You are on page 1of 5

Austenitic Stainless Steels

Austenitic stainless steels are an extraordinary family


of alloys that have exceptional corrosion resistance
and equally impressive mechanical properties. They
have unsurpassed strength, toughness, and formability
among the commercially viable alloys from cryogenic
to elevated temperatures. They are also valued aesthetically and are environmentally benign. Table 1 lists
the most common austenitic grades and two
precipitation hardening grades. Although only about
10.5% chromium is required to render iron alloys
stainless, austenitic stainless steels all contain at least
15% chromium with which nickel, manganese,
carbon, and nitrogen are combined to stabilize the f.c.c.
structure.
Austenitic stainless steels have been in use since the
1920s, but advances in processing implemented in the
1970s made them much more cost effective for common use in applications where coated or plated steels,
aluminum, or copper alloys may otherwise be
preferred. Other inherent factors contributing to their
increased use is long service life with low maintenance
cost, ability to be recycled, and benign effect on the
environment and human health.

1. Processing
The processing advances that have most benefited
stainless steels are the argonoxygen decarburization
(AOD) and the continuous caster processes, but
significant advances have been made in hot and cold
rolling mills as well as in the annealing line.
Stainless steel is typically produced from a combination of recycled scrap and virgin charge materials
melted in an electric arc furnace. After melting it is
transferred to another vessel, an AOD one, for
refining, i.e., to reduce the carbon level from about 1%
to about 0.05%. In the AOD vessel a mixture of
oxygen and argon is injected into the molten metal at
around 1700 mC. The ratio of argon to oxygen is
controlled to keep the partial pressure of carbon
monoxide sufficiently low such that carbon is oxidized
in preference to chromium. At the end of the refining
cycle, the small chromium losses are recovered from
the slag by the addition of silicon.

This process improved the productivity of melt


shops, permitted lower cost raw materials and higher
metallic yields, and resulted in cleaner, lower carbon,
lower sulfur stainless steels. Variants on the AOD
process using vacuum-assisted decarburization also
find occasional use.
Equally important in the development of austenitic
stainless steel was the transfer of continuous casting
technology from carbon steel. Because of the tendency
of the chromium in molten stainless steel to reoxidize,
elimination of the ingot teeming step was a major
improvement in cleanliness. Continuous casting also
eliminated the high cost of slabbing and its poor yield.
Nicodemi (1993) reviews the state of the art.
Hot rolling of slabs is increasingly performed on hot
reversing mills rather than tandem mills because their
flexibility in pass schedule and coil box capability
allows the hot rolling to be better tailored to the high
hot hardness and narrow rolling temperature range of
austenitics.
Austenitics can be used as hot rolled or further
reduced by cold rolling but they are generally annealed
and pickled in either case. Cold reductions of 80% are
possible but require cluster mills because of the high
work hardening rates of austenitics. Final annealing is
conducted either in air or in hydrogen\nitrogen
protective atmospheres.
The heat treatment of austenitic stainless steel is
simply annealing, which is done from about 1000 mC to
1100 mC. The function of annealing is to eliminate the
effects of cold work by recrystallization, to control
grain size, and to dissolve any precipitated carbides.
Cooling from annealing temperatures must be
sufficiently rapid to avoid carbide precipitation. No
martensite can form from quenching. Mechanical
deformation is required to transform these alloys.
2. Structure
Austenitic stainless steels are designed as solid solution
alloys. From a structural point of view it is convenient
to classify the alloying elements by the degree to which
they replicate the effects of chromium in stabilizing
ferrite or nickel in stabilizing austenite. Each element
plays a separate role in corrosion resistance or
strengthening, so alloy design seeks to simultaneously
satisfy the objectives of structure, corrosion resistance,
mechanical properties, cost, and fabricability.

Table 1
Principal grades of austenitic stainless steel.
Grade

Carbon

Nitrogen

Manganese

Silicon

Chromium

Nickel

S20100
S30100
S30400
S31600
S17400

0.15
0.15
0.08
0.08
0.07

0.25

5.57.5
2.0
2.0
2.0
1.0

1.0
1.0
1.0
1.0
1.0

16.018.0
16.018.0
18.020.0
16.018.0
15.017.5

3.55.5
6.08.0
8.010.5
10.014.0
3.05.0

S17700

0.09

1.0

1.0

16.018.0

6.57.75

0.10
0.10

Molybdenum

Other

2.03.0
3.05.0 Cu;
0.150.45 (NbjTa)
0.751.5 Al

Austenitic Stainless Steels


Austenitic stainless steels are designed to initially
contain several percent ferrite after solidification and
cooling to ambient temperatures. The amount of this
ferrite, which dissolves during normal processing, can
be calculated from (Nicodemi 1993)
% ferrite l 3(Crj1.5SijMo)k2.8(Nij0.5 Mn)
k84(CjN)k19.8%

(1)

Suutala (1983) correlated the solidification mode, i.e.,


as primary dendrites of austenite vs. ferrite, as crucial
to hot cracking in welds, and by extension to hot
workability. Suutalas formula for predicting the
solidification mode gives 1.55 as the critical value for
Cr eq\ Ni eq above which solidification occurs in the
preferred ferritic mode which results in fewer
impurities, such as sulfur, phosphorus, selenium, tin,
lead, etc., being rejected to interdendritic grain boundaries. In this relationship the equivalencies are expressed as
Cr eq l % Crj1.37% Moj1.5% Si
j2% Nbj3% Ti (2)
Ni eq l % Nij0.31% Mnj22% C
j14% Nj% Cu (3)
Since the continuous caster puts a premium on hot
workability, austenitic alloys are designed to solidify
as ferrite retaining a small amount of delta ferrite,
which dissolves in subsequent processing. This
approach also makes them weldable without danger
of weld cracking.

3. Austenite Stability
A critical characteristic of the austenitic phase is its
stability. As in carbon steels, stainless steels with
metastable austenite can transform to martensite
either by cooling or by deformation. The martensite
start temperature, Ms, is well below room temperature
in all commercial alloys, but the strain-induced transformation is nearly always present in normal processing involving cold rolling, deep drawing, wire drawing,
grinding, or polishing. The quantity of martensite
formed depends on the temperature, strain rate, and
the composition. The influence of composition is
described by (Angel 1954)
Md30(C) l 497k462(% Cj% N)k9.2% Si
(4)

where Md30 is the temperature at which 50% martensite forms at a true strain of 30%.
2

4. Carbide and Nitride Precipitation


Austenitic stainless steels can contain up to 0.15%
carbon. Carbon is an impurity more than an alloying
element. It increases yield strength and helps stabilize
austenite, but, if not kept in solid solution, it combines
with chromium as M C , where M is principally
#$ ' This carbide precipitates
chromium, but also iron.
initially at grain boundaries, then at twins, then within
austenite grains. The grain boundary precipitation
leaves the surrounding volume significantly depleted
of chromium. This phenomenon is called sensitization and it is particularly detrimental to corrosion
resistance. The kinetics of this precipitation are such
that grades with 0.08% carbon sensitize in minutes
while those with less than 0.03% require 100 times
longer. Increasing nickel and molybdenum and cold
work accelerate the precipitation, while nitrogen and
manganese retard it. Reducing carbon levels has
become the preferred way to avoid sensitization in
service. Some alloys are designed to eliminate the
precipitation of chromium carbides by preferentially
forming titanium or niobium carbides. In certain hightemperature applications this can be more appropriate
for austenitics while it is the normal method in ferritic
stainless steels, where no practical refining method can
reduce carbon levels sufficiently to prevent sensitization.
The carbide precipitation reaction occurs between
600 mC and 1000 mC. At similar temperatures other
intermetallic phases, principally sigma, can precipitate
at much longer times in the more highly alloyed
industrially used grades. This occurs in alloys where
the matrix is rich in chromium equivalents, over 18%,
and low in dissolved carbon. Such phases are of much
more practical importance in ferritic grades where
their kinetics of formation is about 100 times faster.

5. Hardening of Austenite

k8.1% Mnk13.7% Cr
k20% Nik18.5% Mo

All commercially used alloying elements stabilize


austenite, even those that promote ferrite during
solidification. All austenitic grades have roughly comparable yield strengths when annealed and abundant
ductility, but their work hardening rates vary greatly,
because the strain-induced martensite is a much harder
phase than the austenite from which it forms. This
factor is extensively exploited in selecting alloys whose
work hardening rate is optimized for the forming
method preferred.

Interstitial alloying elements, carbon, nitrogen, and to


a limited degree boron, have a pronounced strengthening effect on austenite, compared to the more
modest influence of substitutional solid solution
elements and of grain size. Empirical relationships

Austenitic Stainless Steels


have been developed to quantify these effects (Pickering et al. 1969):
Y0.2% (MPa) l 15.4[4.4j23(C)
j32(N)j1.3(Si)j0.24(Cr)
j0.94( Mo)j1.2( V )
j0.29( W )j2.6( Nb )
j1.7(Ti)j0.82(Al )
j0.16(% ferrite)j0.46dk" ] (5)
#
where d is the mean grain diameter and t is the twin
spacing, both in millimeters. Nitrogen is the most
effective agent for increasing the strength of austenite
because it has the largest strengthening coefficient and
high solubility. Since nitrogen also suppresses carbide
and other phase precipitation, enhances corrosion
resistance, and has negligible cost, it is a crucial
component of modern austenitic alloy design.
One of the most characteristic features of austenitic
stainless steels is their pronounced strain hardening,
which at a minimum of 0.4 is far above that of ferritic
steels. The strain-induced transformation of metastable austenite to martensite can increase that coefficient to nearly 1.0. Cold-worked austenitic stainless
steels are commonly used at tensile strengths
approaching 2000 MPa.

6. Precipitation Hardening Stainless Steels


Another class of austenitic stainless steels is the
precipitation hardenable stainless steels. There are
also a related group of martensitic precipitation
hardenable stainless steels, more familiarly known as
maraging steels, which are discussed elsewhere. The
austenitic precipitation hardenable grades rely on the
precipitation of a coherent intermetallic h phase
from a stable, supersaturated austenite matrix. The
precipitate is composed of nickel, aluminum, and
titanium and a typical precipitation treatment involves
aging at 725 mC for 16 h after a 1000 mC solution anneal.
A286 and Discaloy are two prominent examples of
this alloy family.
A second, more common, group is the semiaustenitic precipitation hardenable stainless steels,
which have a carefully balanced lean austenite. Hightemperature, e.g., 1065 mC, solution treatment produces a metastable austenite, which is soft enough to
fabricate. Subsequent conditioning at 750 mC destabilizes the austenite by precipitating Cr C , which
#$ '
results in an austenite that can thermally transform
to
martensite to produce a primary hardening. This
martensite is in turn aged at 500600 mC to produce a
true precipitation hardening via nickel and aluminum
h. 17-7PH and 15-7PH are representative of this
group.

7. Formability
Because of their exceptional ductility, austenitic stainless steels have excellent formability. The most common cold-forming procedure is drawing and it is
useful to distinguish two modes. In the first, called
deep drawing, a flat blank is pulled freely into a die by
the action of a punch with no added restraint. A
second type of drawing restrains the periphery of the
blank and the metal is stretched into the die. This is
drawing by expansion.
Carbon steels and ferritic stainless steels both deep
draw well because the anisotropic nature of flat-rolled
b.c.c. structures resists thinning in the thickness
direction, giving them a drawability beyond that which
their inherent ductility would suggest. Austenitic
stainless steels are to a first approximation isotropic,
so they lack that benefit, but make up for it with their
exceptional ductility.
Work hardening rate is critical in expansion forming
because localized thinning which would lead to fracture is countered by the increase in strength in the
critical location where stretching is the greatest. The
controllably high work hardening rate of austenite is a
great advantage and permits extraordinary stretch
formability.
Among austenitic stainless steels, those with the less
stable austenite such as 301, e.g., 17% chromium, 7%
nickel, are favored for items such as sinks where part
geometry lends itself to stretch forming. High aspect
ratio cylindrical parts made in multistage dies without
intermediate annealing use richer grades to lessen or
completely eliminate martensite formation. 305 with
up to 19% chromium and 13% nickel represents that
extreme. Between the two a continuum of levels of
chromium and nickel are used depending on the mode
of deformation that must be satisfied.
Surface finish is always a key consideration. Grain
size must be kept sufficiently fine such that the surface
relief produced within grains during deformation does
not become visible. This is the orange peel defect.
The defect of Luders bands, stretcher strains, does not
occur on austenitic stainless steels as it does on mild
steels and ferritic stainless steels. Neither does the
defect known as ridging or roping. As-formed
austenitic parts can have high hardness and high
residual stresses. This can make them susceptible to
brittle delayed failure if they have residual hydrogen
content from previous bright annealing operations.

8. Corrosion
Whatever their other attributes, the austenitic stainless
steels owe their widespread use to their resistance to
corrosion. This corrosion resistance can be attributed
to the tendency of stainless steel to form a passive film.
In environments where passivity cannot be maintained, the corrosion of stainless steels can proceed
3

Austenitic Stainless Steels


rapidly. Fortunately, stainless steel is passive in more
normally encountered environments, so from a design
point of view, it is considered inert and with indefinite
life.
Passivity is the formation on the surface of a thin (a
few nanometers) layer of tightly adherent and chemically stable oxide or hydroxide with low ionic permeability. Because of its low ionic permeability, such a
layer cannot grow, so further consumption of the
underlying metal in the reaction that formed the layer
ceases. Corrosion effectively stops.
Chromium is the element whose concentration in
the passive film gives it its stability. In the least
aggressive aqueous environments about 11% chromium in the alloy is necessary for stable passivity, but
in more aggressive environments, higher concentrations are required. The exact nature of the passive film
is not well understood, but it can be thought of as an
ordered iron\chromium oxide that undergoes transition to a more hydrated form facing the environment.
The film forms spontaneously in air or water and is
self-healing. Films are not stable in all environments,
however. The principle factors that influence film
stability are the difference in potential between the
metal and the solution, the temperature, the acidity,
and the halide concentration. Much empirical data are
available to determine which stainless alloy is stable
in a given environment. From a design point of view
the main objective is always to design for stable
passivity.
The more significant consideration in the use of
stainless steel is not general corrosion, but localized
corrosion, which most commonly takes the form of
pitting. Crevice corrosion is a parallel phenomenon.
Pitting is insidious because with very little loss of
metal, perforation could cause component failure.
Avoidance of pitting requires knowledge of the
environment, which permits the appropriate grade
selection.
The resistance of austenitic stainless steels to pitting,
and by extension to crevice corrosion, is proportional
to their content of those elements that enhance passive
film stability and the absence of those factors that
diminish it. The resistance to pitting based on alloy
composition is commonly expressed as its pitting
resistance equivalent number, PREN (Sedriks 1984),
which is given by
PREN l % Crj3.3(% Mo)j16(% N)

(6)

This relationship explains the widely perceived difference in corrosion performance between ferritic and
austenitic stainless steels. Ferritics contain essentially
no nitrogen in solution, while austenite dissolves an
amount equivalent in performance to 12% of
chromium.
Corrosion resistance can be degraded by any
inhomogeneityonthesteelsurfacethatisnotasresistant
to the environment as the steel that surrounds it. So it
4

can be a site of attack, but further, when dissolved, it


may introduce ions that prevent formation of a passive
film at that site. Such is the case with the very common
MnS inclusion. Unless removed by pickling at the
producing mill or in the field by passivation, i.e.,
cleaning the surface with nitric or other acid, sulfides
seriously degrade pitting performance.
Designers should be especially wary of pipe and
tubing, which often has high sulfur levels to facilitate
tube welding, but at the expense of corrosion
resistance. 316 tubing with 0.015% sulfur has no better
pitting resistance than 304 with less than 0.003%
sulfur. Titanium will form sulfides in preference to
those of manganese and thereby eliminate the harmful
effect of sulfur.
Any disruption to the surface after pickling runs the
risk of degrading pitting resistance. The most common
problem is that caused by abrasive polishing of the
surface to produce a no. 3 or a no. 4 finish. These very
eye-pleasing finishes diminish pitting resistance by
about the same amount as the effect of sulfides cited
above. This can be alleviated by pickling after polishing or by using an embossed finish of the same
appearance.
Crevice corrosion is a related phenomenon in which
mechanical restrictions to fluid flow permit crevices to
become more acidic than the bulk solution with the
consequent danger of loss of passivity. Avoidance of
crevices is thus a design priority while resistance to
crevice corrosion is similar to resistance to pitting
corrosion.
Stress corrosion cracking is a form of brittle delayed
failure, which generally occurs at 100 mC or above in
austenitic stainless steels when a certain threshold
stress is exceeded. A necessary precondition seems to
be pitting, so again if pitting can be avoided in the
initial design, few hazards need be feared from
corrosion-related phenomena.
A full treatment of the corrosion of stainless steel
can be found in Lacombe et al. (1990).

9. Oxidation and Creep


The same tendency of chromium to form an adherent,
stable oxide with low ionic permeability gives austenitic stainless steels excellent oxidation resistance.
Increasing chromium and nickel enhance oxidation
resistance to a maximum at the 25% chromium\20%
nickel level of the 310 grade. Additions of other
elements that form stable oxides, such as silicon and
aluminum, can also enhance performance. Calcium
and rare-earth elements can greatly enhance the
adherence of Cr O and result in reduced oxidation
$
rates, apparently#through
their strong effect on sulfur.
The primary weakness of stainless steels in hightemperature use is the tendency of the protective oxide
layer to spall in thermal cycling. In such cases ferritic
stainless steels can be preferred.

Austenitic Stainless Steels


Austenitic stainless steels have superior creep
resistance to go with their oxidation resistance. The
austenite structure is stable and its deformation less
temperature dependent than that of ferrite. Numerous
factors can enhance creep resistance. Interstitial
solutes, such as carbon, boron, and nitrogen, help.
Substitutional solutes, such as titanium, niobium, and
molybdenum, have a strong effect. Elements such as
vanadium and zirconium, which form stable carbides,
have a predictably good influence. As in other alloys,
large grain size improves creep.

10. Applications
Austenitic stainless steels are used for domestic,
industrial, transport, and architectural products based
primarily on their corrosion resistance but also for
their formability, their strength, and their properties at
extreme temperatures. Because their initial cost is
often higher than that of alternative materials, their
popularity is based on their minimization of cost over
the entire life cycle of their use.
The food, pharmaceutical, chemical, pulp and
paper, and petrochemical industries depend heavily on
austenitic stainless steels because their corrosion
resistance yields low maintenance, lack of product
contamination, high cleanability, and long life. Ease of
welding and fabrication are important in these applications, but stainless is used simply because it is the
most economical material that can do the job. Initial
cost is often a poor measure of a material, such as in
the case of concrete reinforcing bars, where the entire
structure is jeopardized by corrosion of a minor
component. Specifying stainless steel can eliminate
such problems at a small premium to the overall
cost.
Sometimes formability becomes an equally important a requirement as corrosion resistance. Stretchformed parts, such as sinks, are an example of the use
of the more unstable austenitic grades, while the highly
stable grades are used for deeply drawn parts or for
components where low magnetic permeability is
sought. Automotive filter bodies, pen cartridges,
cooking pots, and disk drive parts are examples of the
latter.

The high strength of cold-worked austenitic stainless steel makes it the predominant material for use in
transit cars, but also in springs, seatbelt anchors, and
knife blades.
Since austenitics are tough even to liquid helium
temperatures they are widely used in all cryogenic
applications. Uniquely, they are equally useful for uses
up to 800 mC, where they find wide use in heat
exchangers, boilers, turbines, furnaces, and automotive exhaust systems, where the formability of
ferritics or their creep resistance is insufficient.
Nordberg and Bjorklund (1992) contains numerous
papers on the many industrial uses of stainless steel.
But, like the noble metals, stainless steel can also be
used simply for its aesthetic appeal. The numerous
surface finishes that can be applied to stainless steel,
from mirror to matte, do not degrade over time and
keep their appearance as well as their functionality.
See also: Stainless Steels : Martensitic; Stainless Steels:
Duplex; Ferritic Stainless Steels; Stainless Steels:
Cast

Bibliography
Angel T 1954 J. Iron Steel Inst. London 177, 165
Blank J 1988 Stainless Steels 87. Institute of Metals, London
Boyer H, Gall T 1985 Metals Handbook Desk Edition. American
Society For Metals, Metals Park, OH
Lacombe P, Baroux B, Beranger G 1990 Les Aciers Inoxydables.
Les Editions de Physique, Paris
Nicodemi W 1993 Innoation in Stainless Steel. Associazone
Italiana di Metallurgia, Milan
Nordberg H, Bjorklund J 1992 Applications of Stainless Steel.
Kristianstads Boktryceri, Stockholm
Peckner D, Bernstein I 1977 Handbook of Stainless Steel.
McGraw-Hill, New York
Pickering F, Irvine K, Gladman T 1969 J. Iron Steel Inst. London
1017
Sedriks A 1979 Corrosion of Stainless Steel. Wiley-Interscience,
New York
Sedriks A 1984 Stainless Steels 84. Institute of Metals, Go$ teborg, Sweden
Suutala N 1983 Metall. Trans. 1917

M. F. McGuire

Copyright ' 2001 Elsevier Science Ltd.


All rights reserved. No part of this publication may be reproduced, stored in any retrieval system or transmitted
in any form or by any means : electronic, electrostatic, magnetic tape, mechanical, photocopying, recording or
otherwise, without permission in writing from the publishers.
Encyclopedia of Materials : Science and Technology
ISBN: 0-08-0431526
pp. 406411
5

You might also like