Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Acta Materialia 50 (2002) 1383–1392

www.actamat-journals.com

Prediction of chromium depleted-zone evolution during


aging of Ni–Cr–Fe alloys
H. Sahlaoui a,*, H. Sidhom a, J. Philibert b
a
Laboratoire de Mécanique, Matériaux et Procédés, Ecole Supérieure des Sciences et Techniques de Tunis, 5, Av. Taha Hussein
1008 Montfleury, Tunisia
b
LPCES, Université Paris-Sud, F-91405 Orsay Cedex, France

Received 16 October 2001; received in revised form 26 November 2001; accepted 5 December 2001

Abstract

The phenomenological and analytical study of depleted chromium zone in Ni–Cr–Fe alloys has been carried out in
order to predict the evolution of chromium profiles resulting from carbide precipitation during aging. A diffusional
model has been developed to take into account the two stages of chromium concentration evolution: dechromization
and rechromization. The model has been verified using the available experimental data found in the literature for a
nickel alloy type: Inconel 690. The changes of the chromium concentration at the carbide–matrix interface xiCr(t) and
near the grain boundary xCr(x,t), as a function of annealing conditions, can be assessed with a reasonable accuracy by
the proposed model and compared to the previous thermodynamic and kinetic models.  2002 Acta Materialia Inc.
Published by Elsevier Science Ltd. All rights reserved.

Keywords: Carbides; Nickel alloys; Aging; Depleted zone

1. Introduction concentration in various commercial nickel based


alloys in the as received condition ranges on the
Chromium confers to steels and special alloys in average from 15 to 30 wt%. These alloys are
general and to nickel alloys in particular a good homogenized by a standard processing or by high
resistance to intergranular attack IGA and to inter- temperature annealing [8]. The high chromium
granular stress corrosion cracking IGSCC [1–22]. concentration destabilizes the solution by generat-
This property requires a minimum chromium con- ing complex chromium carbides of type Cr3C2,
tent higher than (or equal to) a critical value xcrit
Cr , Cr7C3, Cr23C6 and intermetallic phases which are
which is a function of the chemical composition of rich in chromium like σ, χ and R phases [1–22].
these alloys [1–8]. For this reason, the chromium These compounds precipitate more or less quickly
between 400 and 900°C according to the content
of carbon, chromium and molybdenum of ternary
* Corresponding author. Tel.: +216-1-49-60-66; fax. +216-
1-39-11-66. alloys Ni–Cr–Fe [1–9]. They often respect a
E-mail address: Habib.Sahlaoui@ipein.rnu.tn (H. sequence of heterogeneous nucleation, beginning
Sahlaoui). with intergranular precipitation, then precipitation

1359-6454/02/$22.00  2002 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
PII: S 1 3 5 9 - 6 4 5 4 ( 0 1 ) 0 0 4 4 4 - X
1384 H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392

on incoherent twin boundaries and on the dislo- account the experimental profiles available in the
cations [2,10]. Their growth, which is controlled literature [2].
by an intragranular thermally activated diffusional
process, depends on the chromium content in the
adjacent precipitates and on the matrix carbon con- 2. Phenomenological analysis of depleted zone
tent [10–21]. The consequence of the intergranular
precipitation of these phases in general, and of the It is well established that the time t0 to start
carbides in particular, is to create a chromium- nucleation of chromium carbides in alloys type Ni–
depleted zone near the grain boundaries. These Cr–Fe and Fe–Cr–Ni depends on exposure tem-
depleted zones get wider as the temperature and/or perature, chemical composition and particularly on
the aging duration is higher [1–9] and reduce the carbon content and alloy microstructure [8]. Chro-
local resistance to intergranular corrosion IGC. mium depleted zones were created following the
Furthermore, they increase the susceptibility to carbides precipitation at the grain boundary
IGSCC in a chlorinated medium [1–8]. For these because bulk diffusivity of chromium is relatively
reasons many studies have been devoted to a quan- low compared to that of carbon. Many studies [1–
titative evaluation of the dechromized zones by 22] report that the chromium content at the car-
experimental analysis or by empirical or analytical bide–matrix interface or grain boundaries
models [1–22]. These models attempt to find by decreases significantly with nucleation and growth
calculation the chromium concentration profiles of the intergranular chromium carbides. The
and their evolutions during aging. These results are growth of carbides during annealing will consume
of great interest for the evaluation of the stages of more chromium atoms which will diffuse from the
sensitization and desensitization to the IGC during interior of grains and leads to a thicker depleted
aging, under the accelerated operating conditions zone which is described by a profile xCr(x,t). This
of the nickel alloys used in industrial equipment profile is characterized by the chromium concen-
tration at the carbide–matrix interface xiCr(t) and by
[1–7]. However, the profiles obtained by the ther-
the width of the depleted zone l which depend
modynamic and kinetic models reveal significant
maily on the annealing conditions of these alloys.
discrepancies when they are compared to the
The evaluation of these characteristics requires pri-
increasingly precise experimental determinations
mary a rigorous analysis of the involved physical
by recent analysis techniques (STEM). These dif-
phenomena.
ferences become more significant as the aging per-
iod is longer [2], and can have risks on the time– 2.1. Evolution of chromium concentration at the
temperature–sensitization (TTS) diagrams predic- grain boundary
tion and consequently on the protection of equip-
ment against localized corrosion. Simplifying The chromium concentration at the carbide–
assumptions seem to be the origin of these discrep- matrix interface or at the grain boundary xiCr(t)
ancies, particularly the hypothesis that the carbides reaches a critical value xcrit
Cr after annealing duration
nucleation is instantaneous at any exposure tem- ts at temperature T. This critical value corresponds
perature and that the chromium content at the car- to the beginning of sensitization at the IGA [1–8].
bide–matrix interface reaches instantaneously its The concentration xiCr(t) decreases continuously to
thermodynamically equilibrium value [1,2,6]. For reach a minimal value xminCr after annealing duration
that reason, we propose to reformulate a model t corresponding to a partial thermodynamic equi-
able to accurately predict the chromium content librium between carbides and matrix (Fig. 1). Since
evolution of the depleted zones during aging of the the diffusion of carbon atoms is very fast, com-
Ni–Cr–Fe alloys. The proposed model will be pared to that of chromium, the concentration of
based on phenomenological and analytical descrip- carbon decreases uniformly with depth [1].
tions of the dechromization process. The model According to the law of the solubility, the decrease
variables will be well identified to take into of carbon activity leads to an increase in local
H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392 1385

the interface carbide–matrix content xiCr as a func-


tion of exposure duration.

3. Analytical description of the depleted zone

3.1. Chromium concentration evolution at the


grain boundary

Fig. 1. Influence of chromium concentration at the carbide– Carbides nucleation and their first growth stages
matrix interface on the process of sensitization–desensitization (stages b and c—Fig. 2) involve atom migration
at the IGA during aging. from the matrix towards the carbides which are
constituted of all metal elements available in the
matrix like carbide formers (chromium and
chromium activity in equilibrium with carbides molybdenum) as well as other elements like nickel,
that constitute the driving force for chromium bulk silicon and manganese [3,8]. The contribution of
diffusion from the interior of the grain towards the these elements to the carbide precipitation modifies
depleted zones [1]. The last phenomenon involves significantly the interface chemical composition
a rise in the chromium concentration at the and particularly its chromium content which is the
depleted zones interfaces which leads to the major element in the carbides. The change of inter-
rehomogenization of alloys. This process of face chromium content can be described, during
rechromization acts against the dechromization stages of germination and growth, by the fre-
process related to the growth of carbide which is quently used relationship (1):
delayed by the reduction of carbon concentration
in the matrix. The desensitization for IGA (self- t
healing) can be obtained when the chromium con- xiCr(t) ⫽ x0Crexp(⫺k ) (1)
t
centration at the carbide–matrix interface reaches
the value xcrit where xiCr(t) can be considered as the solution of
Cr after one duration exposure td (Fig.
1). chemical kinetics equation of a first order interfa-
dxiCr(t)
cial reaction ⫽ ⫺kxiCr(t), with a constant k
2.2. Evolution of the chromium concentration at dt
the depleted zone typical of interfacial reaction.
The rehomogenization stage (stage d—Fig. 2)
High chromium content xcCr of carbides which which is controlled by the chromium bulk diffusion
ranges from 60 to 90 at.% [1–9], is obtained at modifies the chromium content at the carbide–
the expense of the adjacent matrix initial chromium matrix interface and at the grain boundaries. The
content x0Cr which depends on the alloy type. A change of the chromium concentration can be
chromium consumption by carbide growth is bal- described analytically by considering the conser-
anced by the bulk diffusion of this element that vation of matter flow through these interfaces and
leads to the increase of the concentration gradient. assumes an instantaneously exchange of chromium
This phenomenon can be described by chromium atoms between the matrix and the carbide through
concentration profile evolution in the direction per- the interface. This leads to the equality between
pendicular to the grain boundaries as a function of the two areas S1 and S2 (stages c and d—Fig. 2).
annealing time (Fig. 2). Matrix chromium content • The area S1 is given by Eq. (2):
in the depleted zones can be predicted by solution
S1 ⫽ xcCrr (2)
of Fick’s second law. Through the width of the
depleted zone (l), chromium concentration varies Since the growth of carbides is controlled by the
between the initial alloy chromium content x0Cr and chromium diffusion in the grains, their sizes r can
1386 H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392

Fig. 2. Chromium profile concentration evolution during aging.

be expressed according to Zener’s empirical rt ⫽ b冑Dt (5)


relation [5], by Eq. (3):
Combining relations (2) and (5), S1 can be
r ⫽ b冑Dt (3) expressed by:
where b is a constant determined by Zener method S1 ⫽ xcCrb冑Dt (6)
[5] and D is the chromium diffusion coefficient,
which depends on the temperature as a thermally • The area (S2) can be approximated by a tri-
activated process (4): angular shape where the chromium concentration

冉 冊
near the grain boundaries is assumed to have a lin-
⫺Q
D ⫽ D0exp (4) ear profile as:
RT
S2 ⫽ (x0Cr⫺xiCr)l / 2 (7)
D0 is equal to 0.08 cm2/s and Q is equal to 58,500
cal/mol according to [1,4]. the exact value of area S2 can be determined by a
During the annealing period of t, the carbon solution of Fick’s second law equation according
activity in the matrix decreases significantly and to relation (8):
the interface chromium content reaches a minimal


L0
value xmin
Cr , then the growth of carbides is suf- x
S2 ⫽ (x0Cr⫺xiCr) (1⫺erf(
2冑Dt
ficiently slowed down to be neglected and its size )) dx (8)
rt can be considered constant and expressed by: 0
H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392 1387

L0 is a constant higher than or equal to the width


冪t
t
of the depleted zone and could tend to infinity. In xiCr(t) ⫽ x0Cr⫺xcCrb (13)
this study L0 is approximated to half of a grain size.
The finite difference method can be used to cal- It is important to specify that evolution of the inter-
culate accurately the area S2 according to the fol- face chromium concentration, during isothermal
lowing expression (9): exposure, is described by the relation (1) during

冘冋 册
the depletion stage and by the relation (13) during
(xCr(xj) ⫹ xCr(xj+1)
S2 ⫽ x0Cr⫺ ⌬x (9) the self healing (desensitization) stage. Moreover,
2 the size (rt) is not the real carbide final size but
where xCr(xj) ⫽ xiCr(t) ⫹ (x0Cr⫺ its size when the chromium concentration at the
xj interface reaches its minimal value indicating the
xiCr(t)).erf( ),xCr(xj ⫽ xiCr(t) ⫹ (x0Cr⫺ beginning of self healing stage. In this last stage,
2冑Dt it is possible to temporarily neglect the carbide
xj ⫹ 1 growth (and not to infinite) because other phenom-
xiCr(t)).erf( ), ⌬x ⫽ xj ⫹ 1⫺xj. xj values can vary
2冑Dt ena of precipitation can occur such as:
between 0 and L0. The chromium concentration at
the carbide–matrix interface can be obtained by 앫 The precipitation of new inter and intragranular
writing the equality between the two areas obtained carbides and intermetallic phases which are
by Eqs. (6) and (7) or for a better accuracy by (6) more stable for long annealing periods.
and (8 or 9) which gives: 앫 The redissolution of chromium carbides in favor
of more stable other phases such as the σ-
xcCrb冑Dt phase, etc.
xiCr(t) ⫽ x0Cr⫺ L ⬵x0Cr (10)


0
x This assumption could be accepted to determine
(1⫺erf(
2冑Dt
)) dx
the sensitization time (ts) and desensitization time
0
2xcCrb冑Dt
(td) which are principally related to the chromium
⫺ carbide precipitation.
l
the value of integral is calculated by the following 3.2. Evolution of the chromium concentration in
expression (11): the depleted zones

冕 冋
L0

x The evolution of the chromium concentration


(1⫺erf( )) dx ⫽ x
2冑Dt
(11) near the grain boundaries can be obtained by the
numerical or analytical resolution of Fick’s second

冉冑 冊册
0

⫺2冑Dt
x x 1 x2 L0 law for the unidirectional diffusion with the fol-
) ⫹ exp(⫺ )
2 Dt 2冑Dt
erf(
冑p 4Dt 0
lowing initial and limit conditions:
tⱕt0⇒xCr(x,0) ⫽ x0Cr∀x
The relation (10) can be simplified by substituting
the width (l) by its approximate expression: x ⫽ 0⇒xCr(0,t) ⫽ xiCr(t)
l⬇2冑Dt (12) xⱖl⇒xCr(x,t) ⫽ x0Cr
This equality is obtained from the random-walk The chromium concentration xiCr(t) at the car-
theory according to Stawström et al. [1] and Bor- bide–matrix interface is determined, according to
ello et al. [4]. The substitution of l in Eq. (10) model described in the previous paragraph
allows us to write: (Section 3.1).
1388 H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392

3.2.1. Carbide precipitation stage 1 x


xCr(x,t) ⫽ (xiCr(t)⫺x0Cr).erfc( ) ⫹ x0Cr
2冑Dt
During the carbide precipitation (nucleation and (15)
2
growth, stages b and c—Fig. 2), the chromium
concentration profile is computed assuming dif- In this equation, the origin (x ⫽ 0) is always
fusion in a semi-infinite medium, in accordance to taken on the plane p, which intercepts the diffusion
the original work of Stawström and Hillert [1] fol- direction at coordinate l / 2 (with l ⫽ 2冑Dt) and x
lowed by other authors [2–6]. This assumption is ⫺l l
justified by chromium exchange between the car- varies in the interval [ , ].
2 2
bides and the matrix through their common inter-
face during precipitation stages. Consequently, the 3.3. Identification of model constants
chromium concentration profile can be given with
a good approximation by Eq. (14), which is derived Constant t can be given by an empirical relation
from the exact solution of the Fick’s second law: (16) derived from Eq. (5).

冉冊
x
xCr(x,t) ⫽ xiCr(t) ⫹ (x0Cr⫺xiCr(t)).erf( 1 rt 2

2冑Dt
) (14)
t⫽ (16)
D b

3.2.2. Self healing stage In this relation, rt is the average value of the car-
Eq. (14) cannot be adopted during desensitiz- bides size at the end of precipitation, which can be
ation stage (stage d—Fig. 2), where chromium estimated by micrographic observation. The value
exchange through carbide–matrix interface of t, which corresponds to the required duration to
becomes sufficiently slow to be neglected. Conse- reach the minimal chromium concentration at the
quently, we propose a model based on an interdif- carbide–matrix interface xmin Cr can be assessed
fusion mechanism through a plane p so-called Mat- directly from experimental measurements of the
ano plane. This plane divides the depleted zone chromium concentration evolution as a function of
into two equal areas A1 and A2 (Fig. 3). The trans- annealing. Since diffusion phenomenon is a ther-
port of the chromium atoms, through this plane p, mally activated process, the relation time–tempera-
from area A2 towards area A1 will take place with- ture equivalence of Arrhenius type (17), can be
out accumulation. This diffusion leads to an evol- used to derive the value of t for any temperature
ution of chromium concentration as a function of exposure.
healing time xCr(x,t) described by the approximate
Eq. (15). The last equation is derived from the
Fick’s second law exact solution, which supposes
t2 ⫽ t1exp 冋 Q 1 1
( ⫺ )
R T2 T1 册 (17)

a constant interface chromium concentration: Constant k is given by relation (18) and determined

Fig. 3. Chromium profile evolution as a function of exposure condition during desensitization.


H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392 1389

by equating 1 and (13), which expresses the conti- 5. Discussion


nuity condition between the two dechromization
and rechromization stages for t ⫽ t The major contribution of the present model lies
the high accuracy prediction of the interface car-
xcCrb bide–matrix chromium concentration and its evol-
k ⫽ ⫺Ln(1⫺ 0 ) (18)
xCr ution xiCr(t) with annealing time. The value of this
concentration is required for a better prediction of
which is a consequence of intergranular precipi-
tation of chromium carbide in the Ni–Cr–Fe alloys.
4. Model validation The accurate determination of the dechromization
profiles is of particular interest when these alloys
which are usually used in electronuclear, aerospace
The prediction of chromium concentration evol- and agroalimentary applications, are exposed to
ution at the carbide–matrix interface xiCr(t) and near operating conditions which induce dechromization.
the grain boundaries xCr(x,t) during aging of Ni– This last phenomenon reduces the resistance of Ni–
Cr–Fe alloys requires the determination of the fol- Cr–Fe alloys to intergranular corrosion IGC and to
lowing parameters: intergranular stress corrosion cracking IGSCC [1–
22]. The proposed model gives a rather precise
앫 t0: time required for carbide nucleation which is chromium concentration at the carbide–matrix
evaluated from time–temperature–precipitation interface, which was lacking in many empirical,
diagrams; thermodynamic and kinetic models proposed in the
앫 t: determined by application of the Eq. (16) literature [1–7]. Moreover, the assumption of semi-
and/or (17) depending on the available data; infinite diffusion hypothesis, used by these models
앫 k: determined by Eq. (18). does not seem to be well adapted to account for
the rechromization phenomenon at the end of the
carbide precipitation, leading to the significant dif-
To compute chromium concentration profile during ferences between the calculated and measured pro-
isothermal exposure at a temperature T, we pro- files [1,2,6]. The present model gives a better pre-
ceed, depending on the aging time (t), in the fol- diction of chromium profiles during aging, as
lowing manner: shown by the superposition of calculated and mea-
sured chromium profiles for Inconel 690 (Figs. 4–
앫 t ⬍ t: xiCr(t) is determined by Eq. (1) and the 6). The chromium profiles resulting from the
profile xCr(x,t) is computed using Eq. (14); annealing at 700°C of Inconel 690 measured by
앫 t ⬎ t: xiCr(t) is determined by Eq. (13) and the Kai et al. [2] and calculated by the present model
profile xCr(x,t) is computed using Eq. (15). are in good agreement. These profiles diverge as a
function of time from those obtained by the
thermodynamic/kinetic model developed by Was
This established model was validated using data and Kruger [6] and modified by Kai et al. [2] (Fig.
for Inconel 690 which is one of the most sensitive 6). These discrepancies which are more significant
to the phenomenon of chromium carbide precipi- when aging duration is longer, were attributed by
tation at the grain boundaries and the subsequent the authors to the intragranular precipitation of car-
dechromization during aging at their operating bides and to the thermodynamic instantaneous
temperature range (vapor generator, exchanger, equilibrium at the grain boundary hypothesis
turbojet, etc.). Chromium concentration at the car- which is not very founded and in contradictory to
bide–matrix interface and near the grain bound- the carbides nucleation and growth kinetics. This is
aries during aging of this alloy evaluated by the the reason for which all the thermodynamic/kinetic
proposed model is in good agreement with the model profiles find a carbides–matrix interface
experimental data (Figs. 4–6). chromium concentration at the carbide–matrix
1390 H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392

Fig. 4. Chromium carbides–matrix interface concentration evolution during aging of Inconel 690 at T=600°C (t=50, k=0.57) and at
T=700°C (t=6, k=0.40).

Fig. 5. Comparison between measured data [2] and present model calculated chromium depletion profiles during aging at 24 h, 48
h, 100 h and 215 h of Inconel 690 at 600°C.

interface almost constant (苲20 wt%) for all the with the experimental results. Moreover, the
aging duration (Fig. 6). The proposed model, hypothesis considered in this study which consists
which takes into account the evolution of the chro- of neglecting of the carbide–matrix interface dif-
mium concentration at the carbide–matrix interface fusion after a duration t and of using an interdif-
with annealing time is in reasonable agreement fusion solution is justified by the vanishing value
H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392 1391

Fig. 6. Comparison of chromium depletion profiles: measured [2], calculated with former model [2] and with the present model
during aging at 5 h, 10 h, 48 h and 100 h of Inconel 690 at 700°C.

of the horizontal tangent to the profiles at the end the Ni–Cr–Fe alloys can be accurately predicted by
∂C(x,t) a two stage model, i.e. dechromization and rechro-
of the precipitation (( ) ⫽ 0). The last
∂x x ⫽ 0 mization. The model constants were identified and
result was not obtained by other models, which are validated for Inconel 690 heat treated for different
based on the assumptions of semi-infinite dif- periods of time (5 h, 10 h, 24 h, 48 h, 100 h, 215
fusion, although it was observed on experimentally h) and at two temperatures (600°C and 700°C).
measured profiles published in the literature [1–7]. An analytical calculation using approximate sol-
utions of diffusion equations can correctly describe
the phenomena of dechromization and rechromiz-
6. Conclusion ation without the need to resort to a heavy numeri-
cal calculation method (finite differences) to solve
Chromium concentration evolution at the car- the complete equations system describing these
bide–matrix interface and near the grain bound- phenomena. This last method would require in any
aries (chromium depleted zones) resulting from case several approximations, since the final equi-
intergranular carbides precipitation during aging of librium state of the alloy cannot be simply
1392 H. Sahlaoui et al. / Acta Materialia 50 (2002) 1383–1392

described because of the multiplicity of phases M23C6 and austenitic stainless steel. Acta Metall
which can be appear, according to alloy chemical 1974;22:89–94.
[11] Cihal V, Kasova I. Relation between carbide precipitation
composition, exposure conditions, etc. and intercrystalline corrosion of stainless steel. Corrosion
Science 1970;10:875–81.
[12] Slattery GF et al. Metallographic techniques for the exam-
References ination of grain boundary precipitate in type 316 stainless
steel. Pract Metall 1981;18:292–303.
[1] Stawström C, Hillert M. Improved depleted-zone theory [13] Bagnall C, Shiels SA. Equilibrium sructures of type 304
of intergranular corrosion of 18-8 stainless steel. J Iron and 316 stainless steel in the temperature range 800–
Steel Inst 1969;(Jan.):77–85. 1500F: a metallographic review. Microstruct Sci
[2] Kai JJ, Yu G, Tsai CH, Liu MN, Yao SC. The effects 1976;5:53–64.
of heat treatment on the chromium depletion, precipitate [14] Beckitt FR, Clark BR. The shape and mechanism of for-
evolution, and corrosion resistance of Inconel alloy 690. mation of M23C6 carbide in austenite. Acta Metall
Metall Trans A 1989;20A:2057–67. 1967;15:113–29.
[3] Was GS, Tischner HH, Latanision RM. The influence of [15] Singhal LK, Maetin JW. The nucleation and growth of
thermal treatment on the chemistry and structure of grain widmannstätten M23C6 precipitation in an austenitic stain-
boundaries in Inconel 600. Metall Trans A less steel. Acta Metall 1968;16:1159–65.
1981;12A:1397–408. [16] Dacasa C, Nileshwar VB, Melford DA. M23C6 precipi-
[4] Borello A, Casadio S, Saltelli A, Scibono G. Susceptibility tation in unstabilized austenitic stainless steel. J Iron Steel
to the intergranular corrosion of alloy 800. Corrosion-nace Inst 1969;1325–31.
1981;37(9):498–505. [17] Lewis MH, Hattersley B. Precipitation of M23C6 in austin-
[5] Mayo WE. Predicting IGSCC/IGA susceptibility of Ni– itic steels. Acta Metall. 1965;13:1159.
Cr–Fe alloys by modeling of grain boundary chromium [18] Adamson JP, Martin JW. A microanalytical study of
depletion. Mat Sci Eng 1997;A232:129–39. M23C6 carbides as function of aging time in Nb-bearing
[6] Was GS, Kruger RM. A thermodynamic and kinetic basis austenitic steels. Metallography 1974;7:397–402.
for understanding chromium depletion in Ni–Cr–Fe sys- [19] Singhal LK, Martin JW. The growth of M23C6 carbide on
tem. Acta Metall 1985;33(5):841–54. incoherent twin boundaries in austenite. Acta Metall
[7] Van Duysen JC, Guttmann M. A STEM-EDS study of 1967;15:1603–10.
the sensitization to intergranular corrosion of alloy 800. [20] Hillert M, Lagneborg R. Discontinuous precipitation of
Colloque Iternational de Mons. M23C6 in austenitic steels. J Mat Sci 1971;6:208–12.
[8] Boger HE, Gall TL. Metals handbook. American Society [21] Million B et al. Carbon diffusion and thermodynamic
for Metals, 1985. characteristics in chromium steels. Z Metallkd
[9] Was GS, Rajan VB. The mechanism of intergranular 1995;86(10):706–12.
cracking of Ni–Cr–Fe alloys in sodium tetrathionate. Met- [22] Thorwaldsson T, Dunlop GL. The influence of compo-
all Trans A 1987;18A:1313–23. sition on precipitation in stabilized austenitic stainless
[10] Pumphrey PH, Edington JW. The structure of the sem- steels. The Vth international conference strength of metal
icoherent interface between grain boundary nucleated and alloys, vol. 1, 1979.

You might also like