Sei sulla pagina 1di 49

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/314375943

AUSTENITE INSTABILITY AND PRECIPITATION BEHAVIOR OF HIGH


NITROGEN STAINLESS STEELSi

Chapter · January 2016

CITATIONS READS

0 589

3 authors:

Izabel Fernanda Machado Patricia Almeida Carvalho


University of São Paulo SINTEF, Oslo, Norway
78 PUBLICATIONS   352 CITATIONS    241 PUBLICATIONS   1,085 CITATIONS   

SEE PROFILE SEE PROFILE

Angelo Fernando Padilha


University of São Paulo
167 PUBLICATIONS   2,076 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

DiFusion View project

Planar defects in ordered systems View project

All content following this page was uploaded by Patricia Almeida Carvalho on 09 March 2017.

The user has requested enhancement of the downloaded file.


in Stainless Steel: Microstructure, Mechanical Properties and Methods of
Application (Mechanical Engineeting Theory and Applications)
by Alokesh Pramanik (Editor), Animesh Kumar Basak (Editor)
Nova Science Publishers
ISBN-10: 1634820800

AUSTENITE INSTABILITY AND PRECIPITATION BEHAVIOR OF


HIGH NITROGEN STAINLESS STEELS

Izabel Fernanda Machado1), Patrícia Almeida Carvalho2) and, Angelo


Fernando Padilha3) Text

1) Mechatronics and Mechanical System Engineering Department,


Polytechnic School. University of São Paulo. S. Paulo, Brazil. E-mail:
machadoi@usp.br
2) SINTEF Materials and Nanotechnology, Oslo, Norway and Department
of Chemical Engineering, Instituto Superior Tecnico, University of Lisbon,
Lisbon, Portugal. E-mail: patriciaalmeida.carvalho@sintef.no
3) Metallurgical and Materials Engineering Department, Polytechnic
School. University of São Paulo. S. Paulo, Brazil. E-mail: padilha@usp.br

ABSTRACT

Nitrogen is a strong austenite stabilizer and improves the mechanical


properties and the corrosion resistance of austenitic stainless steels. For these
reasons, nitrogen addition to steel has been extensively studied in the last 40 years.
This chapter presents a literature review since 1926, when a pioneer study on the
effects of nitrogen in iron-based alloys was reported. The maximum solubility of
nitrogen in stainless steels is high, although it decreases considerably at
temperatures below 1000ºC. Therefore, depending on the composition, high
nitrogen austenitic stainless steels can undergo different phase transformations
during exposure to temperatures ranging between 500 and 1000oC. Continuous
and discontinuous precipitations of chromium nitride, as well as ferrite and sigma
phase formation, have been observed and, as a result of these phase
transformations, a loss of toughness and lower corrosion resistance are frequently
detected. The aim of this chapter is to present and discuss the austenite stability
and illustrate the precipitation behavior of high nitrogen austenitic stainless steels.
Three different high nitrogen stainless steels are used in order to highlight
stabilization/destabilization effects and microstructural features associated with
the presence/depletion of nitrogen. The nucleation of discontinuous precipitation
of chromium nitride and its growth kinetics are discussed and transformation
models are presented.

Keywords: high nitrogen steels, austenite stability, chromium nitrides


precipitation, ferrite and sigma phase format
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

INTRODUCTION
Austenitic stainless steels present an excellent combination of corrosion
resistance, ductility, toughness and weldability, and account for about 70% of the
world’s  stainless  steel  production.  The  almost  centenary  AISI  304  (German  V2A)  
and AISI 316 are still the most consumed types of stainless steel [1,2]. However,
the strength level of these common types, particularly the yield strength (YS) in
the annealed condition, is relatively low, around 200 MPa [1, 3 ]. Another
important type of stainless steels is the duplex (ferritic-austenitic microstructure),
which is employed where higher yield strength (YS), better pitting corrosion and,
intergranular corrosion resistance are necessary [4]. However, duplex stainless
steels are extremely susceptible to alpha prime (α’) formation at temperatures
ranging from 300 to 500oC and to sigma phase formation between 550 and 950oC.
These phases are highly deleterious to the toughness and corrosion resistance of
the duplex steels and their formation must be avoided during heat treatment,
welding or application [4,5].
High levels of nitrogen addition stabilize austenite and can increase
significantly its YS as well as enhance other properties [ 6 -8 ]. According to
Speidel, 0.4 wt.% is a high nitrogen content for austenitic stainless steels [9],
although the levels for high nitrogen content remain debatable [8]. A pioneer
investigation on the effect of nitrogen addition to iron-chromium alloys was
reported in 1926 by Adcock [10]. A few years later, it was verified that nitrogen
was a very strong austenite stabilizer since its effect was about 20 times stronger
than that caused by nickel, as showed in Tofaute’s   work   [11]. As a result, the
effect of nitrogen was included in a modified Schaeffler diagram for welding
processes [ 12] by DeLong [13] and Espy [ 14]. Nitrogen addition to duplex
(ferritic-austenitic) steels causes also an enlargement of the austenitic field,
making possible their full austenizatization [ 15 - 22 ]. Additionally, nitrogen
improves fatigue and creep resistance [ 23 - 30 ] and inhibits strain-induced
martensite formation [25, 31-33].
The study of the effects of nitrogen in stainless steels showed a significant
enhancement during the decade of 1980 and was followed by the development of

1
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

new or modified processes [8, 15, 21, 34-37]. Nitrogen addition can be carried out
both in the liquid and solid state and overviews of nitrogen addition technologies
have been presented by Simmons [34] and, more recently, by Lo [8]. Therefore,
nickel replacement by nitrogen, and in some instances by manganese, became an
interesting alternative to retain or improve the mechanical properties and
corrosion resistance of stainless steels [6,22,32,33, 38-43].
The benefits of nitrogen in stainless steels occur mainly when this
interstitial element is in solid solution [6, 8, 25, 34, 44-53] as first observed by
Krivobok [44]. The precipitation behavior of chromium nitride in high nitrogen
stainless steels presents peculiarities that make its evaluation important from the
scientific and technological point of views [15,22,34,52,-56]. In addition, other
phase transformations can occur after chromium nitride precipitation, such as
ferrite and sigma phase formation in the nitrogen-depleted regions
[16,17,20,21,39,53]. The aim of this chapter is to address austenite stability and
the precipitation of chromium nitride in high nitrogen stainless steels. For the sake
of clarity the discussion about nitrogen effects will be illustrated for three
different stainless steels, and will focus on the different phase transformation
mechanisms, namely on the nucleation and growth of chromium nitride, on the
precipitation kinetics, and on the consequences of the precipitation for the matrix
stability.

HIGH NITROGEN STEELS DESCRIPTION AND


CHARACTERIZATION

The austenite stability and the precipitation behavior presented in this


chapter will be discussed for three different high nitrogen stainless steels, namely,
(i) a completely austenitic steel, (ii) a functionally graded steel with an austenitic
nitrogen-rich outer region enclosing a duplex (ferritic-austenitic) core and (iii) a
nickel-free high-manganese duplex (ferritic-austenitic) stainless steel with a
casting microstructure:

2
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

(i) Standard type duplex stainless steel DIN W.-Nr.1.4460 in which


nitrogen was added by pressurized electroslag remelting [15]. The resulting
fully austenitic steel was received in the form of a forged bar with the
following composition (wt.%): 25.1Cr, 1.9Mo, 5.5Ni, 1.5Si, 1.20Mn, 0.03C
and 0.87N (designated as AN steel).
(ii) Standard type duplex stainless steel DIN W.-Nr.1.4462 to
which nitrogen was added by solid solution nitriding [18,21]. The result was
an austenitic stainless steel case near the surface and a duplex steel in the
inner region. The sample was a plate with a thickness of 2.5 mm and the
following composition (wt.%): 22.04Cr, 2.9Mo, 5.5Ni, 0.47Si, 1.70Mn,
0.025C, with the nitrogen varying in austenite from about 0.7 on the surface
of the sample to 0.4 wt.% in the inner part [18,21] (designated as ADN
steel).
(iii) Nickel-free duplex stainless steel (ferritic-austenitic
microstructure) to which 0.54 wt.% nitrogen was added [38,39]. The
resulting steel was received in as cast condition with the following
composition (wt.%): 25.8Cr, 0.11Ni, 1.12Si, 17.2 Mn, 0.035C and, 0.54N
(designated as DN steel).

The microstructural characterization was carried out by several


complementary techniques, such as optical microscopy, scanning electron
microscopy (SEM), transmission electron microscopy (TEM), energy-dispersive
spectroscopy (EDS), wavelength-dispersive spectroscopy (WDS) and X-ray
diffraction (XRD), as well as Vickers hardness and magnetic measurements. The
metallographic sample preparation was described in previous works [20,21, 39],
however, it is worth reporting the composition of V2A-Beize, a valuable chemical
etchant for stainless steels, which was used in the metallographic preparation of
the investigated steels prior to optical and scanning electron microscopy
observations: 100 ml of HCl, 100 ml of distilled water, 100 ml of HNO3 and 0.3
ml of Vogel's Sparbeize. Crystal structures and lattice parameters of the
predominant phases have been determined by XRD using  CuKα1 radiation. The

3
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

parameters employed in the hardness measurements can be found elsewhere


[20,21, 39]. The amount of ferromagnetic phases was determined using the
magnetic induction method (Fisher Permascope® M11D with a 0.1% ferrite
detection limit).

SOLUTION ANNEALING AND AGING HEAT TREATMENTS

Austenite or ferrite-austenite (duplex) solid solutions (free of precipitates)


were obtained for each steel using specific heat treatments. Solution annealing of
the AN and DN steels was carried out, respectively, at 1250oC for 1 hour and at
1050oC for 2 hours, whereas a solid solution nitriding process was used for the
ADN steel. This treatment involved heating for 5 hours at 1200oC in N2 partial
pressures of 2 bar in order to obtain a nitrogen-rich layer at the surface. The
samples were water quenched after annealing for nitrogen to remain in solid
solution. The subsequent aging treatments were carried out at temperatures
ranging from 600 to 1100oC, for times between 1.5 and 6000 minutes, and were
followed by water quenching. Throughout the text austenite will be designated by
A or , ferrite by F or sigma phase by S  or  σ  and duplex regions by D or .

NITROGEN SOLUBILITY

Nitrogen solubility in iron and steels is strongly dependent on the chemical


composition and on the crystal structure(s) of the alloy [ 57 ], in addition to
temperature and pressure [54,58]. As a starting point we will summarize the
known solubility behavior for the Fe-Cr-Mn and Fe-Cr-Ni systems, which to a
first approximation can be considered to include the AN, ADN and DN steels.
The effect of alloying elements on the nitrogen solubility for the Fe-Cr-Mn system
is shown in Figure 1, whereas Figure 2 shows the influence of crystal structure
for the Fe-Cr-Ni system.

4
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 1. Influence of alloying elements and temperature on Cr 2N precipitation in the


Fe-Cr-Mn-N system. Adapted from [54].

Figure 2. Influence of crystal structure on the nitrogen solubility for alloys of the Fe-Cr-
Ni system at 0.9579 atm of N2. Adapted from [54].

5
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

The AN and the ADN steels had their nitrogen solubility increased since
the nitriding treatments were carried under pressure. Indeed some technological
processes are carried out at high pressures to promote the introduction of higher
quantities of nitrogen into the steel [15,18,58]. The mass percentage of nitrogen in
the alloy (C0) and its relation with the N2 pressure (PN2) follows Sievert Law [15]:

𝐶 ∝ 𝑃 (1)

which is typically used at pressures of about 1 atm. However, there has been
evidence that this law can also hold for higher pressures, i.e., 𝐶 (𝑃 ≠ 1) =
𝐶 (𝑃 ) [54,59-61] and equation (1) can be considered more general [15]. As a
result of the high nitrogen content in the AN and ADN steels, complete
dissolution of nitrogen was achieved only at high temperatures.
In the AN steel, the chromium nitride precipitation occurs up to 1050oC,
therefore solution annealing treatments have to be carried out at higher
temperatures. On the other hand, the stability domain of ferrite starts at
temperatures as low as 1300ºC, and annealing at higher temperatures results in
duplex microstructures. Annealing at 1250ºC represents the best compromise to
achieve complete dissolution of nitrides and simultaneously guarantee the absence
of ferrite and minimize grain coarsening [16,17,20]. Figure 3 shows the X-ray
diffraction pattern of the AN steel after solution annealing.
Figures 4 and 5 show, respectively, a nitrogen profile across an ADN
sample and the transition microstructure between the austenitic and the duplex
regions formed after the solid solution nitriding process.

6
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 3. X-ray diffraction pattern of AN steel after the solution annealing at 1250oC
for  1  hour  followed  by  water  quenched.  CuKα1  radiation.  

Figure 4. Nitrogen concentration profile of the ADN steel case after the solid solution
nitriding process at 1200oC for 5 hours and at 2 bar N2 pressure. T is the transition region between
the austenitic region and the duplex (ferritic-austenitic) region. X is the distance from the surface.
WDS measurements. Adapted from [17,18,21].

7
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 5. Microstructure of the transition between the austenitic region (nitrogen-rich)


and the duplex region of the ADN steel after the solid solution nitriding process at 1200 oC and 2
bar of N2 for 5 hours. Optical microscopy.

The high levels of chromium and manganese of the DN steel increased the
nitrogen solubility up to 0.5 wt.% [38,39] and the standard solid solution
annealing (1050oC for 2 hours) resulted in a duplex ferritic-austenitic
microstructure free of precipitates as shown by the X-ray diffraction pattern
presented in Figure 6. The volume fraction of ferrite measured by magnetic
induction was 47% (the balance was austenite). After the solid solution treatment
the microhardness of ferrite was HV (0.2 kg) 274±20 and that of austenite was
HV (0.2 kg) 290±16 [39].
The above discussion on solution annealing emphasized the effect of
chemical composition, temperature and pressure on nitrogen solubility. In fact this
is the cornerstone of high nitrogen stainless steels, since the benefits of nitrogen
are directly related to its presence in solid solution [6,8,25,34,52-56]. Therefore
phase transformations must be avoided in manufacturing processes such as
welding, which remains a challenge that demands further research [6,50,52,53].

8
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 6. X-ray diffraction pattern of the DN steel after solution annealing at 1050 oC for 2
hours followed by water  quenched.  CuKα1  radiation.

AUSTENITE STABILIZATION BY SOLID SOLUTION

Nitrogen is a strong austenitizing element as can be seen in Figure 7. Its


influence in terms of nickel equivalent is comparable to that of carbon [13,14].
This effect is clearly observed in the AN e ADN steels, where nitrogen leads to a
change in the equilibrium phase diagram. These steels, which present an initial
ferritic-austenitic microstructure, are fully transformed into austenite after
nitrogen addition. It was demonstrated that the addition of 0.7 wt.% of nitrogen is
enough to produce a fully austenitic microstructure in both cases [17,21,62]. This
result is in agreement with the Espy diagram [14] which can be used on a first
approximation to evaluate the volume fraction of the phases present at room
temperature. In the case of the DN steel the high chromium content and, the
replacement of nickel by manganese does not stabilize completely austenite and a
duplex microstructure is present.

9
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 7. Volume fraction of austenite in the Fe-Cr-Ni system as a function of


temperature and nitrogen content. (a) Corresponds to 0.05, (b) to 0.10 and (c) to 0.20 wt.%
nitrogen. Adapted from [63].

These austenite-stabilizing effects of nitrogen are corroborated by the work


carried out by Presser [ 64 ] and Vandershaeve [ 65 ] on steels with similar
compositions [46,47]. Several studies on the effect of nitrogen in the Fe-Cr-Mn
ternary system [8,32,33,41,43,53] demonstrated also that this element stabilizes
austenite and improves the properties of stainless steels. On the other hand,
Nakada’s  [22] obtained a completely austenitic microstructure in a high chromium
content steel by adding 1 wt.% of nitrogen in the absence of nickel and
manganese.
The diffusion of nitrogen into the ADN steel and the concomitant
austenitization of the nitrided layer can be used to retrieve quantitative parameters
of the system. The boundary conditions for a two phase alloy ( + ) transforming
into a single phase ( ) due to nitrogen enrichment are depicted in Figure 8. As a
result, the thickness (w) of the austenitic layer can be calculated from   Fick’s    
second law [66]:

10
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

/
𝑤 = 4𝐷𝑡 (2)
/

The boundary values can be inferred from the nitrogen profile shown in
Figure 4: the concentration at the surface (Cs) is 0.7 wt.%, the initial
concentration (C0) is 0.14 wt.%. The concentration at the austenite/ferrite interface

(C ) was assumed to correspond to the average concentration at the austenitic-

duplex transition, i.e., 0.35 wt.% [17]. The nitrogen diffusion coefficient at the
annealing temperature (D) is 6.9 10-12 m2/s [67]. Thus for an annealing time (t) of
18000 seconds, w is 643 µm. Since this value is in agreement with the
experimental thickness of the austenitic layer, ~500 µm [17], equation (2) and the

C value can be considered a reasonable approximation.

Figure 8. Schematic of the nitrogen gradient and the associated phase transformation in
the ADN steel (adapted from [66]),   where   γ   designates   austenite,   α   designates   ferrite,   C0 is the

initial concentration in the alloy, C is the concentration at the interface, Cs is the

concentration at the surface, X is the distance from the surface, W is the width of the single phase
region and dW represents the thickness of the layer where the transition from dual to single phase
occurs. Fitting this profile to the experimental one shown in Figure 4 allowed inferring the
boundary  condition  values  necessary  to  solve  Fick’s  Second  Law.

11
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

PHASE TRANSFORMATIONS DURING AGING

The studied steels have high nitrogen content which favors chromium
nitride precipitation at temperatures below the nitrogen solution limit [8,20]. On
the other hand, the high chromium content, which is a ferrite stabilizer and
promotes sigma phase formation [20,21,53], increases the complexity of the
microstructures resulting from the phase transformations. The volume of the
phases present and its evolution during aging was determined by XRD and
magnetic measurements. The time-temperature-transformation diagrams (TTT
diagrams) obtained are shown in Figures 9 to 11 [16,17,20,21,39]. The first phase
formed in austenite is chromium nitride (Cr2N), while sigma phase was the first to
form in ferrite which was almost entirely consumed as the transformation
progressed in the DN steel [17,39].

Figure 9. TTT diagram of AN steel at temperatures ranging from 600 to 1250 oC.
Adapted from [16, 17,20]

12
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

(a)

(b)

(c)

Figure 10. TTT diagram of ADN steel at temperatures ranging from 600 to 1100 oC.
Adapted from [17,21,62].

13
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

(a)

(b)

Figure 11. TTT diagram of DN steel at temperatures ranging from 550 to 1050 oC. (a)
TTT diagram for the ferritic regions and (b) for the austenitic regions. Adapted from [17,39].

Chromium nitride precipitation

Volume or continuous precipitation (CP) involves the nucleation and growth of


individual particles throughout the matrix as the transformation proceeds. The
continuous precipitation is characterized by a continuous gradient of the solute
concentration (and therefore of the lattice parameter) in the matrix surrounding
the precipitate [68,69]. Figure 12 (a) shows a schematic representation of the
composition gradient associated with a growing precipitate. Hornbogen classical
work [70] established the relevance of the crystallographic relation between the

14
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

precipitate and the matrix on the favored sites for continuous precipitation
(vacancies, dislocations, twin and other grain boundaries or homogeneous
nucleation in the matrix). In the presence of a preferred orientation relation the
lattice parameters mismatch results in the generation of strain fields that may be
accommodated by a network of dislocations as illustrated in Figure 12 (b)
adapted from literature [71].

Figure 12. (a) Schematic representation of the continuous change in solute concentration
around a precipitate formed by CP. (b) Illustrative example of Cr2N precipitates formed by a
continuous mechanism in the AN steel aged at 960oC for 2 h. TEM. Adapted from [71]. The stress
field associated with the presence of the precipitates is accommodated by a network of dislocations
in the matrix.

The cellular or discontinuous precipitation (DP) involves nucleation at


grain boundaries followed by a cooperative growth of the precipitates with the
solute depleted matrix, a process which is driven by the grain boundary migration
into an adjacent supersaturated grain as represented schematically in Figure 13. A
composition discontinuity exists between the transformed and untransformed
matrix across the grain boundary [72]. This solid-state transformation produces
two-phase cells or colonies consisting of alternate lamellas of the precipitated
phase and of solute depleted matrix. The accepted initiation mechanisms for DP
are   the   ‘pucker   mechanism’   of   Tu   and   Turnbull   [73,74] and the Fournelle and
Clark mechanism [75].   In   the   ‘pucker   mechanism’   an   embryo,   developing   at   an  
angle with respect to a grain boundary, causes the boundary to be locally deflected
by the torque term associated with the interfacial tension balance at the triple

15
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

point. If the interfacial energy conditions are favorable, the GB tends to migrate
until the higher energy broad face is incorporated in the lower interfacial energy
grain (Figure 14 (a)). The process is iterated to form a group of parallel
precipitates. Although this mechanism has been proven operative, it is critically
dependent on the crystallographic orientation across the GB and on the habit
planes of the precipitates, and it is hence not universal. Nevertheless,
configurations compatible with the Tu and Turnbull model have been found in the
aged AN steel (Figure 14 (b)). In the Fournelle and Clark mechanism,
precipitation occurs on an already moving boundary. The grain boundary is
locally pinned but still allowed to continue migrating between the precipitates
(Figure 14 (b)). In contrast with the previous, this mechanism does not require
any especial interfacial energy conditions, and is probably operating whenever no
preferential matrix/precipitate habit plane can be observed, which is generally the
case (see references [72,76,77]).

Figure 13. (a) Schematic representation of DP adapted from [72],  where  γ0 indicates the
supersaturated  austenite  and  γ  corresponds  to  the  solute  depleted  austenite.  (b)  Illustrative  example  
of a discontinuous precipitation front of Cr2N in the AN steel aged at 1040oC C for 15 min. SEM.

16
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 14. Models for DP initiation. (a) Tu and Turnbull model [73,74]. i refers to the
free energy of the different interfaces between the matrix and the precipitates. (b) and (c)
Configurations of Cr2N precipitation at grain boundaries of the AD steel after 2h at 960°C
compatible with the Tu and Turnbull model. TEM. Adapted from [71]. (c) Fournelle and Clark
model [75].

The growth process associated with DP in alloys containing both


interstitial and substitutional solutes is distinct from the pattern observed in binary
substitutional systems. In these cases the solute transport occurs exclusively by

17
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

interfacial diffusion at the reaction front and therefore the composition of the
untransformed matrix remains constant [72,76,77], which contrasts with the
behavior of the continuous precipitation as depicted in Figure 12. On the other
hand, in interstitial/substitutional systems the key factor governing the
compositional redistribution is the difference between the partitioning processes
of the two types of solutes. In these conditions although diffusion of the
substitutional solute at the reaction front is the rate-controlling effect at the initial
stages of precipitation, long-range volume diffusion of the interstitial solute is the
essential aspect at the later stages [55]: the fast long-range interstitial diffusion
reduces the supersaturation of the untransformed matrix, decreasing the migration
rate of the reaction front and widening the interlamellar spacing. This means that
no steady-state is reached, in contrast with binary substitutional alloys. Therefore,
at later precipitation stages long-range volume diffusion of nitrogen induces
retardation of new nucleation events and inhibition of active reaction fronts that
eventually cease to migrate. This results in incomplete decomposition by
discontinuous precipitation, even though the matrix may still be supersaturated in
nitrogen and the reaction driving force is not equal to zero, which may allow the
continuous precipitation to proceed [55,56,65].
The question of whether continuous or discontinuous precipitation
predominates depends on the mobility and density of discontinuous reaction fronts
as well as on the incubation periods for each mechanism. As the temperature
increases, bulk diffusion becomes faster, favoring volume precipitation [78,79]
and, as a result, close to the solvus temperature continuous precipitation prevails
[80]. Under regimes where discontinuous precipitation is active, any previous or
concomitant volume precipitation reduces the amount of chemical driving energy
available for both initiation and propagation of the discontinuous precipitation.
The Cr2N nitride crystallizes in the hexagonal system and presents an

(111) //(0001)Cr2N and [110] // [1100] Cr2N preferred orientation relation with

austenitic matrices [71]. In high-nitrogen stainless steels the Cr2N nitride can
precipitate both continuously and discontinuously in austenite

18
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[16,17,20,21,34,64,65,81-84], while in duplex steels Cr2N precipitation is favored


at interfaces [85]. The CrN nitride, which crystallizes with the rock-salt
structure, tends to precipitate continuously in ferritic regions according to a Bain
orientation relation [86].
The discussion above explains why the precipitation mechanisms in high
nitrogen stainless steels remain elusive and demand further investigation. The
nucleation and growth of chromium nitrides through the competing mechanisms
are reviewed here for the AN, DN and ADN steels and illustrative examples of the
competing precipitation mechanisms are presented below.
Figure 15 shows initial stages of continuous precipitation of Cr2N
throughout austenite grains in AN and ADN steels (austenitic region). In
particular, Figure 15 (b) demonstrates clearly that the strain fields associated with
dislocations assist the precipitation process. Figure 16 illustrates cases of
precipitation at incoherent and coherent twin boundaries, as well as at grain
boundaries, in AN and ADN steels. Precipitation at the coherent boundaries
occurs later than at the incoherent ones (Figure 16 (a) and (b)), suggesting that
the strain fields at the incoherent twin fronts assist the precipitation process.
Several distinct mechanisms for this behavior have been proposed for carbide
precipitation [87-89]. However, the effect of nitrogen on the stacking fault energy
of austenitic stainless steels is not yet completely understood [90-93]. Stress fields
have been detected in association with precipitates at the grain boundaries (see
arrows in Figure 16 (d)).

19
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 15. Initial stages of continuous precipitation of Cr 2N in austenitic grains. AN


steel aged heat at (a) 700oC for 45 min and (b) and (c) at 800oC for 3 min. TEM.

20
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 16. (a) AN steel aged at 960oC for 2 h. Precipitation of Cr2N at incoherent twin front
(arrows). TEM. (b) AN steel aged at 750oC for 100 h. Precipitation throughout the matrix and at
incoherent and coherent twin boundaries (arrows). SEM. (c) ADN steel aged at 860oC for 10 h.
Precipitation throughout the grains and at twin boundaries. SEM. (d) AN steel aged at 700oC for 45
min. Diffraction contrast near precipitates (arrow) indicates strain fields. TEM. (e) ADN steel aged at
860oC for 10 h. Precipitation at grain boundaries. SEM (f) AN steel aged at 750oC for 10 h for 10 h.
Precipitation at grain boundaries. SEM.

Lamellar colonies resulting from a discontinuous reaction also form in


stainless steels at high nitrogen supersaturations [15,17,20,21,64,65,81,82,84,94].
Due to the similarity in morphology these structures  are  sometimes  called  ‘‘false  

21
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

pearlite’’   [10]. Figure 17 illustrates the morphology of discontinuous


precipitation colonies in the AN steel, where alternate lamellas of Cr2N and solute

depleted austenite (Figure 17 (a)) exhibit the (111) //(0001)Cr2N and [110] //

[1100] Cr2N orientation relation (Figure 17 ((b)). The configuration of the reaction

fronts is shown in Figure 17 (c) and (d) at different magnifications. Figures 18 (a)
to (c) shows the occurrence of well developed cells of discontinuous precipitation
co-existing with grain boundary precipitation (a), precipitation at incoherent (a) to
(c) and coherent (a) twin boundaries. In (a) and (b) the colony produced by DP
surrounds and partly consumes the twin domain while in (c) the lamellar colony
occurs inside, and is confined by, the twin domain. The competition between CP
throughout the austenitic matrix and DP progressing from the austenitic grain
boundaries is illustrated for the AN and ADN steels in Figure 18 (d) and (e).

Figure 17. Discontinuous precipitation in AN steel. (a) Alternating and Cr2N lamellas.
Aged at 960oC for 2h. TEM. (b) Selected area diffraction pattern corresponding to (a). (c)
Discontinuous reaction front. Aged at 960oC for 3 min. SEM. (d) Discontinuous reaction fronts
emerging from a grain boundary. Aged heat treated at 950oC for 3 min. SEM.

22
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 18. (a), (b) ADN steel aged at 900°C for 1 h. DP of Cr2N across a twin domain with

incoherent fronts decorated with precipitates (arrows). SEM. (c) AN steel aged at 900°C for 2 h.
Precipitation colony inside a twin domain with incoherent (arrow) and coherent boundaries
decorated with precipitates. SEM. (d) and (e) DP vs. CP in AN steel aged for 1 h at, respectively,
960°C and 1040°C. SEM. (f) AN steel aged at 900°C for 15 min with decorated grain boundaries
and incoherent (arrows) and coherent twin boundaries in addition to DP vs. CP throughout the
austenitic matrix. SEM.

Figure 19 represents schematically the observed competition between


discontinuous and continuous precipitation in the AN and ADN steels (a) and (b),

23
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

and their relative occurrence in terms of temperature and nitrogen concentration in


the parent phase (c). During aging of austenitic stainless steels the discontinuous
precipitation mechanism tends to prevail at intermediate temperatures
[15,56,65,94, 95 ]. Namely, in the AN steel the discontinuous precipitation is
predominant at temperatures ranging from 850 to 1050oC [16,17,20], while in the
austenitic regions of the ADN steel discontinuous precipitation occurs between
800 and 1000oC [17,21]. Nevertheless, continuous precipitation generally
competes with the discontinuous process even at temperatures for which this
mechanism prevails.
The concentration of nitrogen in duplex stainless steels is typically lower
than that of the DN steel [96] and, as a result of the relatively low nitrogen
content, the Cr2N nitride tends to precipitate in the austenitic regions of duplex
steels by a continuous mechanism [17,39]. In contrast, continuous precipitation of
Cr2N was not detected in the DN steel after aging, although the as-cast
microstructure evidenced colonies resulting from discontinuous precipitation
(Figure 20). Therefore, the transformations that occurred in austenite during
cooling from the melt depleted the solute to the point of inhibiting the continuous
precipitation of Cr2N during the subsequent aging treatments [17,39]. The fact
that grain boundary precipitation was not observed in the as-cast austenite is
noteworthy as this is a required step at the onset of DP. This behavior may be
related with a preferred nucleation of discontinuous precipitation fronts at
subgrain boundaries [97], although this mechanism is still to be established [98].

24
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 19. Schematic representation of (a) discontinuous precipitation of Cr2N and (b)
continuous precipitation competing with the discontinuous process. (c) Continuous and
discontinuous precipitation of Cr2N in austenite as a function of temperature at high and low nitrogen
concentration.

Figure 20. Discontinuous precipitation of Cr2N in the DN steel. (a) Aged at 860oC for
1h. (b) Aged at 850oC for 10h. SEM using backscattered electrons.

25
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

A prior extensive chromium nitride precipitation alters the matrix


composition and induces the formation of ferrite in the austenitic regions [17,20]
(ferrite formation will be discussed later on this chapter). Precipitation of Cr2N
can also occur at the interfaces [17,21,39]. This behavior is justified by the
significant drop in nitrogen solubility below 1100oC and by the favored nucleation
at interfaces [20,21,25,26,71]. Figure 21 (a) shows ferrite plates growing in a
region of the AN steel where nitrides have precipitated continuously. Ferrite
growth involves further nitrogen segregation which in turn contributes to the
formation of nitride precipitates at the interfaces [99]. These precipitates are
able to grow across ferrite plates and into the matrix. EDS analyses and diffraction
experiments identified the precipitates as Cr2N. CrN was not detected in ferrite
contrarily to what is mentioned in the literature [100]. A significant presence of
nitrides was found at prior locations of the interfaces in the ADN steel after
o
aging at 960 C (Figure 21 (b)) in the duplex region. In this case, the Cr2N
precipitation is likely to have changed the equilibrium of the system and, in spite
of the lower N concentration, the austenite volume fraction could increase as the
temperature decreased from 1200oC to 960oC leading to a subsequent migration of
the interfaces [17,21]. This sequence of events is evident in Figure 21 (b).
Precipitation of Cr2N occurred also at interfaces in the DN steel during aging
at 860oC (arrows in Figure 21 (c)).

26
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 21. (a) AN steel aged at 960oC for 10 h where was able to grow due to
changes in composition and Cr2N precipitated at interfaces.. TEM. (b) ADN steel aged at
o
960 C for 10 h showing precipitates at prior locations of the interfaces SEM. (c) DN steel

aged at 860oC for 15 min showing precipitates at interfaces (arrows) and an ongoing +
transformation. SEM. The arrows indicate prior interfaces decorated with precipitates.

Figure 22 summarizes the precipitation behavior of Cr2N observed at


grain boundaries and interfaces; in the discontinuous precipitation the shape of the
precipitates is conditioned by the preferred crystallographic orientation relation
with the growing austenite grain (a), the general precipitation at grain boundaries
does not result in specific shapes or growth directions (b), while at interfaces

27
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

the precipitates tend to grow toward the ferrite side consuming this phase during
growth (c).

Figure 22. Precipitation configurations at grain boundaries. (a) Discontinuous precipitation at


austenitic grain boundaries. (b) Continuous precipitation at austenitic grain boundaries. (d)

Precipitation at interfaces .

Ferrite formation

A high nitrogen concentration changes the equilibrium conditions and


stabilizes austenite. On the other hand, the formation of chromium nitrides, by
means of the different precipitation mechanisms, depletes the matrix of nitrogen
and tends to destabilize austenite in spite of the simultaneous chromium
consumption. The interplay between initial composition and prior or concurrent
phase transformations shapes the final microstructure for each specific thermal
cycle employed and requires careful attention during manufacturing and service.

28
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

In high nitrogen stainless steels ferrite can form during heat treatments by
distinct mechanisms:
(i) Ferrite is stable above 1250oC in agreement with the Fe-Cr-Ni [101] and
Fe-Cr-Ni-N [102] phase diagrams. This shows that at high temperatures the
interstitial element holds a weaker austenite stabilization effect.
(ii) During aging, chromium nitride precipitation depletes nitrogen in the
matrix inducing ferrite formation in the unstable austenite, often inside or near
lamellar colonies of Cr2N [17,20]. In this case ferrite can result: (a) directly
from the allotropic transformation [17,20] or (b) from a (Cr,Fe)2N+
eutectoid reaction as shown by Nakada [22].
(iii) The favored precipitation of Cr2N at interfaces further supports the
growth of ferrite enhancing its volume fraction.

The intense discontinuous precipitation of chromium nitride in the AN steel


at temperatures lower than 1040oC induced the formation of ferrite in the vicinity
of the cellular colonies. A crystallographic orientation relation of the Kurdjomov–
Sachs type has been found between this type of ferrite and the parent austenite
[71]. The morphology of the phases shown in Figure 23 suggests a mechanism of
type (ii(a)).
Figure 24 shows the volume fraction of ferrite after aging for the AN steel.
These results show that formation of ferrite can only be completely avoided in the
rather narrow 1250 - 1040 oC temperature range. The volume fraction of ferrite
can reach values around 10% in the AN steel for long aging times (Figure 24).
This value is compatible with a precipitation of half of the nitrogen and a
consequent chromium and nickel equivalent in the remaining austenite matrix of
26% and 20%, respectively [13]. The decrease in ferrite volume fraction observed
for some of the aging treatments suggests that the phase is consumed during other
transformation, such as sigma phase formation. Since ferrite is already present in
the initial microstructure of the ADN and DN steel, the slight volume fraction
changes during aging cannot be detected as precisely with ferritoscope
measurements for these alloys.

29
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 23. Ferrite formation in the AN steel during aging (a) at 960 oC for 100h and (b)
at 1040oC for 1 h. SEM.

Figure 24. Data relative to the volume fraction (%) of ferrite in the AN steel after the
aging treatments determined with ferritoscope measurements.

Sigma phase formation

Chromium depletion around sigma phase is detrimental for the corrosion


resistance and toughness of the stainless steels, therefore the formation of this
phase must be avoided. In fact, in traditional austenitic stainless steels of the AISI
300 series, sigma phase occurs in low volume fractions and only after hundreds or
even thousands of hours at temperatures ranging between 600 and 900oC [103].
Nevertheless, since the interfaces are preferential nucleation sites for sigma

30
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

phase, larger volume fractions are expected in duplex steels after aging in the
same temperature range. In addition, the formation of sigma is also assisted by the
faster diffusion rates in ferrite [ 104 , 105 ] and by the fact that its chemical
composition is closer to the composition of ferrite than that of austenite.
Sigma phase (S) was detected in austenite and ferrite of the AN, ADN and
DN steels and specific features associated with the presence of this phase will be
addressed below. The following mechanisms are proposed for sigma phase
formation in high nitrogen stainless steels:
(i) Directly from austenite due to austenite destabilization: allotropic
transformation [16,17,20,21,39].
(ii) From ferrite by cooperative growth with austenite through an +
eutectoid reaction [17,20].

Figure 25 shows sigma phase formed around + Cr2N colonies and at grain
boundaries in the austenite of the AN and ADN steels. The corresponding
configurations are schematized in Figure 26. The sigma phase volume fraction
and its transformation kinetics in the austenite of the AN and ADN steels lie
between those of duplex and austenitic stainless steels [17,20,21]. The formation
of ferrite often preceded the appearance of sigma phase, however, ferrite
formation did not necessarily lead to sigma phase for all aging treatments, namely,
ferrite was stable at 960oC whereas this was not the case for the sigma phase. The
regions where sigma phase formed bare similarities to the ones in which ferrite
was formed after nitride precipitation, and microstructural characterization
confirmed that the decrease in ferrite volume fraction in the AN steel (see Figure
24) was associated with sigma phase formation during the aging through a
reaction of type (ii). The AN steel presented a unique microstructure after aging at
lower temperatures, in which sigma phase was more abundant in austenite than in
ferrite. This behavior is justified by the fact that ferrite was not initially present
and was induced by Cr2N precipitation.
In the ADN steel the formation of sigma phase occurred with different
kinetics in the austenitic, transition and duplex regions, as could be expected from

31
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 10. The sigma phase was detected in austenitic regions of the ADN steel
after aging between 700 and 960oC (Figure 25 (b) and (c)) and after 10 hours at
860oC [21]. The resulting morphology and the regions hosting the transformation
were similar in the AN and ADN steel; since ferrite is not expected to be
necessarily implicated in this transformation sigma phase may also have formed
as result of austenite decomposition, i.e., following reaction (i). In austenitic
regions of the DN steel, the transformation also induced a unique microstructural
feature; the + Cr2N colonies formed by discontinuous precipitation were
completely surrounded by a relatively thick layer of sigma phase (bright regions
in Figure 21 (c)). This mechanism was probably induced by the destabilization of
austenite in the nitrogen-depleted regions around the colonies, making sigma
phase formation possible.
Figure 27 illustrates the formation of sigma phase in the ferritic regions of
the ADN steel. In the transition and duplex regions of the ADN steel, the sigma
phase nucleated at interfaces and grew consuming the ferrite following
reaction (ii). Sigma phase formation in the ferritic region of the ADN steel was
faster than in the austenitic region. However, in the transition region, which
contains less nitrogen than in the austenitic region and more than in the duplex
region, the formation of sigma phase was initiated at times similar to those
registered for the duplex region, although the final volume fractions were different
in each region [17,21]. During some heat treatments, the sigma phase consumed
nearly all ferrite available in the ADN steel. In duplex regions, the sigma phase
presented a massive morphology and the particles were larger than in the
(nitrogen-rich) austenitic regions (compare Figure 25 (c) with Figure 27 (b)).
Figure 27 (c) and (d) demonstrate that chromium nitrides and the sigma phase are
easily distinguished using backscattered electron imaging [17].
Figure 28 presents the microstructure of the DN steel where the presence of
sigma phase in ferrite is evident. Ferrite transformation into sigma phase and
austenite took place by eutectoid decomposition (see Figures 21 (c) and 28)
according to a reaction (ii) and resulted in a decrease in ferromagnetism that
correspond to ferrite volume fractions as low as 1% (see Figure 29).

32
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 25. (a) Sigma phase formation in the AN steel aged at 880oC for 100 h. (b) and
(c) Sigma phase formation in austenitic regions of the ADN steel aged at 850 oC for 100 h. SEM.

33
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 26. Schematic representation of sigma phase formation in the austenite after
Cr2N (a) discontinuous and (b) continuous precipitation.

34
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 27. Sigma phase formation in the duplex region of the ADN steel (a) and (b)
aged at 800oC for 100 h and (c) and (d) at 800oC for 10 h. SEM. The prior ferritic regions
transformed through a eutectic reaction ( S + ). The nitrides at interfaces and sigma phase
particles are clearly distinguished using (c) secondary and (d) backscattered electrons .

Figure 28. Ferrite decomposition in the DN steel aged at 860oC for 15 min. SEM
(backscattered electrons).

35
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

Figure 29. Data relative to the volume fraction (%) of ferrite in the DN steel after the
aging treatments determined with ferritoscope measurements.

CONCLUDING REMARKS

The continuing development of austenitic stainless steels has resulted in


complex compositions with substantial amounts of alloying elements, which are
introduced for diverse specific reasons, in order to obtain better mechanical
properties and/or corrosion resistance. However, as usual these additions
invariably induce pitfalls resulting from microstructural instability [1]. The
benefits of nitrogen in austenitic stainless steels arise mainly when this interstitial
element is in solid solution, as it stabilizes austenite, causes strong solid solution
hardening and inhibits the strain-induced martensite formation. However, high
nitrogen levels favor chromium nitride precipitation during heat exposure which
in austenite occurs through competing continuous and discontinuous mechanisms.
Chromium nitride precipitation depletes nitrogen and chromium in the austenitic
matrix and, as result, induces the formation of ferrite and sigma phase in the
microstructure. The study of these phase transformations is interesting from the
scientific point view and vital for technological advances in the field as the
decomposition of austenite must be avoided due to the deleterious consequences
to the mechanical properties and corrosion resistance. Processes like welding and
stress relief heat treatments require special attention. Nevertheless, for low

36
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

temperature applications, such as in biomaterials for implants, where no


precipitation occurs, high nitrogen austenitic stainless steels are most interesting
potential candidates [106].

ACKNOWLEDGMENTS

The authors are grateful to the Conselho Nacional de Desenvolvimento


Científico e Tecnológico (CNPq), to the Coordenação de Aperfeiçoamento de
Pessoal de Nível Superior (CAPES) and to the Fundação de Amparo à Pesquisa
do Estado de São Paulo (FAPESP) for financial support.

REFERENCES

[1] Padilha, A. F., Plaut, R. L., Rios, P. R., (2003). Annealing of cold-worked
austenitic stainless steels (Review). ISIJ International. 43, 135-143.
[2] www.worldstainless.org (website of International Iron and Steel Institute,
accessed on 14 June 2014).
[3] Padilha, A. F. , Rios, P. R., (2002). Decomposition of austenite in austenitic
stainless steels (Review). ISIJ International. 42, 325-337.
[ 4 ] Pimenta Jr, F. C. , Padilha, A. F. , Plaut, R. L., (2003). Sigma phase
precipitation in a superferritic stainless steel. Materials Science Forum. 426-32,
1319-1324.
[ 5 ] Zucato, I., Moreira, M. C., Machado, I. F. , Lebrão, S.M.G., (2002).
Microstrucutural Characterization and the effect of phase transformations on
toughness of the UNS S31803 duplex stainless steel aged treated at 850oC.
Materials Research, 5, 385-389.
[6] Moon, J., Ha, H., Lee, T., (2013).Corrosion behavior in high heat input welded
heat-affected zone of Ni-free high-nitrogen Fe–18Cr–10Mn–N austenitic stainless
steel. Materials Characterization. 82, 113-119.

37
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[ 7] Talha, M., Behera, C. K., Sinha, O. P., (2013). A review on nickel-free


nitrogen containing austenitic stainless steels for biomedical applications.
Materials Science and Engineering C. 33, 3563–3575.
[8] Lo, K. H., Shek, C. H., Lai, J. K. L., (2009). Recent developments in stainless
steels. Materials Science and Engineering R. 65, 39–104.
[9] Speidel, M. O., (1989). Properties and applications of high nitrogen steels. In:
Foct, J. and Hendry, A., ed. High Nitrogen Steels 88. London, Institute of Metals.
92-96.
[10] Adcock, F., (1926). The effect of nitrogen on chromium and some iron-
chromium alloys. Journal of The Iron and Steel Institute. 114, 117-126.
[11] Tofaute, W., Schottky, H., (1940). Zur Frage des Ersatzes von Nickel in
korrosionsbeständigen Chrom-Nickel-Legierungen durch Stickstoff. Technische
Mitteilungen Krupp. 3, 103-110.
[12] Schäffler, A. L., (1949). Constitution diagram for stainless steel weld metal.
Metal Progress. 56, 680-680B.
[13] Long, C. J., Delong, W. T., (1973). The ferrite of austenite stainless steel.
Welding Journal. 52, 281s-291s.
[14] Espy, R. H., (1982). Weldability of nitrogen-strengthened stainless steels.
Welding Journal. 61, 148s-156s.
[15] Pant, P., Dahlmann, P., Schlump, W., Stein, G., (1987). A new nitrogen
alloying technique-A way to distinctly improve the properties of austenitic steel.
steel research. 58, 18-25.
[16] Machado, I. F., (1995). Reações de precipitação no estado sólido em um aço
25%Cr-5,5%Ni contendo 0,87% de nitrogênio. M.Sc. Thesis. Escola Politécnica-
Universidade de São Paulo, São Paulo, Brazil, (in Portuguese)
[17] Machado I. F., (1999). Transformações de fase no estado sólido em alguns
aços inoxidáveis austeníticos e ferríticos austeníticos (dúplex) contendo altos
teores de nitrogênio. Dr.-Ing. Thesis. Escola Politécnica-Universidade de São
Paulo, São Paulo, Brazil. (in Portuguese)

38
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[ 18 ] Kliauga, A. M., (1996). Randschichtbeeinflussung von ferritisch-


austenitischen Chrom-Nickel-Stälen durch Stickstoffeinsatz. Dr.-Ing. Thesis,
Fakultät für Maschinenbau, Ruhr-Universität Bochum, Bochum, Germany. (in
German)
[19] Berns, H., Siebert, S., (1996). High nitrogen austenitic cases in stainless
steels. ISIJ International. 36, 927-931.
[20] Machado, I. F. , Padilha, A. F., (1996). Precipitation behaviour of 25%Cr-
5.5Ni austenitic stainless steel containing 0.87% nitrogen. steel research. 67, 285-
290.
[ 21 ] Machado, I. F. , Padilha, A. F. , Kliauga, A. M., (1998). Precipitation
behaviour of a nitrogenated surface layer of 1.4462 duplex stainless steel. steel
research. 69, 381-386.
[ 22 ] Nakada, N., Hirakawa, N., Tsuchiyama, T., Takaki, S., (2007). Grain
refinement of nickel-free high nitrogen austenitic stainless steel by reversion of
eutectoid structure. Scripta Materialia. 57, 153-156
[ 23 ] Ishizaki, K., Mineura, K., (1989). Load frequency dependent fatigue
properties in ultrahigh steel and iron. In: Foct, J. and Hendry, A., ed. High
Nitrogen Steels 88, Institute of Metals, London. 204-207.
[24] Dhers, J., Foct, J., Vogt, J-B., (1989). Influence of nitrogen on fatigue crack
growth rate at 77 K and 293 K of 316L steel. In: Foct, J. and Hendry, A., ed. High
Nitrogen Steels 88. Institute of Metals, London. 199-203.
[25] Reed, R. P., (1989). Nitrogen in austenitic stainless steels. The Journal of the
Minerals, Metals & Materials Society. 41, 16-21.
[26] Speidel, M. O., (1990). Properties of high nitrogen steels. In: Stein, G. and
Witulski, H., ed High Nitrogen Steels 90. Verlag Stahleisen, Düsseldorf. 128-131.
[27] Douglass, D. L., Thomas, G., Roser, W. R., (1964). Ordering, stacking faults
and stress cracking in austenitic alloys. Corrosion. 20, 15-28.
[28] Fawley, R, Quader, M. A., Dodd, R. A., (1968). Compositional effects on the
deformation modes, annealing twin frequencies, and stacking fault energies of

39
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

austenitic stainless steels. Transactions of the Metallurgical Society of AIME. 242,


771-776.
[ 29 ] Schramm, R. E., Reed, R. P., (1975). Stacking fault energies of seven
commercial austenitic stainless steels. Metallurgical Transactions. 6A, 1345-1351.
[30] Stoltz, R. E., Vander Sande, J. B., (1980). The effect of nitrogen on stacking
fault energy on Fe-Ni-Cr-Mn steels. Metallurgical Transactions. 11A, 1033-1037.
[31] Lee, T., Oh, C., Kim, S., (2008). Effects of nitrogen on deformation-induced
martensitic transformation in metastable austenitic Fe–18Cr–10Mn–N steels.
Scripta Materialia. 58, 110-113.
[32] Hwang, B., Lee, T., Park, S., Oh, C., Kim. S., (2011). Correlation of austenite
stability and ductile-to-brittle transition behavior of high-nitrogen 18Cr–10Mn
austenitic steels. Materials Science and Engineering. A 528, 7257– 7266.
[ 33 ] Ahn, T., Lee, S., Park, K., Oh, K., Han, H., (2014). Strain-induced   ε-
martensite transformation during nanoindentation of high-nitrogen steel.
Materials Science & Engineering. A598, 56–61.
[34] Simmons, J. W., (1996). High-nitrogen alloying of stainless steels. Materials
Science and Engineering. A207, p. 159-169.
[35] Xua, Z., Jia, C., Kuang, C., Chu, K., Qu, X., (2009). Spark plasma sintering
of nitrogen-containing nickel-free stainless steel powders and effect of sintering
temperature. Journal of Alloys and Compounds. 484, 924–928.
[36] Salahinejad, E., Amini, R., Hadianfard, M. J., (2012). Structural evolution
during mechanical alloying of stainless steels under nitrogen. Powder Technology.
215-216, 247–253.
[37] Noneder, H., Merklein, M., (2012). Manufacturing of complex high strength
components out of high nitrogen steels at industrial level. Transactions of
Nonferrous Metals Society of China. 22,  s512−s518
[38] Wischnowski, F., (1995). Einfluβ mikrostruktureller Gefügeveränderungen
auf die Korrosionsresistenz von nichtrostenden ferritisch-austenitischen Duplex-
Stählen. Dr.-Ing. Thesis, Fakultät für Maschinenbau, Ruhr-Universität Bochum,
Bochum, Germany. (in German)

40
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[39] Machado, I. F. , Padilha, A. F., (2000). Aging Behaviour of 25Cr-17Mn High


Nitrogen Duplex Stainless Steel. ISIJ International. 40, 719-724.
[40] Menzel, J., Kirschner, W., Stein, G., (1996). High nitrogen containing Ni-free
austenitic stainless steels for medical applications. ISIJ International, v. 36, p.
893-900.
[41] Peng, H., Wen, Y., Du, Y., Yu, Q., Yang, Q., (2013). Effect of Manganese on
Microstructures and Solidification Modes of Cast Fe-Mn-Si-Cr-Ni Shape Memory
Alloys. Metallurgical and Materials Transactions B. 44, 1137-1143.
[42] Hazraa, M., Rao, K. S., Reddy, G. M., (2014). Friction welding of a nickel
free high nitrogen steel: influence of forge force on microstructure, mechanical
properties and pitting corrosion resistance. Journal of Materials Research and
Technology. 3, 90–100.
[43]Hua-Bing, L., Zhou-Hua, J., Hao, F., Qi-Feng, M., Dong-Ping, Z., (2012).
Aging Precipitation Behavior of 18Cr-16Mn-2Mo-l.lN High Nitrogen Austenitic
Stainless Steel and Its Influences on Mechanical Properties. Journal of Iron and
Steel Research International. 19, 43-51.
[44] Krivobok, V. N., (1935). Alloys of iron and chromium. Transactions of the
American Society for Metals. 23, 1-60.
[ 45 ] Degallaix, S., Foct, J., (1988).   L’azote   dans   les   aciers   inoxydables  
austénitiques 2éme partie: influence sur les caractéristiques mécaniques.
Mémoires et Etudes Scientifiques Revue de Métallurgie. 85, 111-123.
[46] Uggowitzer, P. J., Speidel, M. O., (1991). Ultrahigh-strength Cr-Mn-N steels.
In: International Conference on Stainless Steels. Chiba, The Iron and Steel
Institute of Japan. 762-770.
[47] Gavriljuk, V. G., (1996). Nitrogen in iron and steel. ISIJ International. 36,
738-745.
[48] Rawers, J., Gocken, N. A. Pehlke, R. D., (1993). High nitrogen in Fe-Cr-Ni
alloys. Metallurgical Transactions A. 24, 73-82.

41
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[49] Pickering, F. B., (1989). Some beneficial effects of nitrogen in steel. In: Foct,
J. and Hendry, A., ed. High Nitrogen Steels 88. London, Institute of Metals,.
London. 10-31.
[50]Feng, S., Li-jun, W., Wen-fang, C., Yang, Q., Chun-ming, L., (2009). Aging
precipitation and recrystallization in high-nitrogen austenitic stainless steel.
Transactions of Nonferrous Metals Society of China. 19, s569-s572.
[ 51 ]Yu-Xi, M. , Fan, R., Rong, Z., Yu-Ping, L., Ye-Hua, J., (2007). On
Precipitation of High Nitrogen Containing Austenitic Stainless Steel During
Isothermal Aging at Intermediate Temperature. In: Proceedings of Sino-Swedish
Structural Materials Symposium. Sino-Swedish Advanced Materials Exchange
Centre. China. 344-349.
[ 52 ] Erisir, E., Prahl, U., Bleck, W., (2010). Effect of precipitation on hot
formability of high nitrogen steels. Materials Science and Engineering A. 528,
519–525.
[ 53 ] Lee, T., Kim, S., Takaki, S., (2006). Time-Temperature-Precipitation
Characteristics of High-Nitrogen Austenitic Fe-18Cr-18Mn-2Mo-0.9N Steel.
Metallurgical And Materials Transactions A. 37, 3445-3454.
[ 54 ] Zheng, X., (1991). Nitrogen solubility in iron-base alloys and powder
metallurgy of high nitrogen stainless steels. Ph. D. Thesis (Doctor of Technical
Sciences), Swiss Federal Institute of Technology, Zurich, Swiss.
[55] Kikuchi, M., Kajihara, M., Choi, Si-Kyung., (1991). Cellular precipitation
involving both substitutional and interstitial solutes: cellular precipitation of Cr2N
in Cr-Ni austenitic steels. Materials Science and Engineering A. 146, 131-150.
[56] Santhi Srinivas, N. C., Kutumbarao, V. V., (1997). On the discontinuous
precipitation of Cr2N in Cr-Mn-N austenitic stainless steels. Scripta Materialia.
37, 285-291.
[57] Anson, D. R., Pomfret, R. J. , Hendry, A., (1996). Prediction of the solubility
of nitrogen in molten duplex stainless steel. ISIJ International. 36, 750-758.
[58] Rawers, J, Dunning, J., Petty, A. V., (1990). High-pressure-nitrogen alloying
steels. Advanced Materials&Processes. 138, 50-52.

42
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[ 59 ] Okamoto, M., Tanaka, R., Naito, T., Fujimoto, R., (1962). On the
manufacture of high-chromium steels in high-pressure nitrogen atmosphere and
heat resistance properties of 316L type. Tetsu to-Hanagé Overseas. 2, 25-37.
[60] Hschenk, H., Frohberg, M. G., Heinemann, H., (1962). Untersuchungen zur
Stickstoffaufnahme in flüssigen Eisenlegierungen im Druckbereich bis zu vier
Atmosphären. Archiv für das Eisenhüttenwesen. 33, 593-600.
[61] Horovitz, M. B., Beneduce Neto, F., Garbogini, A. , Tschiptschin, A. P.,
(1996). Nitrogen bearing martensitic stainless steels: Microstructure and
properties. ISIJ International. 36, 840-845.
[ 62 ] Padilha, A. F., Machado, I. F., Randle, V., (1999). Microstructure and
microtexture changes during solution nitriding to produce austenitic case on
ferritic-austenitic duplex stainless steel. Materials Science and Technology. 15,
1015-1018.
[ 63 ] Atamert, S., King, J. E., (1991). Intragranular nucleation of austenite.
Zeitschrift für Metallkunde. 82, 230-239.
[64] Presser, R., Silcock, J. M., (1983). Aging behavior of 18Mn-18Cr high
nitrogen austenitic steel for end rings. Metal Science. 17, 241-247.
[ 65 ] Vanderschaeve, F., Taillard, R., Foct, J., (1994). Peculiarities of grain
boundary precipitation phenomena in high nitrogen austenitic stainless steels. In:
Solid Phase Transformations. TMS (The Minerals, Metals & Materials Society).
USA. 527-532.
[66] Shewmon, P. G., (1989). Diffusion in solids. 2. ed., TMS (The Minerals,
Metals & Materials Society), Pennsylvania.
[ 67 ] Kühl, A., Bergner, D., (1991). Chemische Diffusion von Stickstoff in
hochlegierten austenitischen CrNi(Mo)-Stählen. Materialwissenschaft und
Werkstofftechnik. 22, 462-467.
[68] Christian, J. W., (1965). The theory of transformations in metals and alloys.
Pergamon Press, London.
[ 69 ] Porter, D. A., Easterling, K. E. Mohamed, Y. S., (2009). Phase
transformations in metals and alloys. CRC Press, USA.

43
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[70] Hornbogen, E., (1969). Nucleation of precipitates in defect solid solutions. In:
Zettlemoyer, A. C. ed. Nucleation. Marcel Dekker Inc, New York. 309-378.
[71] Carvalho, P. A., Machado I. F., Solorzano, G., 2008. On Cr2N precipitation
mechanisms in high-nitrogen austenite, Philosophical Magazine. 88, 229–242.
[72] Gust, W. (1969). Discontinuous precipitation in binary metallic systems. In:
Phase Transformations, Chameleon (Institution Metallurgists), London. 1, 27-68.
[73] Tu, K. N.,Turnbull, D., (1967). Morphology of cellular precipitation of tin
from lead-tin bicrystals. Acta Metallurgica. 15, 369-376.
[ 74 ] Tu, K. N., Turnbull, D., (1972). The cellular reaction in Pb-Sn alloys,
Metallurgical Transaction. 3, 2769- 2776.
[75] Fournelle, R. A., Clark, J. B., (1972). The genesis of the cellular precipitation
reaction. Metallurgical Transactions A. 3, 2757-2767.
[ 76 ] Williams, D. B., Butler, E. P., (1981). Grain boundary discontinuous
precipitation reactions. International Metals Reviews. 26, 153-183.
[ 77 ] Yoon, D. Y., (1995). Theories and observations of chemically induced
interface migration. International Materials Reviews. 40, 149-179.
[78] Hillert, M., (1972). On theories of growth during discontinuous precipitation.
Metallurgical Transactions B. 3 2729-2741.
[79] Hillert, M. (1982). An improved model for discontinuous precipitation. Acta
Metallurgica, 30 1689- 1696.
[80] Duly, D. Simon, J.P., Brechet, Y., (1955). On the competition between
continuous and discontinuous precipitations in binary MgAl alloys. Acta
Metallurgica et Materialia. 43, 101-106.
[81] Kajihara, M., Choi, S., Kikuchi, M., Tanaka, R., Seo, Y., Okumura, T.,
Kondoh, Y., (1986). Evidence of long range diffusion of nitrogen in cellular
precipitation of Cr2N in Cr-Ni austenitic steel. Zeitschrift für Metallkunde. 77,
515-518.
[82] Rayaprolu, D. B., Hendry, A., (1989). Cellular precipitation in a nitrogen
alloyed stainless steel. Materials Science and Technology. 5, 328-332.

44
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[83] Kikuchi, M., Urabe, T. Cliff, G. Lorimer, G. W., (1990). The loss of driving
force due to volume diffusion ahead of a migrating boundary in a cellular
precipitation reaction. Acta Metallurgica et Materialia. 38, 1115-1120.
[ 84 ] Matsuoka, S., Mangan, M. A., Shiflet, G. J., (1994). Morphological
development of cellular colonies in a 19Cr-5Ni austenite steel. In: Solid Phase
Transformations. TMS (The Minerals, Metals & Materials Society), USA, 521-
526.
[85] Ramíres Londoño, A. J., (1997). Estudo da precipitação de nitreto de cromo
e de fase sigma por simulação térmica da zona afetada pelo calor na soldagem
multipasse de aços inoxidáveis duplex. M.Sc. Thesis. Escola Politécnica-
Universidade de São Paulo, São Paulo, Brazil. (in Portuguese)
[86] Clauss, A. R., Bischoff, E., Hosman, A. R., Schacherl, R. E., Mittemeijer, E.
J., (2009). Crystal Structure and Morphology of Mixed Cr1–xAlxN Nitride
Precipitates: Gaseous Nitriding of a Fe-1.5 Wt Pct Cr-1.5 Wt Pct Al Alloy,
Metallurgical and Materials Transactions. 40A 1923-1929.
[87] Beckitt, F. R. , Clarck, B. R., (1967). The shape and mechanism of formation
of M23C6 carbide in austenite. Acta Metallurgica. 15, 113-129.
[88] Singhal, L. K. and Martin J. W., (1967). The growth of M23C6 on incoherent
twin boundaries in austenite. Acta Metalurgica. 15, 1603-1610.
[89] Sasmal, B., (1984). Formation of lamellar M23C6 on and near twin boundaries
in austenitic stainless steels. Buletin of Materials Science. 6, 617-623.
[90] Martinez, L. G., Imakuma, K. , Padilha, A. F., (1992). Influence of niobium
on stacking fault energy of all-austenite stainless steels. Steel Research
International. 63, 221-223.
[91] Reick, W., Pohl, M., Padilha, A. F., (1996). Determination of stacking fault
energy of austenite in a duplex stainless steel. Steel Research International. 67,
253-256.
[92] Yakubtsov, I. A., Ariapour, A., Perovik, D. D., (1999). Effect of Nitrogen on
Stacking Fault Energy of F.C.C. Iron-Based Alloys Acta Materialia. 47, 1271-
1279.

45
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[93] Kibey, S., Liu, J. B., Curtis, M. J., (2006). Effect of nitrogen on generalized
stacking fault energy and stacking fault widths in high nitrogen steels. Acta
Materialia. 54, 2991-3001.
[94] Faulkner, R. G., (1993). Discontinuous precipitation kinetics in austenitic
steels. Materials Science and Technology. 9, 118-124.
[ 95 ] Sundararaman D, Shankar P., Raghunathan, V.S., (1996). Electron
microscopic study of Cr2N formation in thermally aged 316 austenitic stainless
steels. Metallurgical and Materials Transactions A. 27, 1175-1186.
[96] ASM Handbook. (1993). Properties and Selection: Irons, Steels, and High
Performance Alloys. 1.
[97] Manna, I., Pabi, S. K., Gust, W., (1991). Discontinuous precipitation in a Cu-
12at% In alloy. Acta Metallurgica et Materialia. 39,1489-1496.
[ 98 ] Hamana, D.; Boumerzoug, Z., (1994). Discontinuous precipitation,
coarsening and dissolution of phases in Cu-In and Cu-Sb alloys. Zeitschrift für
Metalkunde, 85, 479-486.
[99] Ramirez A. J., Lippold J. C., Brandi S. D., (2003). The relationship between
chromium nitride and secondary austenite precipitation in duplex stainless steels.
Metallurgical and Materials Transactions A. 34, 1575-1597.
[ 100 ] Frisk, K. A., Hillert, M., (1989). Thermodynamics of the Fe-Cr-Ni-N
system. in: Foct, J. and Hendry, A., ed. High Nitrogen Steels 88, Institute of
Metals, London. 1-9.
[101] Mundt R., Hoffmeister, H., (1983).   γ   phase   equilibria   in   Iron-rich Iron-
Chromium- Nickel at temperatures between 1200-1350oC. Archiv für das
Eisenhüttenwesen. 54, 253-266.
[ 102 ] Hoffmeister, H., Mundt, R., (1981). Untersuchungen zum Einfluß des
Kohlenstoffs und des Stickstoffs auf die Umwandlung ferritisch- austenitischer
Chrom-Nickel-Stähle. Archiv für das Eisenhüttenwesen. 52, 159-164.
[103] Weiss, B., Stickler, R., (1972). Phase instabilities during high temperature
exposure of 316 austenitic stainless steel. Metallurgical Transactions A. 3, 851-
866.

46
Izabel Fernanda Machado, Patrícia Almeida Carvalho and, Angelo Fernando Padilha

[104] Villanueva, D. M. E., Pimenta Jr, F. C., Plaut, R. L., Padilha, A. F., (2006).
Comparative study of sigma phase precipitation of three types of stainless steels:
austenitic, superferritic and duplex. Materials Science and Technology. 22, 1098-
1104.
[ 105 ] Reick, W. , Pohl, M. , Padilha, A. F., (1998). Recrystallization-
transformation combined reactions during annealing of a cold rolled ferritic-
austenitic duplex stainless steel. ISIJ International. 38, 567-571.
[106] Terada, M., Antunes, R. A., Padilha, A. F., Costa, I., (2007). Corrosion
resistance of three austenitic stainless steels for biomedical applications. Materials
and Corrosion. 58, 762-766.

47

View publication stats

Potrebbero piacerti anche