.<br />

adsorption patterns <strong>for</strong> engineering steels have common as well as individual<br />

features. As a rule, carbon segregates at temperatures lower than<br />

523K, nitrogen—in 523–623K range, phosphorus—in 523–823K range, sulfur<br />

segregates at temperatures higher than 723K.<br />

The substitual <strong>and</strong> interstitial element concurrence promotes blocking<br />

of adsorption centers by mobile impurities <strong>and</strong> impedes P segregation at<br />

Figure 26 The isodose C-curves of multicomponent interface segregation in 0.3C–<br />

Cr–Mn–V steel (see Table 3) under its tempering. Auger electron spectroscopy of free<br />

surface segregations.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 27 The isodose C-curves of multicomponent interface segregation in 0.3C–<br />

Cr–Mn–Si–Ti steel (see Table 3) under its tempering. Auger electron spectroscopy of<br />

free surface segregations.<br />

temperatures lower than 523–673K. The preferential enrichment of GB by P<br />

<strong>and</strong> S becomes possible after dissolution of C <strong>and</strong> N segregations. The alloying<br />

elements change significantly the segregation stability regions <strong>for</strong> various<br />

elements. Fig. 28(a,b) shows the P adsorption isotherms in the<br />

investigated steels at 723K. Molybdenum sharply slows down the P<br />

segregation <strong>for</strong>mation. The differences in diffusion mobility of elements<br />

<strong>and</strong> temperature intervals of segregation stability are the reasons <strong>for</strong> nonequilibrium<br />

enrichment of grain boundaries. The rich segregations are<br />

<strong>for</strong>med at the initial stage of isothermal exposition, <strong>and</strong> they are dissolved<br />

after longer exposition. Comparing the behavior of 0.3C–Cr–Mn–Nb (1)<br />

<strong>and</strong> 0.22C–Cr–Mn–Si–Ni (3) steels at 673K tempering, one can see that<br />

small (lower than 20 min) expositions 0.3C–Cr–Mn–Nb, <strong>and</strong> longer ones<br />

(about 1 hr 20 min) are dangerous <strong>for</strong> 0.22C–Cr–Mn–Si–Ni steel. Analyzing<br />

Table 3 Chemical Composition of Steels<br />

No. Steel<br />

Concentration of elements, wt.%<br />

C Si Mn Cr V Al Ti Nb Ni Mo S P<br />

1 0.3C–Cr–Mn–V 0.32 0.25 0.88 0.92 0.088 0.014 0.024 0 0 0 0.016 0.027<br />

2 0.3C–Cr–Mn–Nb 0.29 0.33 1.04 1.07 0 0.007 0.036 0.025 0 0 0.014 0.027<br />

3 0.3C–Cr–Mo 0.33 0.23 0.56 0.96 0.003 0.014 0.025 0 0 0.25 0.005 0.004<br />

4 0.3C–Cr–Mn–Si–Ti 0.28 0.61 1.15 0.75 0 0.029 0.016 0 0.32 0 0.013 0.022<br />

5 0.2C–Cr–Mn–Ni–Si 0.22 0.43 0.92 0.89 0 0.030 0.015 0 0.91 0 0.015 0.025<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 28 Influence of alloying on the kinetics isotherms of P free surface<br />

segregation at 723K. The following steels were investigated (see Table 3): 1, 3C–Cr–<br />

Mn–Nb; 2, 3C–Cr–Mn–Si–Ti; 3, 2C–Cr–Mn–Ni–Si; 4, 3C–Cr–Mo; 5, 3C–Cr–<br />

Mn–V.<br />

the thermokinetic diagrams <strong>for</strong> ternary Fe–Me–P alloys based on Eqs. (23)<br />

<strong>and</strong> (6), the mutual influence of elements on their binding energy to GB was<br />

determined [36]<br />

E P seg ¼ 20:6 þ 183CP a<br />

4:8C Al<br />

a<br />

7:2C Mo<br />

a<br />

3:4C Ni<br />

a<br />

7141C B a<br />

þ 4:9C Cr<br />

a 444C S a 183E Mo<br />

seg 87E N seg ð24Þ<br />

E S seg ¼ 6:9 151CS a 1:5C Al<br />

a þ 14:5CP a 39E Sn<br />

seg<br />

E N seg<br />

¼ 16 2:6CAl<br />

a<br />

þ 3CMo<br />

a<br />

þ 4:2CCr a<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.<br />

2625C Ti<br />

a<br />

þ 175EMo<br />

seg<br />

ð25Þ<br />

ð26Þ


E C seg ¼ 7:9 1:4CAl a þ 5CMo a þ 676CB a þ 1:2CCr a<br />

130E N seg þ 116EP seg<br />

ð27Þ<br />

E Mo<br />

seg ¼ 0:7 þ 32EN seg 28E P seg ð28Þ<br />

E Ti<br />

seg ¼ 17 þ 3CC a E P seg ð29Þ<br />

E Al<br />

seg ¼ 1:4CAl a<br />

E Sn<br />

seg ¼ 21; ENi seg ¼ 14; EBseg ¼ 54; ECu seg ¼ 20 kJ/mol<br />

ð30Þ<br />

I<br />

where Esegis<br />

segregation energy of the I element, Ca j is bulk concentration of<br />

j impurity.<br />

F. Stability of the Segregation<br />

The equilibrium GBS dissolves as temperature increases. Analysis of the<br />

kinetic development of the equilibrium segregation level of P shown in<br />

Fig. 29 gives the T–t plot of segregation directly. Obviously that segregation<br />

level close to the maximum exists only within a specific temperature range.<br />

This range is characterized by a maximum temperature stability T max, over<br />

which the intensive dissolution of the segregates is observed. This temperature<br />

can be calculated by computer analysis of Eq. (7) at dCb max =dT ¼ 0.<br />

The temperature Tmax depends on Eseg <strong>and</strong> temperature dependencies of<br />

solubility limits, which can be determined from analysis of phase equilibrium<br />

diagrams [43].<br />

Using these dependencies as a generalizing criterion, it is possible to<br />

simplify the analysis of data on element segregation kinetics in iron alloys.<br />

The interrelationship of maximum temperature of stability (T max) of rich<br />

equilibrium segregations <strong>and</strong> segregation energies of different elements is<br />

presented in Fig. 30.<br />

The common features of kinetics show the following groups:<br />

1. enriching grain boundaries at low- <strong>and</strong> medium-tempering<br />

temperatures—B, C, N, <strong>and</strong> Cu;<br />

2. co-segregating with P at high tempering—P, Sn, Ti, <strong>and</strong> Mo;<br />

3. segregating at high temperatures—S <strong>and</strong> Al.<br />

Phosphorus in Fe alloys has abnormally weak dependence of Tmax<br />

on Eseg in reversible temper embrittlement temperature range. In other<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 29 The calculated segregation level of P as a function of temperature<br />

according to Eq. (7). Tmax is the maximal temperature of stability of rich segregation<br />

level. (From Ref. 42.)<br />

words, this means that the temperature of P segregation stability in the<br />

RTE development interval weakly depends on segregation energy or alloy<br />

composition. This circumstance is associated with the specific shape of the<br />

temperature dependence of P solubility in Fe. The established regularity<br />

allows to explain the difficulties with rational alloying of engineering steels<br />

<strong>for</strong> RTE suppression.<br />

G. Nature of Reversibility of Temper Embrittlement<br />

The reversibility of temper embrittlement is usually associated with precipitation<br />

or dissolution of carbide phase at various modes of quenched steel<br />

heat treatment below Ac1 [44–46]. The complex character of multicomponent<br />

GB adsorption—namely interrelation of two opposite processes: concurrence<br />

between impurities, <strong>and</strong> their cooperative segregation—is not<br />

taken into account using this approach.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 30 The interconnection of Tmax-segregation stability temperature <strong>and</strong><br />

Eseg-energy of impurities segregation in Fe-base alloys.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 31 Thermo-kinetics diagrams of multicomponent segregation on free<br />

surface in steel 0.35C–1.58Mn–0.1P–0.6Al.<br />

Figure 31 presents the thermokinetic diagram of element segregation<br />

in 0.35C–1.5Mn–0.1P–0.6Al steel. The chemical composition of free surface<br />

segregations was determined by AES <strong>for</strong> a set of isothermal conditions in<br />

the spectrometer ESCALAB MK2 (VG). The temperature–time interval<br />

of preferential segregation of chemical elements is the result of different diffusion<br />

mobility <strong>and</strong> binding energy of elements with GB. The temperature<br />

interval of P preferential segregation is caused by concurrence of this impurity<br />

with mobile interstitial elements C <strong>and</strong> N. This process determines<br />

temperature <strong>and</strong> exposition necessary <strong>for</strong> RTE development. Direct investigation<br />

of grain boundary composition by AES confirms the conclusion<br />

about the prevailing role of concurrent segregation in RTE. The composition<br />

of several grain boundaries on brittle intercrystalline fracture of<br />

0.35C–Mn–Al steel after heat treatment: quenching from 1223K, tempering<br />

at 923K <strong>for</strong> 1 hr with rapid (a) <strong>and</strong> slow (b) cooling is presented in Fig. 32<br />

[47]. These data are in good correspondence with those in Fig. 31. Accelerated<br />

cooling of steel, does not provide enough time <strong>for</strong> the development of<br />

segregations with high P content, <strong>and</strong> GB are enriched by C. During slow<br />

cooling, phosphorus has enough time to enrich the grain boundaries. In this<br />

case, the carbon concentration on GB is sufficiently lower than at rapid<br />

cooling of steel. Carbon segregations are unstable at temperatures higher<br />

than 500–673K, <strong>and</strong> they are dissolved. At slow cooling, P segregates to<br />

grain boundaries, decreasing the GB redundant energy. This circumstance<br />

lessens the thermodynamic stimulus <strong>for</strong> carbon segregation as the temperature<br />

decreases. Carbon <strong>and</strong> phosphorus in steels are responsible <strong>for</strong> RTE<br />

development. They have high surface activity <strong>and</strong> diffusion mobility that<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 32 Chemical composition of GB in steel 0.35C–1.58Mn–0.1P–0.6Al (AES);<br />

(a) tempering at 923K, water cooling; (b) tempering at 923K, cooling with furnace.<br />

(From Ref. 47.)<br />

predetermines their segregation on GB at heat treatment. Difference of diffusion<br />

mobility as well as difference of maximum temperature of segregation<br />

stability is the reason <strong>for</strong> preferential segregation of an impurity. This is the<br />

reason <strong>for</strong> the characteristic temperature region of RTE development.<br />

Reversibility (i.e. disappearance) of temper embrittlement is associated<br />

with full dissolution of rich GBS of phosphorus at high temperatures <strong>and</strong><br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


enrichment of GB by carbon at rapid cooling [48]. Undoubtedly, carbide<br />

trans<strong>for</strong>mation, internal stresses, substructure trans<strong>for</strong>mations are very<br />

important <strong>for</strong> RTE. One should take into account such circumstances<br />

where kinetics of C <strong>and</strong> P segregation are dependent significantly on steel<br />

alloying.<br />

IV. DYNAMIC SIMULATION OF GRAIN BOUNDARY<br />

SEGREGATION<br />

A. Interface Adsorption During Tempering of Steel<br />

1. Decomposition of Martensite<br />

The common laws of multicomponent GBS <strong>and</strong> analysis of experimental<br />

diagrams on elements segregation kinetics in iron alloys are used to develop<br />

the computer models of these processes. The exact solution of McLean’s<br />

diffusion Eq. (21) accounting <strong>for</strong> temperature dependant of diffusion <strong>and</strong><br />

element solubility is a complex problem. In low-alloyed steels, the concentration<br />

of surface-active impurities (S, P, <strong>and</strong> N) is rather small, <strong>and</strong> based<br />

on this reason, it is possible to analyze the diffusion of each element separately.<br />

The model takes into account mutual influence of bulk <strong>and</strong> surface<br />

concentration of elements with respect to segregation energies.<br />

Carbon in solid solution has maximum influence on phosphorus GBS<br />

kinetics. Concentration of C in martensite changes significantly during<br />

quenched steel tempering <strong>and</strong> mainly depends on alloying element content.<br />

Based on this reason, one should take into account the solid solution composition<br />

altering segregation processes modeling during tempering.<br />

Investigations of martensite tetragonality at alloyed steel tempering<br />

[6,7] are the basis <strong>for</strong> calculations of mutual influence of alloying elements<br />

on martensite decomposition kinetics <strong>and</strong> carbon content in solid solution.<br />

The carbon content change in solid solution during tempering of engineering<br />

steels is well described by equation<br />

where<br />

DX C a<br />

X C a ð0Þ ¼ 1 exp KDot exp<br />

Q<br />

RT<br />

n<br />

ð31Þ<br />

DX C a ¼ XCa ð0Þ XCa ðtÞ ð32Þ<br />

X a C (0) <strong>and</strong> Xa C (t) are the carbon content in quenched steel <strong>and</strong> after a time t;<br />

Do is the carbon diffusion coefficient; Q is the activation energy associated<br />

with the interstitial diffusion of carbon atoms; K is the constant associated<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Table 4 Coefficient A in Eq. (28) <strong>for</strong> Low-Alloying Engineering Steels<br />

Coefficient A<br />

with the nucleation; n is the constant independent of both temperature <strong>and</strong><br />

X a C (0); R is the gas constant <strong>and</strong> T is the temperature.<br />

Influence of C <strong>and</strong> alloying elements on parameters Q, K, <strong>and</strong> n in<br />

Eq. (31) is determined <strong>for</strong> various steels. The activation energy Q in lowalloyed<br />

steel depends on the concentration of carbon <strong>and</strong> alloying elements<br />

in solid solution:<br />

Qðcal=molÞ ¼8571:5X C a<br />

þ A XMe<br />

a<br />

Alloying element<br />

Ni Si Mn Cr Mo<br />

433.56 1,432.54 726.35 2,898.91 971.51<br />

þ 18; 000 ð33Þ<br />

where Xa C <strong>and</strong> Xa Me are concentrations of C <strong>and</strong> alloying elements, mass%;<br />

A is a constant depending on alloying element. The values of coefficients in<br />

Eq. (31) are presented in Tables 4 <strong>and</strong> 5. The diffusion activation energy of<br />

Table 5 Influence of Carbon <strong>and</strong> Alloying Elements on Parameters Q, K,<br />

<strong>and</strong> n in Eq. (31)<br />

Steel, wt.% Q, cal=mol Ln K n<br />

0.4C–0.24Ni 21,532 15.364 0.26<br />

0.39C–3.0Ni 22,643 17.481 0.22<br />

0.37C–5.6Ni 23,599 18,575 0.24<br />

0.4C–0.32Mn 21,196 15.737 0.24<br />

0.4C–1.32Mn 20,298 14.241 0.22<br />

0.4C–2.43Mn 19,406 13.713 0.24<br />

0.4C–0.2Cr 20,848 15.366 0.21<br />

0.4C–2.1Cr 15,348 10.076 0.24<br />

0.4C–3.6Cr 10,992 5.481 0.42<br />

0.4C–6.7Cr 2,005 1.698 2.32<br />

0.4C–0.37Si 21,929 16.351 0.19<br />

0.38C–1.75Si 23,764 15.234 0.22<br />

0.4C–2.75Si 25,368 10.050 0.15<br />

0.4C 21,429 16.72 0.24<br />

1.4C 30,000 40.881 0.07<br />

1.2C–2.0Mo 26,343 29.768 0.08<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 33 Change of carbon concentration in solid solution with temperature <strong>and</strong><br />

time of tempering. Steel 0.43C–2.43 Mn (mass%). Isodose curves <strong>for</strong>: 1, 1 at.% C; 2,<br />

0.5 at.% C; 3, 0.1 at.% C; 4, 0.05 at.% C; 5, 0.03 at.% C.<br />

carbon decreases on the growth of carbide-<strong>for</strong>ming element (Mn, Cr, <strong>and</strong><br />

Mo) concentration. The contrary effect is observed <strong>for</strong> Ni <strong>and</strong> Si.<br />

Obviously, it is associated with the different influence of these elements<br />

on thermodynamic activity of carbon in ferrite. These dependencies are<br />

basic <strong>for</strong> calculations of segregation kinetics of C since carbon is the element<br />

that influences on P segregation highly. The kinetics of carbon content in<br />

solid solution change during tempering of quenched steel 0.43C–2.43Mn<br />

(mass%) are shown in Fig. 33. These data are obtained by computer modeling<br />

using Eqs. (31–33) <strong>and</strong> those from Tables 4 <strong>and</strong> 5.<br />

This model provides the possibility of calculating the influence of<br />

alloying on cementite <strong>for</strong>mation temperature interval, growth rate of its<br />

particles, <strong>and</strong> many other parameters of martensite decomposition at tempering<br />

[49].<br />

Fig. 34 presents the calculation results of effective growth rate of Fe 3C<br />

nucleus at tempering of engineering alloyed steels. The calculations were<br />

carried out using expression [49]:<br />

VmaxR ¼ð27D=256pÞN ð34Þ<br />

where R is the cementite particle radius; N is the right part of Eq. (30). Manganese<br />

decreases martensite stability significantly promoting its decomposition<br />

at low temperatures. Silicon, at a concentration greater than 1%,<br />

activates martensite decomposition at 700–800K <strong>and</strong> inhibits it at lower<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 34 Change of effective growth rate of Fe 3C nucleus with alloying of 0.4% C<br />

steel: 1, unalloyed steel with 0.4C(mass%); 2, alloyed with 0.35Si; 3, alloyed with<br />

2.1Cr; 4, alloyed with 1.75Si; 5, alloyed with 2.43Mn.<br />

content. Chromium does not change the temperature of intensive cementite<br />

growth.<br />

2. Calculation of Thermokinetic Diagrams of Impurities’<br />

Segregation During Tempering of Steel<br />

<strong>Modeling</strong> of multicomponent adsorption kinetics is carried out using a<br />

sequence of computer calculations.<br />

At the initial stage, thermodynamic characteristics of surface activity<br />

in Fe-base binary <strong>and</strong> ternary alloys are determined.<br />

Analysis of phase equilibrium diagrams permits the determination of<br />

the impurity segregation energy Eseg<br />

i<br />

<strong>and</strong> temperature dependence of ulti-<br />

mate solubility X8c. These two parameters are very important <strong>for</strong> determination<br />

of equilibrium GB concentration of impurity X i b using Eq. (17).<br />

Examples of such calculations <strong>for</strong> binary <strong>and</strong> ternary alloys have been<br />

presented. Mutual influence of alloy components on their surface activity<br />

could be refined experimentally. The equations <strong>for</strong> binding impurity segregation<br />

energy with solid solution composition could be obtained by regression<br />

analysis of multicomponent adsorption diagrams. These experiments<br />

allow the determination of the effective diffusion coefficient of elements.<br />

The diagram of the equilibrium impurity concentration calculation on grain<br />

boundary in engineering steel is presented in Fig. 35.<br />

Carbon concentration in martensite changes drastically during tempering,<br />

as it depends on chemical composition of steel, temperature, <strong>and</strong> duration<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 35 Calculation scheme of equilibrium impurity GBS. X a i (0) is the initial<br />

concentration of ith element in the steel; Xa C (T,t) is the running carbon concentration<br />

in martensite during its tempering; Xb i (T) is the maximal equilibrium GBS of ith<br />

element; E i seg Fe–i is the segregation energy of ith element in two-component Fe–I<br />

alloy; E i seg Fe–i–j is the segregation energy of ith element in multicomponent alloy;<br />

Di(T) is the diffusion coefficient of ith element in austenite, martensite, <strong>and</strong> ferrite.<br />

of treatment. This factor influences on thermodynamic activity of all steel<br />

components <strong>and</strong> on their energy of GB segregation.<br />

The second important stage of GBS modeling includes calculation<br />

of C volume concentration in martensite Xa C (T), depending on steel<br />

chemical composition Xa i (0) <strong>and</strong> parameters of tempering. New segregation<br />

energy values of each element at changing of treatment temperature or<br />

time <strong>and</strong> new equilibrium GBS level have been calculated in this way (see<br />

Fig. 35).<br />

The final stage of modeling includes a set of independent calculations<br />

of various element diffusion to GB zone, <strong>and</strong> their desorption. The limited<br />

capacity of boundary <strong>and</strong> its effective width (about 0.5 nm) are shown. It is<br />

assumed that interstitial <strong>and</strong> substitial impurities occupy different positions<br />

on GB. Time t of reaching the definite concentration of impurity in segregation<br />

X b(t) at given temper temperature T is calculated by (22), <strong>and</strong> it is controlled<br />

by diffusion D i(T).<br />

Adsorption in multicomponent system is accompanied by concurrence:<br />

arrival of some surface-active impurity decreases GB energy <strong>and</strong>, in<br />

this way, the thermodynamic stimulus <strong>for</strong> segregation of other impurities.<br />

Dissolution of segregations is observed at increasing temperature. Impurity<br />

desorption to grain bulk is analogous to adsorption, however it is tied not<br />

with concentration Xi(0) but with Xb(t), <strong>and</strong> it is also controlled by diffusion<br />

D i(t).<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


The model is restricted to initially homogeneous bulk concentrations<br />

X i b ð0Þ ¼Xi a<br />

ð35Þ<br />

The kinetics of segregation to surfaces or grain boundaries from the<br />

bulk are determined by volume diffusion of impurities with bulk concentrations<br />

X i a(t) which can be treated as a one-dimensional problem. Since both<br />

bulk concentrations are very small, Arrhenius type diffusion coefficients:<br />

Di ¼ D i 0 exp<br />

Qi<br />

RT<br />

ð36Þ<br />

can be used which are independent of X i a(t). In the case of site competition,<br />

the GB impurities concentration is<br />

qi ¼<br />

Xi<br />

1 P<br />

J Xj<br />

exp<br />

Ei<br />

KT<br />

ð37Þ<br />

The equations describing the time evolution of segregation <strong>for</strong> homogeneous<br />

initial condition [60] are<br />

pffiffiffi pd<br />

XiðtÞ ¼Xið0Þþ2 X0 i<br />

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

Z t<br />

0<br />

Diðt 0 Þ dt 0<br />

1<br />

pffiffiffi pd<br />

Z t<br />

0<br />

qiðt0ÞDiðt0 Þ<br />

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

R t0 t Diðt00Þ dt00 q ð38Þ<br />

In the case of constant temperature (i.e. Di ¼ const), Eq. (38) can be<br />

simplified:<br />

pffiffiffiffi<br />

2 D<br />

XiðtÞ ¼Xið0Þþpffiffiffi ½X<br />

pd<br />

0 i<br />

pffiffi<br />

qiðtÞŠ t<br />

ð39Þ<br />

Diffusion coefficient <strong>for</strong> impurities in Fe <strong>and</strong> Fe-base alloys in ferrite<br />

interval is present in Table 6.<br />

The calculated diagrams of multicomponent adsorption in steels 0.3C–<br />

Cr–Mo, 0.3C–Cr–Mn–V, 0.3C–Cr–Mn–Si–Ti (see Table 3) are presented in<br />

Figs. 36–38. Comparing these diagrams with the experimental ones (Figs.<br />

24, 26, <strong>and</strong> 27), a good correlation of segregation kinetic features<br />

<strong>for</strong> various elements is observed, that confirms the basic principles of the<br />

proposed model of GBS in steels. According to this model, the main<br />

role of carbide precipitation in GBS consists of changing solid solution<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Table 6 Coefficients of Diffusion <strong>for</strong> Impurities in a-Fe <strong>and</strong> Steels<br />

Solute System Temperature, K D, m 2 Sec 1<br />

D 0 (m 2 Sec 1 )<br />

Q,<br />

kcal=mol Reference<br />

C 0.3C–10Ni<br />

Martensite<br />

723–873 5.26 10 5<br />

15.2 [50]<br />

C a-Fe 623 5.3 10 14<br />

[51]<br />

P a-Fe 723 2.8 10 19<br />

[52]<br />

P a-Fe 748 7.7 10 19<br />

[52]<br />

P a-Fe 773 2.0 10 18<br />

[52]<br />

P a-Fe 798 4.8 10 18<br />

[52]<br />

P a-Fe 9.55 10 6<br />

50.6 [53]<br />

P Fe–2.1Mn 1.43 10 4<br />

54.2 [53]<br />

P Fe–Ni–P 0.51 10 4<br />

55 [54]<br />

P a-Fe 0.108 exp( 288=RT) [55]<br />

P 0.1P–0.15Cr a 0.336 exp( 296=RT) [55]<br />

P 0.1P–0.13Si a 18 exp( 329=RT) [55]<br />

P 0.1P–0.17Mn a 0.235 exp( 292=RT) [55]<br />

P 0.1P–0.14Mo a 3.23 10 6 exp( 434=RT) [55]<br />

P 0.1P–0.14Ni a 43.9 exp( 336=RT) [55]<br />

S Fe–3Si 973 1.7 10 2<br />

61.2 [56]<br />

Sn a-Fe 973–1303 5.4 55.5 [57]<br />

Cr Fe–Cr 1048 2.33 10 4<br />

57.1 [58]<br />

Co Fe–6.8Co 903–1073 4.69 10 5<br />

44.7 [59]<br />

C Fe–0.79Si 803 3.6 10 12<br />

[60]<br />

C Fe–0.79Si 873 1.4 10 11<br />

[60]<br />

C Fe–0.79Si 973 1.9 10 10<br />

[60]<br />

C Fe–0.6Ni 803 4 10 4<br />

[60]<br />

C Fe–0.6Ni 873 16 10 4<br />

[60]<br />

C Fe–0.6Ni 973 6.9 10 4<br />

[60]<br />

C Fe–0.56Mo 873 7.2 10 4<br />

[60]<br />

C Fe–0.56Mo 923 21 10 4<br />

[60]<br />

C Fe–0.56Mo 973 55 10 4<br />

[60]<br />

composition, as it is exactly this factor that controls mutual influence of elements<br />

on their surface activity. Only such elements that segregate in near<br />

temperature ranges mutually influence GBS. The computer calculations of<br />

segregation kinetic diagrams predict these effects with small changes of steel<br />

chemical composition. Figures 39 <strong>and</strong> 40 present the modeling data on influence<br />

of sulfur content in 0.3C–Cr–Mn–Si–Ti steel on phosphorus segregation<br />

kinetics. Sulfur <strong>and</strong> Phosphorus are strong surface-active elements,<br />

<strong>and</strong> they can compete at grain boundaries. Desulfurization of steel significantly<br />

slows down GBS of S. Indeed, P adsorption increases with a decrease<br />

of S content. According to calculations (see Fig. 40), the time of 6% P GBS<br />

<strong>for</strong>mation exceeds 4000 sec at a sulfur content more than 0.02 at.%. This<br />

time it is significantly longer than the usual duration of quenched steel<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 36 The isodose C-curves of multicomponent interface segregation in 0.2C–<br />

Cr–Mn–Ni–Si steel (see Table 3) under its tempering. Computer simulation.<br />

tempering. Deeper cleaning of steel by S activates GB adsorption of P, drops<br />

down time of segregation <strong>for</strong>mation, <strong>and</strong> increases the maximum temperature<br />

of segregation stability (see Figs. 39 <strong>and</strong> 40). Such calculations<br />

are very useful <strong>for</strong> the design of optimal alloying <strong>and</strong> purification degree on<br />

harmful impurities, since they permit the determination of the influence of<br />

alloying on ultimate concentration of harmful impurities.<br />

Figure 37 The isodose C-curves of multicomponent interface segregation in 0.3C–<br />

Cr–Mn–V steel (see Table 3) under its tempering. Computer simulation.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 38 The isodose C-curves of multicomponent interface segregation in 0.3C–<br />

Cr–Mn–Si–Ti steel (see Table 3) under its tempering. Computer simulation.<br />

B. Interface Adsorption During Quenching of<br />

Engineering Steels<br />

Mathematical models of GBS [61] <strong>and</strong> phase trans<strong>for</strong>mations permit the<br />

analysis of heat treatment with respect to the accompanying phenomena<br />

in a greater detail than that of a simple summary of the experimental<br />

knowledge.<br />

Figure 39 Dependence of Tmax of P GBS as a function of sulfur containing in<br />

0.3C–Cr–Mn–Si–Ti steel under its tempering. Computer simulation.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 40 Dependence of time of 6 at.% GBS of phosphorus <strong>and</strong> sulfur as a<br />

function of sulfur concentration in 0.3C–Cr–Mn–Si–Ti steel during its tempering at<br />

700K. Computer simulation.<br />

The results of mathematical modeling provide backgrounds <strong>for</strong> reasonable<br />

planning of full-scale experiments when seeking <strong>for</strong> the optimum<br />

technological procedures <strong>and</strong> steel composition <strong>and</strong> they enable the<br />

extrapolation of the consequences of variations in the technological conditions<br />

even outside the boundary of the empirical experience we have available.<br />

Interaction of GB segregation enrichment <strong>and</strong> phase trans<strong>for</strong>mations<br />

during heat treatment of steels in the austenitic region is hard to imagine.<br />

Nb <strong>and</strong> V carbonitride precipitation in microalloyed austenite, precipitation<br />

of free ferrite, change chemical composition of austenite, <strong>and</strong> influence<br />

on GBS kinetics to a large extent. The experiments show that nonequilibrium<br />

grain boundary phenomena occur <strong>for</strong> a rather short time up<br />

to 100 sec. The minimum time of 5% volume fraction of Nb <strong>and</strong> V carbonitride<br />

precipitation is about 1000 sec [62,63]. Precipitation of free ferrite<br />

needs from several seconds to several minutes depending on steel chemical<br />

composition. There<strong>for</strong>e, the non-equilibrium GBS in steels with a wide<br />

region of undercooled austenite stability independently from phase trans<strong>for</strong>mations.<br />

This computer model has some limitations but redistribution<br />

of harmful impurities between grain bulk <strong>and</strong> boundaries permits the analysis<br />

of steel quenching.<br />

The modeling of non-equilibrium GB phenomena allows during investigation<br />

of such short-time changes of chemical composition that could not<br />

be measured experimentally <strong>and</strong> that has an extreme importance <strong>for</strong> modern<br />

heat-treatment processes with high heating <strong>and</strong> cooling velocities in<br />

controlled media.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 41 The presentation of C-curve on simulating TTT diagrams. (Scheme.)<br />

Parameters U <strong>and</strong> S correspond to Table 7.<br />

1. Phase Trans<strong>for</strong>mations of Undercooling Austenite<br />

At present, many computer models of evolution of structure <strong>and</strong> phase composition<br />

of steels during quenching have been developed. Most of them are<br />

based on physical models of phase trans<strong>for</strong>mations [64–66]. But physical<br />

models cannot describe adequately all kinetic features of undercooled austenite<br />

trans<strong>for</strong>mations. The computer models based on regression analysis of<br />

experimental data can best predict steel phase composition changes during<br />

steel cooling.<br />

It was introduced directly by Davenport <strong>and</strong> Bain [67] <strong>and</strong> the time–<br />

temperature-trans<strong>for</strong>mation (TTT) diagram was the predominant tool to<br />

describe the isothermal decomposition kinetics of supercooled austenite.<br />

In most TTT diagrams, general S- or C-curves are used to represent the<br />

kinetics of a number of isothermal trans<strong>for</strong>mation products: ferrite, pearlite,<br />

upper bainite, lower bainite, <strong>and</strong> martensite. Conversely, many experimental<br />

results demonstrate that each type of trans<strong>for</strong>mation product has a separate<br />

C-curve.<br />

To build a mathematical model, all TTT diagrams published in Refs.<br />

[68–71] were analyzed. The rationalization of the kinetics of isothermal<br />

decomposition of austenite permitted the establishment of a metastable<br />

product (phase) diagram of a number of steels of different compositions<br />

with 6% of total content of all alloying elements.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Most isothermal trans<strong>for</strong>mations take place by nucleation at the austenite<br />

grain boundaries, so the original austenite grain size will affect the isothermal<br />

decomposition kinetics of austenite.<br />

From the total number of factors characterizing austenite matrix, the<br />

present day experimental knowledge allows only an approximate examination<br />

of the statistically recrystallized proportion <strong>and</strong> estimation of the size<br />

of de<strong>for</strong>med austenite grains.<br />

The grain growth kinetics satisfy the law [73]<br />

dðtÞ ¼d0 þ kt exp<br />

Q<br />

RT<br />

ð40Þ<br />

where d(t) is averaged grain size at moment t; d0 is initial grain size; Q is activation<br />

energy; k is a constant.<br />

The algorithm of calculating the size of austenite grains is described in<br />

Ref. [73].<br />

The procedure <strong>for</strong> calculation of the structural proportions of anisothermal<br />

decomposition of austenite at engineering steel cooling is given<br />

in Tables 7 <strong>and</strong> 8 <strong>and</strong> shown in Fig. 41.<br />

The cooling curve is approximated partially by a constant function<br />

<strong>and</strong> at the individual time intervals Dt <strong>and</strong> the rate of decomposition is calculated<br />

as isothermal trans<strong>for</strong>mation corresponding to the mean temperature<br />

of that interval. The required kinetic data are available from the TTT<br />

diagrams [68–71] that can be digitized (see Table 8) by procedures shown<br />

in Fig. 41, using equation<br />

S S0<br />

SN S0<br />

1=2 U U0<br />

¼ e<br />

UN U0<br />

1=2<br />

exp<br />

1<br />

2<br />

U U0<br />

UN U0<br />

ð41Þ<br />

where S ¼ Int-time interval, s; U ¼ 1000=(T þ 273).<br />

Since it is necessary to distinguish between the parts of the C-curves<br />

representing the <strong>for</strong>mation of ferrite, pearlite, <strong>and</strong> bainite, only those<br />

diagrams having readily distinguishable component curves were used in<br />

the analysis.<br />

The calculation method includes the effect of the size of austenite<br />

grains on the kinetics of phase trans<strong>for</strong>mations. The main precondition is<br />

knowledge of this effect on the course of C-curves showing the start <strong>and</strong><br />

end of trans<strong>for</strong>mations in the graph of isothermal decomposition of austenite<br />

<strong>for</strong> the relevant steel.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Table 7 The Algorithm of Calculation of the Structural Proportions<br />

1. Temperature of start of trans<strong>for</strong>mations yAc3, yAc1, yBa, yMs<br />

2. t 0; y(0) y0; V1(0) 1; Vi(0) 0, i ¼ 2, 3, 4, 5; i ¼ 1-austenite,<br />

2-ferrite, 3-pearlite, 4-bainite, 5-martensite<br />

2.1 Mean temperature at the interval of ht, t þ Dti<br />

y ¼ (y(t) þ y(t þ Dt))=2;<br />

if y < ¼ yMs pass to 3;<br />

if y < ¼ min (yjs) then n j;<br />

2.2 <strong>for</strong> i ¼ 2, ..., n carried out as follows:<br />

– calculation of the trans<strong>for</strong>mable proportion of austenite<br />

Vmi(t) <strong>for</strong> i ¼ 2:<br />

Vm2(t)¼0; <strong>for</strong> y ¼ > yA3,<br />

Vm2ðtÞ ¼V 0 yA3 y<br />

m2 yA3 yA1 ; <strong>for</strong> yA1 < y < yA3;<br />

; <strong>for</strong> y 2 <strong>and</strong> Vmi ¼ 1<br />

– calculation of the start <strong>and</strong> end of trans<strong>for</strong>mation tsi, tfi <strong>and</strong> exponent<br />

ki <strong>for</strong> y<br />

ki ¼ 6.127=ln (tsi=tfi); <strong>for</strong> ferrite k2 ¼ 1;<br />

– the fictive volume fraction of the trans<strong>for</strong>med proportion<br />

Xi ¼ Vi(t)=[Vi(t) þ Vi(t) Vmi(t)]; – the fictive time of isothermal trans<strong>for</strong>mation required <strong>for</strong> reaching<br />

the proportion Xi<br />

t 0 i ¼<br />

tki lnð1 XiÞ<br />

si<br />

bS 1=ki;<br />

– the fictive volume proportion of the structural component at time<br />

t þ Dt<br />

Xiðti þ DtÞ ¼1 exp b2 tiþDt<br />

tsi<br />

;<br />

– the volume proportion of the structural component at time t þ Dt<br />

Viðt þ DtÞ ¼Xiðt0 i þ DtÞ½ViðtÞþViðtÞVmiðtÞŠ;<br />

2.3 the new value of residual content of austenite<br />

V1ðt þ DtÞ ¼1 Pn i¼2 Viðt þ DtÞ; if V1ðt þ DtÞ


Table 8 The Constants Blk <strong>for</strong> Calculations of the YL Parameters of C-Curves<br />

Trans<strong>for</strong>mation YL<br />

k<br />

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16<br />

B lk<br />

Const C Mn Si Ni Cr Mo GS Const C Mn Si Ni Cr Mo GS<br />

Ferrite-start T8C T OFS 727 13 9 17 22 3.5 0 229 30 66 22 70 5 0<br />

T NFS 572 33 44 9 25 50 1.01 48 57 25 15 15 25 3.90<br />

Time IntOFS 10.13 2.5 2.84 0.46 4.8 5.9 0.095 1.92 2.80 3.55 1.50 3.00 3.50 0.0096<br />

IntNFS 1.04 0.6 0.19 0.07 4.8 5.9 0.038 2.58 0.20 1.10 0.04 2.55 3.50 0.021<br />

Pearlite-start T8C TOPS 727 13 9 17 22 3.5 0 0 15 5 5 22 0<br />

T NPS 572 33 44 9 25 50 1.01 48 32 20 5 27 23 3.20<br />

Time Int OPS 10.13 2.5 2.84 0.46 4.8 5.9 0.095 1.92 10.12 3.55 3.30 3.00 4.20 0.096<br />

Int NPS 1.04 0.6 0.19 0.07 4.8 5.9 0.038 2.58 1.31 0.20 0.20 2.55 2.50 0.041<br />

Pearlite-finish T8C TOPF 727 13 9 17 22 3.5 0 0 0 15 5 5 22 0<br />

TNPF 577 2 35 7 53 58 0.52 6 4 5 5 36 25 1.31<br />

Time IntOPF 10.55 6.86 3 1.81 4.8 12.3 0.095 0.03 14.08 0.76 1.99 3.00 4.35 0.0096<br />

IntNPF 0.15 2.0 0.19 0.07 2.65 10.6 0.038 0.03 1.31 1.56 0.65 0.98 4.35 0.021<br />

Bainite-start T8C T OBS 570 12 16 12 20 22 0 107 12 –5 –5 –5 –40 0<br />

T NBS 485 12 2 7 40 32 0.50 107 30 35 5 36 2 0.37<br />

Time Int OBS 3.52 2.14 1.07 0.05 2.3 3.1 0.070 5.44 1.46 1.64 0.40 3.49 3.00 0.0626<br />

IntNBS 0.96 1.30 0.30 0.02 2.2 2.2 0.024 1.40 1.29 1.64 0.20 2.04 6.60 0.0248<br />

Bainite-finish T8C TOBF 570 12 16 7 22 22 0 76 12 5 7 5 51 0<br />

TNBF 488 48 25 7 47 61 0.52 76 30 28 7 40 96 0.37<br />

Time Int OBF 6.90 5.23 2.8 0.7 6.2 9.5 0.070 7.73 0.75 5.44 0.40 0.90 6.88 0.0813<br />

Int NBF 0.15 3.00 2.8 0.7 3 5.9 0.024 1.98 1.17 5.84 0.02 2.47 4.00 0.0340<br />

M S T MS 539 423 30.4 0 17.7 12.1 7.5 0 0 0 0 0 0 0 0<br />

Ck<br />

C1 C2 C3 C4 C5 C6 C7 C8<br />

1 %C %Mn %Si %Ni %Cr %Mo 2<br />

YL ¼ P8 k¼1 Blk ½ þ Dlk þ 8ð0:8 C2ÞŠCk;<br />

GS—grain size (ASTM).<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.<br />

D lk


The program involves the calculation of temperatures of trans<strong>for</strong>mations<br />

of bainite <strong>and</strong> twinned, athermal <strong>and</strong> lamellar martensite [74]<br />

BS ¼ 720 585:63ðCÞþ126:6ðCÞ 2<br />

31:66ðCrÞþ2:17ðCrÞ 3<br />

42:37ðMoÞþ9:16ðCoÞ 0:125ðCoÞ 2<br />

66:34ðNiÞþ6:06ðNiÞ 2<br />

91:68ðMnÞþ7:82ðMnÞ 2<br />

M TM<br />

S ¼420 208:33ðCÞ 72:65ðNÞ 43:36ðNÞ2 16:08ðNiÞ<br />

M LM<br />

S<br />

M A S<br />

þ0:78ðNiÞ 2<br />

36:02ðCuÞ ð42Þ<br />

0:025ðNiÞ 3 2:47ðCrÞ 33:428ðMnÞþ1:296ðMnÞ 2<br />

þ30:0ðMoÞþ12:86ðCoÞ 0:2665ðCoÞ 2 7:18ðCuÞ ð43Þ<br />

¼ 540 356:25ðCÞ 260:64ðN 24:65ðNiÞþ1:36ðNiÞ2<br />

17:82ðCrÞþ1:42ðCrÞ 2<br />

47:59ðMnÞþ2:25ðMnÞ 2<br />

þ 17:5ðMoÞþ21:87ðCoÞ 16:52ðCuÞ ð44Þ<br />

¼ 820 603:76ðCÞþ247:13ðCÞ2<br />

31:1ðCrÞþ2:348ðCrÞ 2<br />

0:196ðCoÞþ0:165ðCoÞ 2<br />

55:72ðNiÞþ3:97ðNiÞ 2<br />

66:24ðMnÞ 24:29ðMoÞ<br />

The size of ferritic grain is expressed as follows [75]<br />

da ¼ 11:7 þ 0:14dg þ 37:7V 0:5<br />

C<br />

31:88ðCuÞ ð45Þ<br />

ð46Þ<br />

where dg is the size of austenitic grain, (mm); VC is the cooling speed,<br />

(8Cmin 1 ).<br />

The interlamellar distance of pearlite can be estimated as follows [75]:<br />

S ¼ X<br />

"<br />

18:0DVPðyiÞ=ð996<br />

#<br />

yiÞ =VP<br />

ð47Þ<br />

i<br />

where DVP(yi) is the volume proportion of pearlite trans<strong>for</strong>med at yi temperature.<br />

The thickness of the ferritic <strong>and</strong> carbide lamellae of pearlite is approximately<br />

lf ¼ 0.885S; lc ¼ 0.115S<br />

The size of martensitic <strong>and</strong> bainitic particles is identical with the<br />

original size of the austenite grains. The course of the anisothermal<br />

decomposition of austenite in several steels has been calculated by the<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


just-described method <strong>and</strong> by applying the digitized TTT diagrams of<br />

Table 7.<br />

2. Determination of the Kinetics of Carbonitride<br />

Precipitation in Austenite<br />

Microalloying of steels with Ti, V, Nb, <strong>and</strong> Zr affects decomposition of<br />

supercooled austenite, its recrystallization <strong>and</strong> grain boundary segregations<br />

of harmful impurities. These changes of material properties are associated<br />

with carbonitride precipitation <strong>and</strong> changes of austenite chemical composition.<br />

Based on these reasons, the modeling of the kinetics of carbonitride<br />

precipitation is important.<br />

The nucleation time t of carbonitrides per unit volume N at any temperature<br />

T, can be expressed as [76]<br />

t ¼ C exp Q<br />

RT exp<br />

B<br />

T3ðLnKSÞ 2<br />

!<br />

t ¼ C exp Q<br />

RT exp<br />

B<br />

T3ðLnKSÞ 2<br />

!<br />

ð48Þ<br />

where C ¼ 6 10 13 <strong>for</strong> homogeneous nucleation; activation energy of Nb<br />

diffusion Q ¼ 270 kJ=mol;<br />

B ¼ 16pg 3 V 2 3<br />

mN0=3R ð49Þ<br />

Vm ¼ 1.28 10 5 m 3 =mol; g ¼ 0.5 J m 2 ; N0 are numbers of nucleus by<br />

radius R per molar volume Vm; KS is supersaturation [77]<br />

LgðKSÞ ¼ A<br />

þ B ð50Þ<br />

T<br />

where thermodynamics parameters A <strong>and</strong> B <strong>for</strong> various carbides <strong>and</strong><br />

nitrides are calculated in Ref. [78] <strong>and</strong> presented in Table 9. Thus, the calculation<br />

of carbonitride nucleation time necessary to reproduce the C-curves<br />

Table 9 Thermodynamics Parameters in Eq. 50 (From Ref. 77)<br />

Chemical compound<br />

Parameter AlN VC VN TiC TiN NbC NbN ZrC ZrN<br />

A 7,130 9,500 7,985 8,872 15,573 7,714 10,440 8,464 13,968<br />

B 1.463 6.72 3.09 4.04 3.82 3.27 3.87 4.96 3.08<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


corresponds to start or finish of this trans<strong>for</strong>mation in austenite under heat<br />

treatment of steel.<br />

3. Calculation of Thermokinetic Diagrams of Impurity<br />

Segregation During Quenching of Steel<br />

Computation of grain boundary multicomponent adsorption kinetics could<br />

be simplified <strong>for</strong> steels with high undercooled austenite stability. The GBS<br />

develops in this case in austenite in short time <strong>and</strong> has no dependence on<br />

phase <strong>and</strong> structure trans<strong>for</strong>mations at steel quenching. Enrichment of grain<br />

boundaries by various impurities as well as their desorption is treated as a<br />

result of multicomponent diffusion of impurities from near-boundary<br />

volume to the boundary. Impurity binding energy with GB includes<br />

mutual influence of elements in grain bulk <strong>and</strong> on the boundary in accordance<br />

with Guttmann’s theory [Eqs. (18) <strong>and</strong> (19)]. Auger electron spectroscopy<br />

is the technique <strong>for</strong> experimental investigation of GBS kinetics. These<br />

experiments are basic <strong>for</strong> analysis of correlation of impurity segregation<br />

energy with the content of other elements in the bulk <strong>and</strong> on boundaries<br />

(see Section 2.5, Eqs. (23)–(29).<br />

Adsorption <strong>and</strong> desorption of impurities on GB (qi) at steel quenching<br />

is modeled well using the equation<br />

qi ¼ qið0Þþ 2q0 i<br />

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

Z t<br />

pffiffiffiffiffi<br />

Diðt<br />

pd<br />

0Þ dt0 s<br />

1<br />

pffiffiffiffiffi<br />

pd<br />

0<br />

Z t<br />

0<br />

C i a ðt0 ÞDiðt 0 Þ<br />

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

R t<br />

t0 Diðt00Þ dt00 q dt 0<br />

ð51Þ<br />

where d is the grain boundary thickness; Di(t 0 ) is the diffusion coefficient of<br />

impurity which depends on the temperature <strong>and</strong> phase composition (austenite,<br />

martensite, <strong>and</strong> ferrite); in the case of adsorption<br />

C i a ¼<br />

i<br />

where CGB<br />

Ci GB<br />

1 P<br />

j Cj<br />

expð Gi=kTÞ ð52Þ<br />

GB<br />

is the element i concentration on grain boundary; Ca i is the con-<br />

centration of ith element in the adjacent bulk layer; Gi is segregation energy.<br />

Desorption is determined by GB concentrations of impurities, <strong>and</strong> in this<br />

case, the parameter C a i in Eq. (51) is equivalent to GB concentration Xb I in<br />

Eqs. (12) <strong>and</strong> (13).<br />

The change of temperature at cooling or isothermal exposition is<br />

described by equation<br />

TðtÞ ¼½Tð0Þ Tð/ÞŠ expð rtÞþTð/Þ ð53Þ<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Table 10 Cooling Rates <strong>for</strong> Some Metallurgical Technologies<br />

Name of the treatment Cooling rates r (K sec 1 )<br />

Quenching 100–10<br />

Controlled cooling 10<br />

Air cooling of hot-rolled metal 10–0.1<br />

Cooling with furnace 0.01<br />

Controlled cooling of large-size <strong>for</strong>ging 0.001<br />

Table 10 presents cooling rates r <strong>for</strong> heat-treatment processes. The block<br />

diagram of multicomponent intercrystalline adsorption model is shown in<br />

Fig. 42. Adsorption of P, C, <strong>and</strong> S is determined by parameters K1, K2,<br />

K3, <strong>and</strong> their desorption by parameters K2, K4, K6. The parameters Ci<br />

are equivalent to GB concentration of element i. This model allows the computation<br />

of the condition when there is change of GB composition in steels<br />

<strong>and</strong> alloys at preselected arbitrary mode of cooling including isothermal<br />

exposition.<br />

Given below are the examples of investigation of phosphorus <strong>and</strong> sulfur<br />

grain boundary adsorption in Cr–Ni–Mo steel (see Table 11).<br />

The components of steel mutually influence their diffusion mobility<br />

<strong>and</strong> GBS activation energy. Based on this reason, one should take into<br />

account the stochastic fluctuations of diffusion flows of various impurities<br />

on GBS kinetics. For this purpose, the r<strong>and</strong>om fluctuation of diffusion<br />

coefficients up to 30% of its mean value was used in the model. Figures 43<br />

<strong>and</strong> 44 present the GBS kinetics calculation results at cooling of various purity<br />

steels cooling that were carried out using the stochastic model. As one<br />

can see, the self-regulation of adsorption is observed which is developing<br />

despite significant short-time oscillations of impurity concentration on grain<br />

boundaries. The significant non-equilibrium enrichment of GB by impurities<br />

is observed at initial stage of the heat treatment. This effect is determined<br />

by cooling velocity as well as impurities content. Increasing cooling<br />

velocity from 0.001 to 1000K s 1 decreases the non-equilibrium GBS of P<br />

<strong>and</strong> S. Formation of non-equilibrium rich GBS of harmful impurities at<br />

small cooling times could be established only by using computer modeling<br />

methods. The experimental verification of such phenomena needs special<br />

techniques which allow to open grain boundaries: hydrogenation of<br />

quenched samples or delayed fracture tests. Since these techniques are<br />

conducted in air <strong>and</strong> could not be applied in the vacuum chamber of<br />

electron spectrometer; <strong>for</strong> most of engineering steels, the regularities of<br />

non-equilibrium GBS <strong>for</strong>mation at quenching could only be estimated by a<br />

computer experiment.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 42 Calculation scheme of three-component GBS (phosphorus, sulfur, <strong>and</strong><br />

carbon).<br />

Table 11 Chemical Composition of Cr–Ni–Mo Steels<br />

Smelting<br />

number<br />

Chemical composition, mass%<br />

C S P Ni Cr Mo<br />

Time of austenite<br />

stability at 600 hr<br />

82 0.38 0.027 0.054 3.95 3.0 0.51 2.0<br />

83 0.38 0.01 0.006 4.02 3.0 0.50 2.0<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 43 The change of GBS during quenching of Cr–Ni–Mo steel containing<br />

0.027S <strong>and</strong> 0.054P (mass%). Computer simulation of fast (a) <strong>and</strong> slow (b) cooling.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 44 The change of GBS during quenching of Cr–Ni–Mo steel containing<br />

0.01S <strong>and</strong> 0.006P (mass%). Computer simulation of fast (a) <strong>and</strong> slow (b) cooling.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 45 Chemical composition of GB in Cr–Ni–Mo steel containing 0.027S <strong>and</strong><br />

0.054P (mass%) after austenitization at 1373K (30 min), interim cooling up to 873K<br />

<strong>and</strong> quenching in water (a) <strong>and</strong> in furnace (b). Auger electron spectroscopy of<br />

intergranular fracture.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


Figure 46 Chemical composition of GB in Cr–Ni–Mo steel containing 0.01S <strong>and</strong><br />

0.006P (mass%) after austenitization at 1373K (30 min), interim cooling up to 873K<br />

<strong>and</strong> quenching in water (a) <strong>and</strong> in furnace (b). Auger electron spectroscopy of<br />

intergranular fracture.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


The validation of calculation reliability was done <strong>for</strong> steel composition<br />

82 <strong>and</strong> 83 (see Table 10) by Auger spectroscopy. The samples after austenitization<br />

at 1373K (30 min) were in the interim cooled to 873K with further<br />

cooling in water or with furnace cooling. The undercooled austenite in this<br />

steel has high stability <strong>and</strong> does not trans<strong>for</strong>m in ferrite region <strong>for</strong> 2 hr.<br />

After cooling the samples had martensite–baintite structure. To investigate<br />

the chemical composition of grain boundaries by Auger spectroscopy, special<br />

samples were crushed in the electron spectrometer ESCALAB MK2 at<br />

vacuum at about 10 8 Pa at temperature 83K. The fields with intercrystalline<br />

fracture type were investigated on the fracture surface. The variation<br />

of phosphorus <strong>and</strong> sulfur content in GBS in Cr–Ni–Mo steel of several melts<br />

after heat treatment is shown in Figs. 45 <strong>and</strong> 46. At accelerated cooling the<br />

GB are significantly enriched by carbon. The P concentration in GB<br />

increases only at slow cooling of samples, <strong>and</strong> P segregation is strongly suppressed<br />

in pure steel. A good correspondence of calculated <strong>and</strong> experimental<br />

results is observed <strong>for</strong> all cases to be analyzed.<br />

The results of numerical modeling give in<strong>for</strong>mation about the equilibrium<br />

<strong>and</strong> non-equilibrium character of a GB adsorption processes, which<br />

are frequently unavailable from experiments. Moreover, these simulation<br />

methods explain the phenomenon of reverse temper embrittlement as the<br />

result of non-equilibrium concurrent GBS of carbon <strong>and</strong> phosphorus. These<br />

results explain many questions in the multicomponent GB adsorption<br />

kinetics in engineering steels that were dynamically developed in the last<br />

10 years. Further investigations in this direction are required especially<br />

<strong>for</strong> competitive internal adsorption in engineering steels treated by using<br />

newest schemes of heat treatment.<br />

REFERENCES<br />

1. Briant, C.L.; Banerji, S.K. Intergranular failure in steel. Int. Met. Rev. 1978,<br />

4, 164–196.<br />

2. Arharov, V.I.; Ivanovskaya, S.I.; Kolesnikova, K.M.; Farafonova, T.A. The<br />

nature of phosphorous influence on temper embrittlement. Fiz. Met.<br />

Metallioved. 1956, 2, 57–65.<br />

3. Hondros, E.D.; Seah, M.P. Segregation to interfaces. Int. Met. Rev. 1977, 22,<br />

12,261–12,303.<br />

4. Seah, M.P. Grain boundary segregation. J. Phys. F. 1980, 10 (6), 1043–1064.<br />

5. Guttmann, M. Equilibrium segregation in ternary solution: a model <strong>for</strong> temper<br />

embrittlement. Surf. Sci. 1975, 53, 213–227.<br />

6. Kaminskii, E.Z.; Stelletskaya, T.I. Kinetic of martensite dissolution in carbon<br />

steel. Problems of Fisical Metallurgy; Metallurgy: Moscow, 1949, 192–210.<br />

7. Bokshtein, S.Z. Structure <strong>and</strong> <strong>Mechanical</strong> Properties of Alloyed Steel;<br />

Metallurgy: Moscow 1954.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


8. Briant, C.L.; Banerji, C.K. Intergranular failure of ferrum alloys in inagressive<br />

environment. In Treatise on <strong>Material</strong>s Science <strong>and</strong> Technology, Vol. 25;<br />

Embrittlement of Engineering Alloys; Briant, C.L., Banerji, S.K., Eds.;<br />

Academic Press: 1983; 29–59.<br />

9. Erhart, H.; Paju, M. Phosphorus segregation in austenite. Scripta Met. 1983,<br />

17, 171–174.<br />

10. Kovalev, A.I. Influence of grain boundaries phosphorus segregation <strong>and</strong> steel<br />

microstructure on fracture after quenching <strong>and</strong> middle tempering. Fiz. Met.<br />

Metalloved. 1980, 49, 818–826.<br />

11. Glikman, E.; Cherpakov, Ju.; Bruver, R. Dependence of fracture toughness <strong>and</strong><br />

surface energy of grain boundaries on grain size of Si-steel under reversible<br />

temper embrittlement. Fiz. Met. Metalloved. 1978, 42, 864–870.<br />

12. Howe, H.M. Proc. Inst. Mech. Eng. 1919, Jan–May, 405.<br />

13. Carr, F.L.; Goldman, M.; Jaffe, L.D.; Buffin, D.C. Trans. AIME. 1953, 197,<br />

998.<br />

14. McLean, D. Grain Boundaries in Metals; Ox<strong>for</strong>d University Press: London<br />

1957.<br />

15. Hondros, E.D.; Seah, M.P. The theory of grain boundary segregation in terms<br />

of surface adsorption analogues. Met. Trans. A 1977, 8, 1363.<br />

16. Abraham, F.F.; Brundle, C.R. Surface segregation in binary solid solutions: a<br />

theoretical <strong>and</strong> experimental perspective. J. Vac. Sci. Technol. 1981, 18 (2),<br />

506–519.<br />

17. Abraham, F.F. Bond <strong>and</strong> strain energy effects in surface segregation. Scripta<br />

Met. 1979, 13 (5), 307–311.<br />

18. Fowler, R.H.; Guggenheim, E.A. Statistical Thermodynamics; University Press:<br />

Cambridge 1939.<br />

19. Grabke, H.J. Adsorption, segregation <strong>and</strong> reactions of non-metal atoms on<br />

iron surfaces. Mat. Sci. Eng. 1980, 42, 91–99.<br />

20. Grabke, H.J. Surface <strong>and</strong> grain boundary segregation on <strong>and</strong> in iron. Steel Res.<br />

1986, 57, 4178–4185.<br />

21. Grabke, H.J. Surface <strong>and</strong> Grain Boundary segregation on <strong>and</strong> in Iron <strong>and</strong><br />

Steels. ISIJ Int. 1989, 29, 7,529–7,538.<br />

22. Kovalev, A.I.; Mishina, V.P.; Stsherbedinsky, G.V.; Wainstein, D.L. EELFS<br />

method <strong>for</strong> investigation of equilibrium segregation on surfaces in steel <strong>and</strong><br />

alloys. Vacuum. 1990, 41 (7–9), 1794–1795.<br />

23. Bruver, R.E. Investigation of impurities influence on intergranular low<br />

temperature failure of metallic solid solution. Ph.D. Dissertation, Tomsk, 1970.<br />

24. Chelikowsky, J.R. Predictions <strong>for</strong> surface segregation in intermetallic alloys.<br />

Surf. Sci. 1984, 139, L197–L203.<br />

25. Miedema, A.R.; Boer, F.R. et al. Enthalpy of <strong>for</strong>mation of transition metal<br />

alloys. Calphad 1977, 1, 4341–4359.<br />

26. Miedema, A.R. On the heat of <strong>for</strong>mation of plutonium alloys. In Plutonium <strong>and</strong><br />

Other Actinides; Blank, H.; Linder, R., Eds.; North-Holl<strong>and</strong> Publ.: Amsterdam,<br />

1976; 3–20.<br />

27. Bennett, L.H.; Watson, R.A. A database <strong>for</strong> enthalpies of <strong>for</strong>mation of binary<br />

transition metal alloys. Calphad 1980, 5 (1), 19–23.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


28. Watson, R.E. Optimized prediction <strong>for</strong> heats of <strong>for</strong>mation of transition metal<br />

alloys. Calphad 1981, 5 (1), 25–60.<br />

29. Birnie, D.; Machlin, E.S.; Kaufman, C.; Taylor, K. Comparison of pair<br />

potential <strong>and</strong> thermochemical models of the heat of <strong>for</strong>mation of BCC <strong>and</strong><br />

FCC alloys. Calphad 1982, 6 (2), 93–126.<br />

30. Machlin, E.S. Correlation terms to pair potential model values of the energy of<br />

<strong>for</strong>mation <strong>for</strong> transition element–polyvalent element phases. Calphad 1980, 5<br />

(1), 1–17.<br />

31. Zadumkin, S.N. Modern theory of surface energy of pure metals. In. Surface<br />

Phenomena in Liquid metals <strong>and</strong> Solid Phases; Kabardino-Balkarskoe Publ.<br />

Press: Nalchik 1965; 12–28.<br />

32. Zadumkin, S.N. Surface energy on the interface of metals. In Surface<br />

Phenomena in Liquid metals <strong>and</strong> Solid Phases; Kabardino-Balkarskoe Publ.<br />

Press: Nalchik, 1965; 79–88.<br />

33. Morita, Z.; Tanaka, T. Effect of solute-interaction on the equilibrium<br />

distribution of solute between solid <strong>and</strong> liquid phases in iron base. Trans.<br />

ISIJ 1984, 24, 206–211.<br />

34. Kamenetskaya, D.S. Influence of molecular interaction on phase diagrams.<br />

Problems of Physical Metallurgy <strong>and</strong> Metal Science; Metallurgia: Moscow 1949;<br />

113–131.<br />

35. Toshihiro, T. Equilibrium distribution coefficient of P in Fe-base alloys. J. Iron<br />

Steel Inst. Jpn. 1984, 70 (4), 220.<br />

36. Guttmann, M.; Dumolin, Ph.; Wayman, M. The thermodynamics of interactive<br />

co-segregation of phosphorous <strong>and</strong> alloying elements in iron <strong>and</strong> temper-brittle<br />

steels. Met. Trans. A 1982, 13, 1693–1711.<br />

37. Morito, Z.-I.; Tanaka, T. Effect of solute-interaction on the equilibrium<br />

distribution of solute between solid <strong>and</strong> liquid phases in iron base ternary<br />

system. Trans. ISIJ. 1984, 24, 206–211.<br />

38. Morito, Z.-I.; Tanaka, T. Equilibrium distribution coefficient of phosphorus in<br />

iron alloys. Trans. ISIJ 1986, 26, 114–120.<br />

39. Okamoto, T.; Morito, Z.; Kagawa, A.; Tanaka, T. Partition of carbon<br />

between solid <strong>and</strong> liquid in Fe–C binary system. Trans. ISIJ 1983, 23,<br />

266–271.<br />

40. Yamada, K.; Kato, E. Effect of dilute concentrations of Si, Al, Ti, V, Cr, Co,<br />

Ni, Nb <strong>and</strong> Mo on the activity coefficient of P in liquid iron. Trans. ISIJ 1983,<br />

23, 51–55.<br />

41. Michina, V.P. Atomic interaction at grain boundary segregation in Fe alloys.<br />

Author’s Abstract of Doctor’s Thesis, Moscow, 1988.<br />

42. Seah, M.P. Grain boundary segregation <strong>and</strong> the Tt dependence of temper<br />

brittleness. Acta Met. 1977, 25, 345–357.<br />

43. Bannih, O.A.; Budberg, P.B.; Alisova, C.P. Equilibrium diagrams of two- <strong>and</strong><br />

poly-component systems on base of Fe. H<strong>and</strong>book; Metallurgia: Moscow 1986.<br />

44. Arharov, V.I.; Konstantinova, T.S. The nature of reversible temper embrittleness<br />

of 0.35C–Cr–Mn–Si <strong>and</strong> 12C–Cr–Ni steels. Fiz. Met. Metalloved. 1974, 38<br />

(1), 169–175.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


45. Otani, H.; Feng, H.C.; McMahon, C.J. New in<strong>for</strong>mation on the mechanism<br />

of temper embrittlement of alloy steels. Met. Trans. 1974, 5 (2), 516–518.<br />

46. Lei, T.C.; Tang, C.H.; Su, M. Temper brittleness of 0.3C–Cr–Mn7–Si–2Ni steel<br />

with various initial microstructures. Metal Sci. 1983, 17, 75–79.<br />

47. Kovalev, A.I.; Michina, V.P. Role of grain boundary segregation in RTE of<br />

steels. Physical basis of construct of physical <strong>and</strong> mechanical properties of steels<br />

<strong>and</strong> alloys. In Proceedings of CNIICHERMET; Bardin, I.P., Ed., Metallurgia:<br />

Moscow, 1990, 43–46.<br />

48. Kovalev, A.I.; Mishina, V.P.; Stsherbedinsky, G.V. Features of intergranular<br />

adsorption of carbon <strong>and</strong> phosphorus in Fe-alloys. Fiz. Met. Metalloved 1986,<br />

62 (1), 126–132.<br />

49. Zuyao, X.; Siwei, C. Mechanism of embrittlement in tempered martensite. Mat.<br />

Sci. Technol. 1985, 1, 1025–1028.<br />

50. Zemskii, S.V.; Litvinenko, D.A. Diffusion of C in two-phase system. Fiz. Met.<br />

Metalloved 1971, 32 (3), 591–596.<br />

51. Golikov, V.M.; Matosyan, M.A.; Estrin, E.I. Influence of pressure on carbon<br />

diffusion. Protect. Coat. Met. 1971, 4, 74–78.<br />

52. Guttmann, M.; Dumolin, Ph.; Wayman, M. The thermodynamics of interactive<br />

co-segregation of phosphorus <strong>and</strong> alloying elements in iron <strong>and</strong> temper-brittle<br />

steels. Met. Trans. A 1982, 13, 1693–1711.<br />

53. Grabke, H.J.; Hennesen, K.; Moller, R.; Wei, W. Effect Mn on the grain<br />

boundary segregation, bulk <strong>and</strong> grain boundary diffusitivity of P in ferrite.<br />

Scr. Met. 1987, 21 (10), 1329–1334.<br />

54. Yongbin, I.M.; Daniluk, St. A surface segregation study in P <strong>and</strong> S-doped type<br />

304 stainless steel. Met. Trans. A. 1987, 18 (1), 19–26.<br />

55. Mastsuyama, T.; Hosokawa, H.; Suto, H. Tracer diffusion of P in iron alloys.<br />

Trans. JIM 1983, 24 (8), 589–594.<br />

56. Gruzin, P.L.; Mural, V.V. S -diffusion in 3%Si–Fe alloy. Fiz. Met. Metalloved<br />

1971, 32 (1), 208–212.<br />

57. Treheus, D.; Marchive, D.; Delagrange, J. Determination of the coefficient of<br />

diffusion of Sn of infinite dilution in a-Fe. Compt. Rend. Acad. Sci. C 1972, 274<br />

(13), 1260–1262.<br />

58. Huntz, A.M.; Guiraldeng, P.; Aucouturier, M.; Lacombe, P. Relation between<br />

the diffusion of radioactive Fe <strong>and</strong> Cr in Fe–Cr alloys with 0–15% Cr <strong>and</strong> their<br />

a=g trans<strong>for</strong>mation. Mem. Sci. Rev. Met. 1969, 66, 85–104.<br />

59. Hirano, K.; Cohen, M. Diffusion of Co in Fe–Co alloys. Trans. JIM 1972, 13<br />

(2), 96–102.<br />

60. Gruzin, P.L.; Babikova, Yu.F.; Borisov, E.B.; Zimskii, S.V., et al. Investigation<br />

of diffusion of C in steel <strong>and</strong> alloys by C14 isotope. Problems of Metals Science<br />

<strong>and</strong> Physical Metallurgy; Metallurgy: Moscow 1958; 327–365.<br />

61. Militzer, M.; Wieting, J. Theory of segregation kinetics in ternary systems. Acta<br />

Met. 1986, 34 (7), 1229–1236.<br />

62. Dutto, B.; Sellars, C.M. Effect of composition <strong>and</strong> process variables on<br />

Nb(CN) precipitation in niobium microalloyed austenite. Mat. Sci. Technol.<br />

1987, 3 (3), 197–206.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.


63. Rios, P.R. Expression <strong>for</strong> solubility product of niobium carbonitride in<br />

austenite. Mat. Sci. Technol. 1988, 4 (4), 324–327.<br />

64. Adrion, H. Thermodynamic model <strong>for</strong> precipitation of carbonitrides in high<br />

strength low alloy steels containing up to three microalloying elements with or<br />

without additions of aluminum. Mat. Sci. Technol. 1992, 8 (5), 406–420.<br />

65. Reed, R.C.; Bhadeshia, H.K.D.H. Kinetics of reconstructive austenite to ferrite<br />

trans<strong>for</strong>mation in low alloy steels. Mat. Sci. Technol. 1992, 8, 421–436.<br />

66. Rees, G.I.; Bhadeshia, H.K.D.H. Bainite trans<strong>for</strong>mation kinetics Part 1. Mat.<br />

Sci. Technol. 1992, 8, 985–993.<br />

67. Davenport, E.S.; Bain, E.C. Trans. AIME. 1930, 90, 117–144.<br />

68. Isothermal Trans<strong>for</strong>mation Diagrams, 1943, 1st Ed., 1963, 2nd Ed.; United<br />

States Steel Corporation: Pittsburgh, PA.<br />

69. Atlas of Isothermal Trans<strong>for</strong>mation Diagrams of BS En Steels, 1949, 1st Ed.,<br />

1956, 2nd Ed.; The Iron <strong>and</strong> Steel Institute, British Iron <strong>and</strong> Steel Research<br />

Association: London.<br />

70. Popov, A.A.; Popova, L.E. Isothermal <strong>and</strong> Thermokinetics Diagrams of<br />

Decomposition of Supercooled Austenite; Metallurgia: Moscow 1965.<br />

71. Atlas of Isothermal Trans<strong>for</strong>mation <strong>and</strong> Cooling Trans<strong>for</strong>mation Diagrams;<br />

ASM:Metals Park, OH: 1977.<br />

72. BISRA Atlas of Isothermal Trans<strong>for</strong>mation Diagrams of BS En Steels, Special<br />

Report no. 56, 2nd Ed.; The Iron <strong>and</strong> Steel Institute: London, 1956.<br />

73. Priestner, R.; Hodson, P.D. Ferrite grain coarsening during trans<strong>for</strong>mation of<br />

thermomechanically processed C–Mn–Nb austenite. Mat. Sci. Technol.<br />

1992, 8 (10), 849–854.<br />

74. Sellars, C.M. The physical metallurgy of hot working. In Hot Working <strong>and</strong><br />

Forming Processes, Proceedings of International Conference, University of<br />

Sheffield, July 17–20, 1979; The Metalls Soc: London, 1980, 3–15.<br />

75. Zhao, J. Continuous cooling trans<strong>for</strong>mations in steel. Mat. Sci. Technol.<br />

1992, 8 (11), 997–1003.<br />

76. Licka, S.; Wozniak, J. Mathematical model <strong>for</strong> analyzing the technological<br />

conditions of hot rolling of steel. Hutnicke Aktuality 1981, 22 (9), 1–49.<br />

77. Dutta, B.; Sellars, C.M. Effect of composition <strong>and</strong> process variables on<br />

Nb(C,N) precipitation in niobium microalloyed austenite. Mat. Sci. Technol.<br />

1987, 3, 197–206.<br />

78. Adrion, H. Thermodynamic model <strong>for</strong> precipitation of carbonitrides in high<br />

strength low alloy steels containing up to three microalloying elements with or<br />

without additions of aluminum. Mat. Sci. Technol. 1992, 8, 406–420.<br />

Copyright 2004 by Marcel Dekker, Inc. All Rights Reserved.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!