You are on page 1of 31

International Journal of Plasticity 22 (2006) 1234–1264

www.elsevier.com/locate/ijplas

Constitutive modelling and identification


of parameters of the plastic strain-induced
martensitic transformation in 316L stainless
steel at cryogenic temperatures
a,*
Cédric Garion , Bła_zej Skoczeń b, Stefano Sgobba a

a
CERN, European Organization for Nuclear Research, CH-1211 Geneva 23, Switzerland
b
CUT, Cracow University of Technology, Al. Jana Pawla II 37, Krakow, Poland

Received 2 April 2005


Available online 2 November 2005

Abstract

The present paper is focused on constitutive modelling and identification of parameters of the
relevant model of plastic strain-induced martensitic transformation in austenitic stainless steels at
low temperatures. The model used to describe the FCC ! BCC phase transformation in austen-
itic stainless steels is based on the assumption of linearization of the most intensive part of the
transformation curve. The kinetics of phase transformation is described by three parameters:
transformation threshold (pn), slope (A) and saturation level (nL). It is assumed that the phase
transformation is driven by the accumulated plastic strain p. In addition, the intensity of plastic
deformation is strongly coupled to the phase transformation via the description of mixed kine-
matic/isotropic linear plastic hardening based on the Mori–Tanaka homogenization. The theory
of small strains is applied. Small strain fields, corresponding to phase transformation, are decom-
posed into the volumic and the shear parts. The grade AISI 316L, stainless steel often used in
cryogenic applications, has been chosen as a good example of the austenitic structure. The mag-
netic permeability of fine gauge stainless steel sheets (thickness 0.15–0.25 mm) subjected to mono-
tonic straining was measured as a function of strain. The detailed methodology of relevant
measurements is presented in the paper. Tuning of the constitutive model is described and the
relevant parameters are identified. The model has been applied in the design of thin-walled

*
Corresponding author.
E-mail address: Cedric.Garion@cern.ch (C. Garion).

0749-6419/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2005.08.002
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1235

bellows expansion joints for the large Hadron Collider (LHC), at present under construction at
CERN.
 2005 Elsevier Ltd. All rights reserved.

Keywords: A. Phase transformation; A. Strengthening mechanisms; B. Elastic–plastic material; B. Constitutive


behavior; C. Mechanical testing

1. Introduction

1.1. General

The present paper is focused on constitutive modelling and identification of the param-
eters of a consistent model of austenitic stainless steels (SS) for cryogenic applications. The
model describes one out of three phenomena, driven by the plastic strains, that occur in
metastable stainless steels at low temperatures: the phase transformation, from the FCC
to BCC lattice. The other two phenomena of dissipative nature: the discontinuous yielding
and the evolution of microdamage, are not taken into account in the present paper. The
martensitic transformation yields the initially homogeneous material strongly heteroge-
neous as a result of the presence of martensite platelets embedded into the austenitic ma-
trix. Since the BCC a 0 -martensite is much more rigid than the FCC c-austenite, its presence
influences the plastic flow and the process of hardening. The intensity of phase transfor-
mation affects the level of stress for a given total strain imposed on a sample. In the present
paper, the elastic–plastic material with mixed kinematic-isotropic linear hardening is taken
into account (Garion and Skoczeń, 2001, 2002, 2003). The volume fraction of martensite
affects both the parameters of kinematic and isotropic hardening. The hardening process
of two-phase material has been described on the basis of the Mori–Tanaka homogeniza-
tion (Garion and Skoczeń, 2002). The kinetic law of phase transformation has been pre-
sented in a convenient linearized form, where the rate of transformation is proportional to
the rate of the accumulated plastic strain. One of the possible applications for such a mod-
el are thin-walled corrugated shells called bellows, subjected to kinematically controlled
expansion/compression cycles between room and cryogenic temperatures. SS bellows, of-
ten used at cryogenic temperatures as compensation elements (Skoczeń, 2004), develop
strong plastic strain fields in the convolutions. Such fields contribute to the strain-induced
martensitic transformation that modifies the FCC c-austenite into the BCC a 0 -martensite.
The c–a 0 transformation, usually localised at crest and at root of convolutions, has an im-
pact on the local evolution of ductile damage and fatigue life of bellows but also on the
magnetic permeability of the shell. Since the a 0 martensite is known to be ferromagnetic,
an intensive phase transformation yields the structure highly susceptible to magnetic field.
This has a crucial impact on the performance of the expansion joints applied in the con-
struction of particle accelerators for high energy physics. The bellows, designed for particle
accelerators, are often located in close proximity of the superconducting magnets and in
their stray field. High magnetization of the bellows convolutions may affect the correct
functioning of the accelerator. Therefore, a correct choice of material, oriented towards
maximum stability at cryogenic temperatures (minimum phase transformation), is an
important issue. In order to minimize the amount of the c–a 0 phase transformation due
1236 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

Table 1
Results of chemical analysis for fine gauge 316L plates, thickness 0.15 and 0.25 mm
Element Specification Thickness 0 .15 mm content (%) Thickness 0 .25 mm content (%)
C 0.030 max. 0.027 0.029
Si 1.00 max. 0.596 0.492
Mn 2.00 max. 1.557 0.872
P 0.030 max. 0.022 0.025
S 0.010 max. 0.0009 0.0022
Cr 16.00–18.50 17.99 17.91
Ni 11.00–14.00 10.29 11.03
Mo 2.00–2.50 2.032 1.97
Co 0.20 max. 0.122 0.195
N 0.050 max. 0.0075 0.0074

to deformation between room temperature and 1.9 K a special 316L grade has been devel-
oped (see Table 1) with increased stability with respect to martensitic transformation. Fine
gauge samples of stainless steel were strained at room and at low temperatures and the
magnetic permeability was measured. Since the magnetic permeability is proportional to
the volume fraction of martensite, evaluation of the amount of the phase transformation
was possible. The results of this experimental work were then used to identify the param-
eters of the relevant constitutive model (Garion and Skoczeń, 2002). Limitation of the vol-
ume fraction of martensite is specified with respect to the so-called magnetic failure, that
means the mechanism of magnetic permeability exceeding a predefined allowable level.

1.2. Stainless steels subjected to martensitic transformation

Fe–Cr–Ni austenitic SS are commonly used to manufacture components of supercon-


ducting magnets and cryogenic transfer lines since they retain their ductility at low temper-
atures. They are initially paramagnetic in their fully austenitic state. A fully austenitic
microstructure is the result of proper balance of chemical composition and thermo-
mechanical transformation processes. Some of the nitrogen strengthened stainless steels
of 300 series belong to the group of metastable austenitic alloys. Under certain conditions
the steels undergo martensitic transformation at cryogenic temperatures which may lead to
a considerable evolution of material properties and to a ferromagnetic behaviour. The
martensitic transformations are induced mainly by the plastic strain fields and amplified
by the high magnetic fields. Spontaneous transformations due to the cooling process occur
in the least alloyed grades of the family such as 304L, while N-strengthened grades, such as
304LN and 316LN do not generally show spontaneous transformation during cooling
(Morris et al., 1992). At room temperature, the SS of the 300 series shows the classical
c-phase of FCC austenite. This phase may transform either to a 0 phase of body centered
tetragonal (BCT) ferrite or to a hexagonal e phase. The most often occurring c–a 0 trans-
formation leads to the formation of the martensite sites dispersed in the surrounding aus-
tenite matrix. In the course of the strain-induced transformation, the martensite platelets
modify the FCC lattice leading to local distortions. The amount of martensite depends on
the chemical composition, temperature, stress state, plastic strains and exposure to mag-
netic field. It is known that alloying elements considerably stabilize the c-phase. For in-
stance, the strain-induced martensitic content in the grades 304LN, 316LN at a given
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1237

temperature and for the same level of plastic strain, is much lower than in the grades 304L,
316L (Morris et al., 1992).
Tensile properties of stainless steels at low temperatures are strongly influenced by the
plastic strain-induced martensitic transformation. As a result of the transformation, the
initially homogeneous c-phase looses its homogeneity due to the progressive development
of the harder martensite phase. The martensite platelets embedded in the soft austenitic
matrix provoke local stress concentration and block the movement of dislocations. There-
fore the existence of martensite induces additional strain hardening. The effect of strain-
induced phase transformation on the mechanical properties of stainless steels at low
temperatures was investigated by many authors, both for monotonic and for cyclic loads.
The influence of the plastic strain-induced martensitic transformation on the low-cycle
fatigue of stainless steel (Fe18Cr6.5Ni0.19C) was studied by Baudry and Pineau (1977).
The authors applied both the magnetic measurements and the X-ray technique in order
to detect and identify the a 0 martensite. For the cyclic loads, they introduced a critical va-
lue of the plastic strain range ðDcp Þ. Below this threshold, the material did not transform,
whereas above this threshold the presence of a 0 martensite in the austenitic matrix strongly
influenced the fatigue life. The authors suggested that the fatigue crack propagation was
accelerated due to the formation of the a 0 phase. The process of cyclic loading was kine-
matically controlled. Another extensive study of mechanical strength of grades 304L and
316L subjected to monotonic and cyclic loads at 77 and 4 K was carried out by Suzuki
et al. (1988). The cyclic hardening curves and the fatigue life curves were presented in
the paper. The magnetic permeability has been measured as a function of the tensile strain.
The evolution of magnetic permeability with monotonic tensile straining and with cyclic
straining was analysed. The authors presented both the classical mechanical failure crite-
rion and a new magnetic failure criterion (unacceptable increase of magnetic permeabil-
ity). Similar curves showing the increase of a 0 martensite as a function of strain for
grades 304L and 304LN at 77 K were reported by Morris et al. (1992).
The strength of austenitic stainless steel (German designation X2 CrNi 19 11) subjected
to cyclic deformation at 103 K was experimentally analysed by Bayerlein et al. (1992). The
authors applied the strain-controlled process, with the plastic strain range of 1% and 2%.
Symmetric and asymmetric strain programmes were carried out. Formation of martensite
was detected by using the transmission electron microscopy. The cyclically deformed sam-
ples were in the second stage of analysis subjected to tensile tests at room temperature. It
turned out that the cyclic hardening, correlated with the phase transformation, led to a
rather fast saturation process (18 cycles in the case of symmetric loading and 23 cycles
for asymmetric loading). A substantial increase of the 1% offset yield stress (up to
200%) and tensile strength (by 75%) measured during tensile test at RT and after cycling
at low temperature was reported.
The NDT techniques were used by Grosse et al. (2001) to detect the martensite produced
in the austenitic piping steels exposed to low-cycle fatigue damage. The phase transforma-
tion has been observed in the metastable steel as a function of damage evolution in the pre-
crack stage. The transformation was driven by the accumulated plastic strain. Density and
distribution of martensite sites was measured by using the neutron diffraction technique.
One of the most important problems related to multi-phase materials is the question
how does the microstructure affect the fatigue crack propagation. The effect of the micro-
structural factors on the fatigue crack propagation in two-phase steels was studied by
Chen et al. (1989). The authors investigated different morphologies of martensite in
1238 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

as-rolled and heat-treated steels in the laboratory air and in 3.5% NaCl solution. It turns
out that the martensite morphologies affect in a substantial way the threshold values of the
stress intensity range and the fatigue crack propagation rates. Microstructures containing
the acicular or netted martensite show higher threshold values and lower fatigue crack
growth rates when compared to the granular martensite. In the near-threshold regions,
the fracture is characterized by the cyclic cleavage facets or the intergranular mode for
the samples tested in the air or 3.5% NaCl, respectively. In the regions of higher stress
intensity range, the fracture is represented by cyclic cleavage facets, secondary cracks
and striations. Enhancement of the threshold values has been observed in the corrosive
environment. The authors explain this phenomenon by higher fracture surface roughness
and influence of the corrosion products on the stronger crack closure.
The problem of fracture toughness of TRIP steels has been extensively discussed by
Iwamoto and Tsuta (2002). The phase transformation near the crack tip has been ana-
lysed. The fracture toughness for a metastable material with and without phase transfor-
mation was computed by using the path integrals technique. The favourable effect of the
strain-induced martensitic transformation on the fracture toughness of TRIP steels has
been highlighted. An interesting concept of optimization of the mechanical properties of
TRIP steels by using the strain-induced martensitic transformation was recently presented
by Iwamoto (2004). The author postulated a strong dependence of the mechanical prop-
erties of TRIP steels on the size, shape, orientation and spatial distribution of martensitic
inclusions in the microstructure of the material. The mechanical properties can be im-
proved by controlling the parameters of microstructural inhomogeneity of the material.
The model developed by the author is based on the Eshelby type ellipsoidal inclusion lo-
cated in the austenitic matrix. In order to obtain the macroscopic properties, the homog-
enization technique has been applied.
Finally, it is worth mentioning a special class of materials (TRIP steels) with the con-
tinuously graded microstructure, called the magnetically graded materials. They are ob-
tained by using the c ! a 0 phase transformation techniques. A smooth distribution of
the magnetization function across the specimen is characteristic of this class of materials.
As the c phase is paramagnetic and the a 0 phase is ferromagnetic the increase in martensite
fraction promoted by plastic deformation can be detected by measuring the magnetic per-
meability l. As a result of the controlled phase transformation, a required profile of the
magnetization along the sample can be obtained. However, it is necessary to control simul-
taneously the temperature and the strain distribution in the course of the process, as
shown by Watanabe and Sakai (2003).

1.3. Models of the martensitic transformation for stainless steels

The physically based transformation kinetics has been developed by Olson and Cohen
(1975). The authors identify the strain-induced martensite nucleation sites as shear-band
intersections. The analysis leads to the following equation for the volume fraction of mar-
tensite (n) as a function of plastic strain (p):
n
n ¼ 1  expðbð1  expðap ÞÞ Þ; ð1Þ
where a denotes the rate of shear-band formation, b represents the probability that a
shear-band intersection will become a martensite nucleation site and n is a fixed exponent.
The transformation curves (volume fraction of martensite versus plastic strain) show
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1239

Fig. 1. Volume fraction of martensite versus plastic strain at cryogenic temperatures.

typical sigmoidal shapes with saturation levels usually well below 100% (Fig. 1). The
constitutive modelling of the strain-induced martensitic transformation in SS is to a large
extent a consequence of the pioneer work carried out by Olson and Cohen. A constitutive
equation predicting the stress as a function of strain, strain rate and temperature for meta-
stable austenites has been derived by Narutani et al. (1982). The contributions of two ma-
jor factors controlling the flow stress: static hardening and dynamic softening were
separated. The transformation strain corrected rule of mixtures was used to describe the
static hardening. The dynamic softening was derived by treating the martensitic transfor-
mation as a deformation mechanism. The quantitative description of both contributions
to the flow behaviour led to the constitutive model for which the knowledge of the trans-
formation curves and the flow properties of both phases is needed. Another constitutive
model developed by Stringfellow et al. (1992) for non-thermoelastic alloys introduced a
generalized version of the Olson–Cohen transformation kinetics, where the evolution of
martensite was a function of temperature, plastic strain and the stress state. The authors
assumed that the transformation process generates a nucleation strain that can be decom-
posed into the deviatoric and the hydrostatic components. Isotropic hypoelastic formula-
tion is applied to describe the stress state evolution. The strain softening resulting from the
transformation strain was incorporated into the model. The self-consistent approach
describing the deviatoric plastic strain rate has been applied. The Eshelby solutions for
isotropic, incompressible spherical inclusions embedded in an infinitely extended incom-
pressible isotropic matrix were used as a basis for the localization law. The law has been
extended to the case of nonlinear viscous materials, for which the deformation within the
inclusion was no longer uniform.
Another approach, based on microplasticity coupled with the phase transformation was
proposed by Fisher et al. (1998). The formulation is based on the conditional extremum
problem. The dissipation inequality under the yield and transformation constraints consti-
tutes the core of the formulation. Full coupling between the phase transformation and
plasticity has been taken into account in the constitutive equations. A specific form of
the Gibbs potential as well as the yield and transformation conditions have been pre-
sented. A quantitative prediction of the volume fraction of martensite in austenitic stain-
less steel under thermo-mechanical loading was derived by Diani and Parks (1998). The
authors focus their attention on the deformation mechanisms in the grain. The analysis
1240 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

is restricted to the austenitic steels with low stacking fault energy. It is assumed that
formation of the shear-bands is an irreversible process, the bands being locked by the
a 0 -martensite phase. Furthermore, for sufficiently low temperatures only two mechanisms
of shear-band formation are activated: twining or formation of e-martensite. In addition
the authors assume that within a given grain maximum of two out of 12 directions of pos-
sible coordinated movement of atoms are simultaneously active at a given increment of
loading. As the shear-band mechanisms do not imply significant volume change, the clas-
sical theory of plasticity is applied. The transformation strain (eigen-strain) is not taken
into account. The elastic strains are neglected as small when compared to plastic strains.
The only mechanism of inelastic deformation allowed for is the crystallographic slip, rep-
resented by the homogenized shear rate c_ a within the slip system a. The effective slip pair
or single slip are chosen within each grain; 66 pairs in the case of double slip and 12 single
slips are taken into account. The material forms a polycrystalline aggregate subjected to
macroscopic strain. The macroscopic velocity gradient is an additive combination of the
deformation rate D ~ and the spin W
~ . The deformation rate is projected onto the 5-dimen-
sional second-order deviatoric space, generated by the independent Schmid tensors related
to the slip systems. The global/local relation is postulated for the deformation rate,
whereas the local spin is assumed equal to the macroscopic one. The resolved shear stress
is calculated from the homogenized shear rate by using the power law in the form:
 a m
_ 
a c
s ¼ s   signð_ca Þ;
a
ð2Þ
c_ 0
where sa denotes the slip system hardness that obeys evolution law based on the shear rates
and hardening moduli. One of the main problems in this micromechanical model is the opti-
mum choice of no more than two active slip systems for each grain. The criterion is based on
the maximum shear principle for the pairs of slip systems and maximum positive resolved
shear strain for single slips. The criterion leads to the selection of different active systems
from grain to grain, as a function of grain orientation. Finally, at the intersection of shear
bands the nucleation process of a 0 -martensite takes place. The volume fraction of martensite
in the grain is derived by assuming that all bands of a given slip system are equidistant and
periodically spaced. The model is valid for small amounts of the transformed phase and re-
flects the essential features of the experimental results obtained by Angel (1954). Also, the
warm-rolling textures were used as initial conditions to perform several numerical tests of
low-temperature uniaxial tension in the rolling direction. The initial rolling textures increase
substantially the amount of a 0 -martensite produced in the austenitic matrix.
A mesoscopic continuum thermo-mechanical approach applied to the strain-induced
martensitic transformation was developed by Levitas et al. (1992). Finite plastic and trans-
formation strains and small elastic deformations were taken into account. The model was
based on the multiplicative decomposition of the total deformation gradient into elastic,
transformation and plastic parts. The generalized Prandtl–Reuss equations for isotropic
elasto-plastic materials, including large plastic and transformation strains, were applied.
Transformation criterion and extremum principle for determination of location and vol-
ume of transformed domains were formulated. A revised formulation of the transforma-
tion-induced plasticity has been presented by Fisher et al. (2000). Both the influence of
transformation-induced plastic strain and the influence of shear (orientation effect) on
the irreversible deformation were taken into account. A new constitutive description for
the elasto-plastic materials subjected to the phase transformation has been developed.
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1241

Cherkaoui et al. (2000) developed a micromechanical analysis of the internal stress


sources, resulting either from the incompatible transformation strain accompanying
the phase transformation or from the plastic flow of both phases due to the motion
of dislocations. The authors underline that the main complexity in the modelling of
the behaviour of TRIP steels consists in evaluation of the effect of internal stresses on
the c ! a 0 phase transformation. They take into consideration a microscale mechanism
of martensitic phase transformation. Furthermore, formation of martensite microdo-
mains with moving boundaries inside inhomogeneous plastic strain fields is assumed.
Coherent interfaces with discontinuities of stress and strain fields across the multiple
moving boundaries are introduced. Helmholtz free energy is chosen as a potential
describing the thermodynamic state of the two-phase system. The morphology of micro-
domains is represented by an inclusion ellipsoidal in shape, where one dimension is
much smaller than the other two dimensions. Eshelby description of an inelastic ellipsoi-
dal inclusion has been adopted. Finally, a suitable kinetics of the martensitic phase
transformation has been developed and a complete constitutive description was applied
to an austenitic single crystal. Elasto-plastic self-consistent algorithm is used to make a
transition to the polycrystalline TRIP steels. Numerical simulation of the behaviour of a
polycrystalline TRIP steel has been presented.
Tomita and Shibutani (2000) approach refers to the earlier model by Tomita and Iwam-
oto (1995), where the deformation behaviour by formation of shear bands is highly strain
rate dependent. The evolution law for the volume fraction of martensite is given by the
equation:
0 a 0
na ¼ Að1  n_ Þ_pslip
aeq ; ð3Þ
where _ pslip
aeq is the equivalent strain rate of slip deformation in the austenitic matrix. The
function A is expressed by the following formula:
n1
A ¼ apngf sb ð1  f sb Þ; ð4Þ
sb
where f is the volume fraction of shear bands, p denotes the probability that an intersec-
tion of shear bands forms a martensitic embryo (see Olson and Cohen (1975)), n, g are geo-
metric constants and a is a function related to the stacking fault energy. The total strain
rate has been decomposed into elastic and plastic components. The plastic strain rate has
been further decomposed into the strain rate induced by slip in both phases _ pslip and a
component induced by the phase transformation _ ptrans . The latter has been split into
the deviatoric _ pshape and dilatational _ pdilat parts. The dilatational part is directly related
to the volume change Dv during the phase transformation (typical value for austenitic
stainless steels reported by the authors is 0.02–0.05). On the other hand, the strain rate rep-
resenting the change of shape is assumed to be coaxial with the deviatoric stress. In the
constitutive description, the stress rate has been derived as
 ¼ E : ð_  _ p Þ  bT_ ;
r ð5Þ

where r  represents the Zareba–Jaumann derivative of the Kirchhoff stress and b is the
thermal expansion tensor. The authors apply the Eshelby theory to describe the mar-
tensitic inclusions in the austenitic matrix (two-phase composite). In order to identify
the parameters of the constitutive model the uniaxial tests of the 304 stainless steel
have been performed. Also, numerical simulations of the deformation behaviour of
1242 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

notched bars at 77 K were carried out and the essential features of the deformation
were confirmed.
Some further research on the influence of phase transformations on the properties of
austenitic stainless steels were made by Lebedev and Kosarchuk (2000). The commercial
18Cr–10Ni austenitic stainless steel grades were studied. For one of the steels, with lower
content of Ni, the plastic strain-induced phase transformation occurred already at room
temperature. A detectable amount of a 0 martensite was present for plastic strains below
10%. The authors analyse presence of the a 0 martensite and the hexagonal  martensite
in the austenitic matrix. Three types of loading were taken into account: tension, compres-
sion and torsion. The authors indicate that the intensity of the phase transformations de-
pends strongly on the type of loading. Tension causes more intense formation of the 
phase than torsion or compression. Both, the kinetics of c !  and c ! a 0 phase transfor-
mations were considered. The maximum of  phase fraction reached in the tests was of the
order of 8–9%, whereas for the a 0 phase and under torsion the maximum was close to
100%. The effect of phase transformations on the strain hardening of Cr–Ni stainless steels
was discussed. Three characteristic stages of the strain hardening process were distin-
guished on the curves of microhardness against plastic strain intensity:

 rapid increase due to hardening of the austenite and formation of  martensite (stage I),
 decrease of the strain hardening rate, no more increase of the  martensite and forma-
tion of small quantities of phase a 0 (stage II),
 strong hardening of the steel due to intensive c ! a 0 phase transformation (stage III).

Finally, the authors indicate that with the increase of the stress triaxiality parameter
and decrease of the Lode parameter the c ! a 0 phase transformation significantly inten-
sifies. Thus, the Lode parameter turns out to be of fundamental importance for the
martensitic transformation at low temperatures. A new and broader constitutive
description including the effect of strain rate, temperature as well as the applied stress
has been presented by Tomita and Iwamoto (2001). The model was tested on grade
304 austenitic stainless steel and for the temperature range 77–353 K. The phase trans-
formation (evolution of martensite) has been described as a function of temperature,
prestrain, applied stress as well as the stress range. Numerical tests on steel bars with
ringed notches under cyclic loads were performed. A general formulation of the finite
thermoplasticity model including the phase transformation has been presented by
Dachowski and Boehm (2004). The model is based on the concept of isomorphism of
elastic ranges. The evolution of different phases is described in terms of the mass frac-
tions, treated as the internal variables. As the model is general, it can be used to de-
scribe different types of thermo-mechanical processes, where the phase transformation
takes place.
Quite recently a number of papers on the transformation kinetics as well as on the
behaviour of TRIP steels were published. A microstructure-based computational model
has been developed by Heung Nam et al. (2004). Again the shear-band intersections were
assumed as the locations of the martensite nucleation sites. The activation of the nucle-
ation sites was function of the interaction energy between the stress state and the lattice
deformation. A self-consistent approach was applied to describe the deformation behav-
iour of each material phase. The uniaxial tension and simple shear were used as the test
cases for comparison between the experiments and the numerical simulation.
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1243

In the model by Oberste-Brandenburg and Bruhns (2004), a macroscopic driving force


for the movement of the phase boundary has been introduced. It is independent of the lo-
cal description and represented by the Eshelby tensor lR. Since the second-order tensor
has been chosen as the thermodynamic driving force for the phase transformation, a cor-
responding second-order tensor ftf representing the thermodynamic flux and linked to the
driving force by the equation:
o~
u
ftf ¼ ð6Þ
olR

has been introduced. Here, u


~ is the thermodynamic potential function.
The macroscopic description of the phase transformation kinetics is based on the tensor
l, which represents difference of the chemical potential between the austenitic and the
martensitic phase
l ¼ lR;a  lR;m ð7Þ
and a kinetic variable related to the growth of the martensite inclusions (sites)
R;m
n_ ¼ n_ . ð8Þ
In the description of the transformation kinetics, an assumption is made that the transfor-
mation-induced dissipation reaches its maximum in the transformation process
l : n_ ! max . ð9Þ
n
A transformation threshold function F has been introduced and during the transforma-
tion Fn = 0. A normality law (similar to the associated flow rule) has been postulated (as a
result of a constrained maximization problem)
n
oF
n_ ¼ kn . ð10Þ
ol
Finally, a set of internal variables (similar to hardening variables in plasticity) has been intro-
duced and the relevant evolution laws were defined. During the transformation the function
Fn behaves in a similar way like the surface of plasticity in the model with mixed kinematic
and isotropic hardening. The model is general enough to cover a wide range of experimen-
tally confirmed transformation kinetics. The model is very attractive, however a large set
of parameters needs to be identified for each material and for chosen transformation kinetics.
Iwamoto (2004) concentrated on the deformation behaviour of TRIP steel with growth of
martensite sites due to strain-induced martensitic transformation. The effect of the geomet-
rical configuration of martensite sites on the macroscopic behaviour of TRIP steels has been
investigated. Growth of ellipsoidal martensitic inclusion embedded in the austenitic matrix
has been analysed using a homogenization method. The author assumes that the mechanical
properties of TRIP steels can be enhanced due to controlled geometrical characteristics of the
martensite fraction in the austenitic matrix. In the analysis, the transformation-thermo-cou-
pled asymptotic homogenization was applied. The model includes the transformation strain
rate and latent heat induced by the phase transformation. The following kinetic law governs
the growth of martensitic inclusion in the austenitic matrix:
   
_ tÞ ¼ ½1  nðx; tÞ Aða; g; p; nsb Þh_p ðx; tÞi þ B g; dp ; nsb gðx;
nðx; _ tÞ ; ð11Þ
eq
dg
1244 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

where n is the volume fraction of martensite, nsb is the volume fraction of shear bands, p is
the probability that an intersection of shear bands forms a martensitic embryo (see Olson
and Cohen, 1975), a is a parameter related to the stacking fault energy, g is a geometric
parameter, h_peq i denotes the average of the equivalent plastic strain rate of austenite
and g stands for the driving force of the phase transformation. The author takes into
account exclusively the growth of the martensitic inclusion. The nucleation process is
not accounted for. The thermo-elasto-viscoplastic analysis has been performed by using
the finite element method. The effect of strain, strain rate, temperature, the shape and
the spatial orientation of the martensitic inclusions as well as the effect of the direction
of loading on the deformation of TRIP steels have been discussed. The author indicates
the importance of work hardening in the martensitic phase to enhance the strength of
TRIP steels.
If the strain-induced transformation occurs at very low temperatures (liquid nitrogen
77 K, liquid helium 4.2 K) then the steep part of the transformation curves (see Fig. 1) re-
mains in the domain of relatively small strains (below 0.2). In such a case, the constitutive
modelling can be considerably simplified and stays within the scope of the classical rate
independent theories of plasticity. It has been shown by Hecker et al. (1982) that below
the von Mises true effective strain of 0.25 the strain rate sensitivity of the production of
volume fraction of martensite in grade 304 austenitic stainless steel was much lower than
for the strain levels exceeding 0.25. Most probably a similar phenomenon takes place in
the other austenitic stainless steel grades. The relevant elasto-plastic model with linear
mixed (isotropic/kinematic) hardening, including the effect of strain-induced martensitic
transformation, has been developed by Garion and Skoczeń (2002). The model is attrac-
tive in view of its simplicity and relatively small number of parameters to be identified at
cryogenic temperatures. Especially the tests in liquid helium (at 4.2 K) or in superfluid he-
lium (beyond the scope of the present paper) are laborious and require rather complex
cryogenic installations. In the present paper, the parameters of the constitutive model
for 316L fine gauge stainless steel sheets (thickness: 0.15 and 0.25 mm) are identified at
three temperature levels: 293 (room temperature), 77 and 4.2 K.

2. Constitutive model of 316L stainless steel for low temperature applications

2.1. General equations

The basic thermo-elastoplastic constitutive equations including phase transformation


were formulated by Garion and Skoczeń (2002). All variables are defined at the meso-
scopic level, i.e. in the representative volume element (RVE, Fig. 2).
A simplified evolution law for the martensite content has been introduced with respect
to the linear part (region II) of the sigmoidal curve (Fig. 3) in the following form:
dn ¼ AðT ; _ p ; rÞdpH ððp  pn ÞðnL  nÞÞ; ð12Þ

where A is a function of the temperature T, the stress state r and the plastic strain rate _ p , p
denotes the accumulated plastic strain, pn is the accumulated plastic strain threshold (to
trigger the formation of martensite) and nL is a limit of martensite evolution, above which
the martensitic transformation rate is considered equal to 0. Here, H represents the Heav-
iside function and p_ denotes the accumulated plastic strain rate defined by
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1245

Fig. 2. The representative volume element (RVE): definition of the parameter n.

Fig. 3. Linearization of the region II of the martensitic transformation.

rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 p
dp ¼ d : dp . ð13Þ
3

The following constitutive law is applied:


 
r ¼ E :   p  th  PT ; ð14Þ

where p and th stand for the plastic and the thermal strain tensors, respectively. PT is the
mesoscopic strain tensor representing the phase transformation. The tensor PT is related
to the microscopic eigen-strain tensor ePT by the following equation:
Z
PT 1
 ¼ ePT dV ; ð15Þ
V V
where V is the volume of the RVE.
Taking into account that, in the austenitic phase,
ePT ¼ 0; ð16Þ
Eq. (15) transforms to
1246 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264
Z
Va 1
PT ¼ ePT dV a ¼ n~ePT ; ð17Þ
V Va Va

where ~ePT is a tensor representing the average (over the RVE) of eigen-strain in the mar-
tensitic phase. In the case of c–a 0 martensitic transformation (displacive transformation),
the eigen-strain tensor has the following form (Wechsler et al., 1953; Fischer, 1997):
0 1
0 0 2c
B C
ePT ¼ @ 0 0 0 A ð18Þ
c
2
0 Dv ð~
x;~
y;~

in the local co-ordinate system ð~


x;~ zÞ. Here, the ð~
y;~ yÞ-base plane corresponds to the habit
x;~
plane of a martensite variant with its normal ~ z. c represents the transformation shear and
Dv denotes the volume change during the transformation.
Assume that the material is initially isotropic, i.e. the grains are randomly oriented in
the RVE. As the martensite orientation is determined with respect to the initial austenitic
phase, also the martensite platelets are randomly oriented in the RVE. Thus, PT and ~ePT
turn out to be purely isotropic tensors. Since the first invariant of the tensor, ~ePT reads
~ePT ¼ 13DvI ð19Þ
the mesoscopic free strain tensor, Eq. (17) takes the form
PT ¼ n13DvI. ð20Þ
For isotropic material, the elastic stiffness tensor E is given by
E ¼ 3kJ þ 2lK; ð21Þ

where
8
< J ¼ 13 I  I ðJ ijkl ¼ 13 dij dkl Þ;

:K ¼ I  J and I ijkl ¼ 12 dik djl þ dil djk .


E E
Here l ¼ 2ð1þmÞ , k ¼ 3ð12mÞ , expressed as functions of the modulus of elasticity E and the
Poisson coefficient m, denote the shear and the compressibility moduli, respectively.
The yield surface for the rate independent plasticity is given by
fc ðr; X ; RÞ ¼ J 2 ðr  X Þ  ry  R; ð22Þ
where
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3
J 2 ðr  X Þ ¼ ðs  X Þ : ðs  X Þ ð23Þ
2
is the second invariant of the stress tensor. Here X is the back stress tensor, ry and R stand
for the yield point and the isotropic hardening variable, respectively. It is assumed that the
quasi-isotropic two-phase material obeys the normality rule with the yield function postu-
lated as the potential of plasticity. Thus the plastic flow rate is given by
ofc 3 sX
dp ¼ dk ¼   dk; ð24Þ
or 2 J2 r  X
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1247

where k is the plastic multiplier (time-like parameter). Since, the hardening variables R and
X are altered by the presence of martensite, the corresponding evolution laws are postu-
lated in the following form:
dR ¼ f ðnÞdp; ð25Þ
dX ¼ dX a þ dX aþm ¼ 23Cdp þ 23GðnÞdp ¼ 23gðnÞdp . ð26Þ
It is assumed that the back stress increment is composed of a classical term which corre-
sponds to the behavior of the austenitic phase dXa and a term related to the presence of
martensite in the austenitic matrix (dXa + m). Furthermore, it is assumed that the mecha-
nism of plastic flow at low temperatures is based on the motion of dislocations in the lat-
tice. If the massive motion of dislocations occurs they are stopped by the martensite
inclusions and the corresponding local stress fields.
The principal components that constitute the two-phase material model are the elasto-
plastic matrix (austenite) and the elastic inclusions (martensite platelets). Linear kinematic
hardening law is applied to model the plastic behavior of the pure austenite phase
dX a0 ¼ 23C 0 dp ; ð27Þ
where C0 is the hardening modulus of the original non-transformed purely austenitic
phase. For the two-phase transformed material, the hardening modulus C0 is replaced
by the modulus C, that is higher than C0 because of the interactions between the disloca-
tions in the austenite and the martensite inclusion
C ¼ C 0 uðnÞ for 0 6 n 6 nL ; uð0Þ ¼ 1. ð28Þ
The function u(n) has been linearized and takes the form
uðnÞ ¼ hn þ 1; ð29Þ
where h is a material-dependent parameter. The function u(n) can be interpreted as this
part of the hardening process that corresponds to the increase in volume fraction of mar-
tensite and enhanced probability that a dislocation will stack on an inclusion. The back
stress increment corresponding to the behavior of austenite can be decomposed in the fol-
lowing way:
dX a ¼ dX a0 þ dX an ¼ 23C 0 dp þ 23C 0 hndp ; ð30Þ

where dXan corresponds to the interactions between the dislocations in the austenitic ma-
trix and the martensite inclusions.
For the pure austenitic phase, a linearization of the stress/strain relations in the vicinity
of the current state is expressed by
Dra ¼ Et : D; ð31Þ

where Et is the tangent stiffness tensor. A similar principle can be applied to the two-phase
continuum. However, the stress increment is obtained by the homogenization process
Draþm ¼ EH : D. ð32Þ

The hardening increment induced by the presence of the martensite is given by


Dr ¼ Dra  Draþm ¼ ðEH  Et Þ : D. ð33Þ
1248 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

For the plastically active processes, the linearization in the vicinity of the current state al-
lows us to apply the homogenization technique. For a given load increment, the local lin-
earized stiffness of the austenite is described by the corresponding tangent modulus
Eta ¼ 3k ta J þ 2lta K; ð34Þ

where
Et Et EC
lta ¼ ; k ta ¼ ; Et ¼ . ð35Þ
2ð1 þ mÞ 3ð1  2mÞ EþC
As the inclusions are isotropic and elastic, the corresponding modulus of elasticity is given
by
Em ¼ 3k m J þ 2lm K; ð36Þ

where
E E
lm ¼ ; km ¼ . ð37Þ
2ð1 þ mÞ 3ð1  2mÞ
The inclusions are assumed spherical and uniformly distributed in the matrix. Application
of the Mori–Tanaka homogenization yields (Mori and Tanaka, 1973)
EH ¼ EMT ¼ 3k MT J þ 2lMT K; ð38Þ

with EMT obtained from the following relation:


1 X 1
EMT þ E ¼ fi ½Ei þ E  ; ð39Þ
i¼a;m

where fi is the volume fraction of the component ‘‘i’’ and E* stands for the Hill influence
tensor. After some rearrangements the following equations are obtained
 1
1n n
3k MT þ 3k  ¼ þ ; ð40Þ
3ðk ta þ k  Þ 3ðk 1 þ k  Þ
 1
 1n n
2lMT þ 2l ¼ þ ; ð41Þ
2ðlta þ l Þ 2ðlm þ l Þ
4 l ð9k ta þ 8lta Þ
k  ¼ lta and 2l ¼ ta . ð42Þ
3 3ðk ta þ 2lta Þ
As the strain increment is mainly due to the plastic strains: D @ Dp, Eq. (33) becomes
 
Dr ¼ EH  Et : Dp . ð43Þ

The plastic strains are represented by a deviatoric tensor, therefore


J : Dp ¼ 0; K : Dp ¼ Dp . ð44Þ

Finally, the stress increment due to the formation of martensite is equal to


Dr ¼ 2ðlMT  lta ÞDp . ð45Þ
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1249

If pure kinematic hardening is considered, the evolution of the back stress for two-phase
material obeys the following equation:
DX aþm ¼ Dr ð46Þ
or in the incremental form

dX aþm ¼ 2ðlMT  lta Þdp . ð47Þ

On the other hand, if pure isotropic hardening is considered, the evolution of the harden-
ing parameter is obtained by imposing the second invariant of the stress tensor
DR ¼ DRaþm ¼ J 2 ðDrÞ ¼ 3ðlMT  lta ÞDp; ð48Þ
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where Dp ¼ 23 Dp : Dp . In the incremental form, one obtains
dR ¼ dRaþm ¼ 3ðlMT  lta Þdp. ð49Þ
This linearized approach to the evolution of isotropic hardening is replaced by a more gen-
eral nonlinear formulation, suitable for higher martensite content
dR ¼ ðR1 ðnÞ  RÞdp; ð50Þ
where R1 is the parameter that defines the ultimate size of the yield surface
R1 ðnÞ ¼ 3ðlMT  lta Þ. ð51Þ
Eq. (48) can be reduced to Eq. (39) in the vicinity of the initial state.
In order to establish a proportion between the kinematic and the isotropic hardening,
the Baushinger parameter b is introduced
r0 þ r0
b¼ ; 0 6 b 6 1; ð52Þ
2ðr0 þ r0 Þ

where r 0 denotes the stress level at unloading and r 0  is the stress level corresponding to
the reverse active process. The parameter varies between 0 for the isotropic hardening (no
Bauschinger effect) and 1 for the kinematic hardening (perfect Bauschinger effect). Thus,
the mixed hardening is described by the following model:

dX ¼ 2bðnÞbðlMT  lta Þdp ; ð53Þ


dR ¼ bðnÞð1  nÞðR1 ðnÞ  RÞdp; ð54Þ

The term b(n) = 1  n has been added in order to compensate for the assumption that the
martensite inclusions are elastic, whereas they behave in elastic–plastic way. The process of
phase transformation-induced hardening stops when n = 1.
The above-presented mixed hardening model has the following features:

(1) principal constituents of the two-phase material model are the elastic–plastic matrix
(austenite) and the elastic inclusions (martensite platelets),
(2) the evolution of back stress is driven by the interaction between the dislocations and
the martensite sites dispersed in the austenitic matrix,
(3) the evolution of back stress depends on the properties and amount of both fractions
via the Mori–Tanaka homogenization,
1250 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

(4) for the plastically active processes the Mori–Tanaka homogenization is performed
by using the tangent stiffnesses obtained via the linearization in the vicinity of the
current state,
(5) the evolution of the isotropic hardening parameter tends asymptotically to the ulti-
mate size of the yield surface,
(6) the ultimate size of yield surface depends on the properties and amount of both frac-
tions via the Mori–Tanaka homogenization.
(7) proportions between the kinematic and the isotropic hardening are defined by the
Bauschinger parameter.

2.2. Algorithm of elastic–plastic analysis including the c ! a 0 phase transformation

Generally, the plastic straining phenomena at low temperatures, accompanied by the


c ! a 0 phase transformation, may be analysed at four different levels:

 Atomic level: reconfiguration of atoms due to the FCC–BCC transformation.


 Microscopic level: perturbation of lattice including: a 0 inclusion, interface inclusion–lat-
tice, direct proximity of the inclusion.
 Mesoscopic level: average state of stress and strain in the representative volume element
resulting from the phase transformation. quasi-isotropic behavior.
 Macroscopic level: average state of stress and strain in the macroscopic portion of two-
phase material. Fully isotropic behavior.

The present paper is based on the mesoscopic and the macroscopic levels. It is as-
sumed that the basic brick in the material is the RVE and the macroscopic state of
stress and strain results from integration of the constitutive equations defined at the
mesoscopic level. Analysis of the plastic strain-induced phase transformation in the
austenitic stainless steels follows the algorithm presented in Fig. 4. Initially, 10 material
parameters (identified by the appropriate experiments) as well as the loading conditions
are introduced. Among the material parameters both transformation thresholds (pn, nL)
as well as the slope (A) have to be defined. All the material parameters are tempera-
ture-dependent and may evolve as a function of the temperature. As soon as the pro-
cess becomes plastically active the accumulated plastic strain (p) is computed and the
first transformation threshold (pn) is checked. When the threshold is reached the pro-
cess of phase transformation begins and the initially homogeneous material becomes
heterogeneous (two-phase). This implies the corresponding evolution of material prop-
erties (hardening), computed by using the Mori–Tanaka homogenization principle. As
the algorithm is based on the updated Lagrangian approach, the increments of the
state variables are computed and an updated state of material is found. The process
continues until the moment when the second transformation threshold is reached. It
is assumed that above the second threshold there is no further transformation and
the volume fraction of martensite (n) remains constant. The material becomes partially
transformed, as a function of the experimentally confirmed second threshold nL. In the
case of unloading, the current level of the volume fraction of martensite is frozen and
as soon as the process becomes plastically active further transformation can take place.
As the transformation is governed exclusively by the accumulated plastic strain
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1251

Fig. 4. Algorithm of plastic strain-induced phase transformation at cryogenic temperatures.

(Odqvist parameter), the process is irreversible and cannot be healed. Finally, the model
allows analysis of transient states with the temperature evolution taken into account.
It is worth pointing out that the model is valid under the assumption of small strains,
1252 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

therefore a limit on the accumulated plastic strain has been imposed. It is assumed that
a practical value which constitutes this limitation is equal to 0.2.

3. Measurements of intensity of martensitic transformation at cryogenic temperatures

3.1. Evaluation of martensitic transformation through magnetic measurements

An increase in martensite content promoted by plastic deformation can be detected by


measuring the relative magnetic permeability l. Fig. 5 (Larbalestier and King, 1973) shows
a selected reference diagram which allows estimation of the martensite content (volume
fraction of martensite). It converts the susceptibility (or permeability) into martensite con-
tent. This diagram is based on the 304L, 321 and 347 SS and is accurate for magnetic mea-
surements, performed under a field of 1000 Oe (0.1 T). Since the results for different
stainless steel grades are consistent, the same diagram is considered valid for 316L grade
as well.
However, extrapolation of this diagram is uncertain above 15% of plastic strain since
the highest experimental point has been obtained for the martensitic content of around
11%. It is worth pointing out that the range of plastic strain needed to validate the con-
stitutive model is 0–20%. Thus, the available range of measurement of the magnetic sus-
ceptibility against plastic deformation corresponds to 75% of the range for the
constitutive modelling. There is another aspect to be considered in the low permeability
range. The diagram shall be regarded as the increase of the permeability measured under
a field of 1000 Oe (l1000) with respect to the initial value that is not 1. Typical values of the
initial permeability of untransformed fully austenitic SS are of the order of 1.002–1.005.
Therefore, the value identified by means of the diagram shown in Fig. 5 has to be added
to the known initial permeability level.
The advantage of this method of estimation is to convert a value of average measured
permeability into an average value of volume fraction of a 0 martensite, while avoiding del-
icate assessment of volume fraction by local methods (Magnetic Force Microscopy,
microhardness testing etc.). Indeed, results issued from local methods might be strongly
dispersed, depending on the size of the evaluated region.

Fig. 5. Permeability versus volume fraction of a 0 martensite for grades 304L, 321 and 347, strained at 4.2 K.
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1253

In order to study the properties and stability of 316L at cryogenic temperatures, fine
gauge cold rolled sheets were manufactured. The thickness of sheets under consideration
ranges from 0.15 to 0.25 mm. On these sheets, a series of measurements aimed at quanti-
fying the phase transformation at low temperatures has been carried out.

3.1.1. Chemical composition


Samples issued from two different heats of 316L have been studied, corresponding to
two different thicknesses (0.15 and 0.25 mm). The chemical composition is reported in Ta-
ble 1. The composition is compared to the allowed range. Impurities, such as P and S, can
lead to hot cracking during welding. In the analyzed material, P and S content are slightly
higher than expected (their sum is 0.0229%). The Ni content for 0.25-mm thick plates is
11.03%. Ni improves the stability of austenite versus martensitic transformations under
thermo-mechanical cycling. Mo content is smaller than expected (2.032% and 1.97%). It
is closer to the lowest specified limit of 2% (Table 1). This constituent increases the stabil-
ity of stainless steel with respect to strain-induced martensitic transformation, as well as
the resistance against corrosion.
Two temperatures of martensite formation are well defined and identified in the litera-
ture (Reed, 1983). Ms defines the temperature at which spontaneous martensitic transfor-
mation occurs during cooling to low temperatures and Md30 is the temperature at which
50% of the martensitic transformation occurs under 30% plastic strain. Semi-empirical
relationships, assessed by Reed (1983), describe the effect of the chemical composition
of the steel on Ms and Md30. The temperatures Ms and Md30 (Table 2) were calculated
by using the empirical formulas, based on the chemical composition.
The calculated values of Ms are below 0 K which has no physical meaning apart from
the conclusion that the material is stable with respect to the spontaneous transformation
down to 0 K.

3.1.2. Tensile test


Tensile tests are aimed at understanding the behavior of austenite/martensite two-phase
composite induced by plastic deformation in 316L SS. It is important to identify the point
where martensitic transformation begins. For different strain levels: 5%, 10%, 20%, 30%
and 40%, the rate of martensitic transformation was assessed for three different tempera-
tures: room temperature (RT), 77 and 4.2 K. At each level of deformation, samples were
removed from the nitrogen or helium bath in order to perform the measurement of the
magnetic permeability and to deduce the amount of martensite formed in the sample. Ten-
sile testing was performed on flat specimens (section 0.45 or 0.75 mm2) by using a universal
electromechanical testing machine, UTS 200.4 (load cell 25 kN). Two different extensom-
eters were used: an inductive sensor for the elastic range (Sensorex SX 9W3) and a carbon
potentiometer for the plastic range (MCB RX 13-50). The tests were kinematically con-
trolled (total strain).

Table 2
Temperatures Ms and Md30 estimated for two different 316L heats
Thickness (mm) Ms (K) Md30 (K)
0.15 <0 265
0.25 <0 266
1254 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

Fig. 6. Partially transformed austenite of an AISI 316L austenitic stainless steel sample strained at 6.5% at 4.2 K.

Martensite is concentrated in bands (under the white boundary in Fig. 6(A)), develop-
ing during tensile deformation at 4.2 K. A detail of the austenite–martensite microstruc-
ture is shown in Fig. 6(B).

3.1.3. Measurements of the magnetic permeability and deduction of volume fraction of


martensite
The volume fraction of martensite is quantified by magnetic methods (Trémolet de Lac-
heisserie, 2000). The set-up, used for magnetization measurements, consists of a magnetic
superconducting coil. The system is equipped with temperature stabilization and relevant
instrumentation to measure the magnetization (M) and the magnetic field (H). Measuring
the magnetic flux induced by the sample inside the measuring coil enables the quantifica-
tion of the magnetization. The flux is calculated on the basis of the reciprocity theorem
(Trémolet de Lacheisserie, 2000). Several techniques of assessment of the volume fraction
of martensite in stainless steels are proposed in the literature. Measurements of magnetic
flux density using a thin film Flux Gate sensor (FG) have been advantageously applied to
assess material degradations in Inconel 600 and 304-stainless steel (Takaya and Miya,
2005). FG does not need any cooling system and has higher spatial resolution than a
Superconducting QUantum Interference Device (SQUID). SQUID is a second well as-
sessed technique to measure magnetic properties, recently used (Zhang et al., 2004) to as-
sess the degradation of 304 type stainless steels used in nuclear reactors due to radiation
induced segregation, possibly resulting in martensitic transformations. For our study, two
different methods have been used to measure the magnetic flux: the first method called the
VSM (Vibrating Sample Magnetometer), for which a periodic movement is applied to the
sample, and the second method with an extraction magnetometer in which the sample is
subjected to a monotonic movement. Table 3 shows the features of the relevant set-ups.

Table 3
Features of the relevant devices for measuring the magnetic field
Sensitivity VSM 108 to 1010 (A m2) Extraction 107 (A m2)
Range of magnetic field 2 T to +2 T 10.8 T to +10.8 T
Temperature range 1.5–300 K 1.5–300 K
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1255

Susceptibility of the sample is determined by a graphical method (method of secants)


and represents the slope of the following curves:
l0 M
v¼ and l ¼ 1 þ v; ð55Þ
l0 H
where v is the relative susceptibility, l0 the vacuum permeability, M the magnetization, H
the magnetic field. It is worth pointing out that the non-deformed material has already a
susceptibility different from zero. The increase of susceptibility is calculated by subtracting
this value from the graphical result. Besides, each value has its own error margin which
needs to be taken into account.

3.2. Experimental results: magnetic susceptibility versus plastic strain

The results of measurements of the susceptibility versus plastic strain are given in Figs.
7–10. Fig. 7 shows the susceptibility measured for different strain levels at three different

Fig. 7. Magnetic susceptibility measured at RT versus strain applied at RT, 77 and 4.2 K for the sheet 0.25 mm
thick.

Fig. 8. Martensite content measured at RT as a function of strain applied at RT for the sheet 0.25 mm thick.
1256 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

Fig. 9. Martensite content measured at RT as a function of strain applied at 77 K for the sheets 0.25 mm thick.

Fig. 10. Martensite content measured at RT as a function of strain applied at 4.2 K for the sheets 0.15 mm and
0.25 mm thick.

temperatures (RT, 77 and 4.2 K). Figs. 8–10 show the martensitic content identified at
these temperatures, respectively.

4. Identification of parameters of the constitutive model

The constitutive model has been implemented by using a FE code. The radial return
algorithm was applied when integrating the constitutive equations for an active plastic
process (Ortiz and Simo, 1986). The model and the algorithm were initially validated by
using the data available in the literature: Morris et al. (1992); Iwamoto et al. (1998).
The first validation was carried out with respect to the experimental data obtained by
Morris et al. (1992) for grade 304L stainless steel tested at 77 K. The following set of data
needed in the model was identified (Table 4).
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1257

Table 4
Identification of the parameters for grade 304L stainless steel (from Morris et al., 1992)
E (GPa) m ry (MPa) C0 (MPa) h A pn Dv
190 0.3 580 750 1.9 4.23 0.004 0.05

Based on this set of data, the numerical results were compared with the experimental
curves presented by Morris et al. (1992) for the samples subjected to tensile monotonic
loading. In Fig. 11, both the volume fraction of martensite and the stress–strain simulation
are plotted against the experimental curves. The convergence between the experimental
and the numerical results turns out to be good at least for the strains in the range 0–0.2.
Similar validation has been done for the grade 304 stainless steel tested at 128 K (exper-
imental data published by Iwamoto et al., 1998). As the tests were carried out both under
tension and under compression two sets of parameters were identified (Table 5).
Also in this case, the model generated good results (for tension and for compression) as
shown in Fig. 12. It is worth pointing out, that in both validation cases the same value of
Baushinger parameter equal to 0.45 was applied (estimated from Suzuki et al., 1988).
Generally, 10 parameters are needed as input to the model (Fig. 4), among which three
are linked to the kinetics of the phase transformation. However, for the engineering pur-
poses a simplified model including only two parameters in the kinetics of phase transfor-
mation (pn, A) can effectively be used. Thus, the number of parameters to be identified

Fig. 11. Validation of the model for grade 304L stainless steel under tensile monotonic loading (data from Morris
et al., 1992).

Table 5
Identification of the parameters for grade 304 stainless steel (from Iwamoto et al., 1998)
E (GPa) m ry (MPa) C0 (MPa) h A pn Dv
Tens. Comp. Tens. Comp.
190 0.3 600 1200 1.8 6.3 6.3 0.028 0.005 0.05
1258 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

Fig. 12. Validation of the model for grade 304 stainless steel under tensile and compressive loads (data from
Iwamoto et al., 1998).

reduces to 9. Another problem is related to the thickness of the samples. In the case of fine
gauge samples, where the skin layers play an important role, only tensile tests are possible
(for stability reasons). Nevertheless, identification of the parameters of the model under
tensile loads provides already a very useful information that can be exploited in the struc-
tural analysis.

4.1. Identification of the parameters of the kinetic law of strain-induced martensitic


transformation

The parameters have been identified from the curve of martensite content versus the
inelastic strain. The inelastic strain is decomposed into plastic strain and the bain strain
according to the relationship
inel ¼ p þ PT . ð56Þ
The volume fraction of martensite is plotted as a function of plastic strain. Linearization of
the region II of the sigmoidal curve is carried out by using the least-square method. The slope
of the line and the intersection with the x-axis correspond to the parameter A and threshold
pn, respectively. As an example, identification of the parameters for grade 316L stainless steel
at 77 K is shown in Fig. 13. For the example shown in Fig. 13, the parameters pn and A mea-
sured at 77 K are equal to 0.09 and 4.37, respectively. The identification has been carried out
for three temperatures (RT, 77, 4.2 K) and the corresponding values are presented in Table 6.
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1259

Fig. 13. Identification of parameters of the kinetic law of martensitic transformation at 77 K.

The evolution of two parameters describing the plastic strain-induced martensitic transfor-
mation, pn and A is given in Fig. 14.

4.2. Influence of the strain-induced martensitic transformation on the tensile curve

The c–a 0 phase transformation at cryogenic temperatures induces a significant strain


hardening, that can be measured experimentally and computed. Fig. 15 shows an example
of an experimental tensile curve measured at 77 K compared with a curve computed from
the previously identified set of material parameters, presented in Table 7 (for the same
temperature). The stress measured during the experiment corresponds to the second
Piola–Kirchhoff stress tensor. The stress, which is obtained form the numerical simulation,
corresponds to the Cauchy stress tensor. It was transformed to the Piola–Kirchhoff stress
when constructing Fig. 15. Parameter C0 has been identified in the range, where the mar-
tensitic transformation is low (p < 0.1). Parameter h has been identified as the best fit be-
tween experiment and numerical simulation.
Assuming linear interpolations for all material parameters between three temperature
levels (4.2 K, 77 K, RT), stress–strain and martensite content curves have been plotted
for isothermal processes (Fig. 16).

Table 6
Parameters of the kinetic law of martensitic transformation (0.25 mm thick SS sheet)
Temperature pn A
RT 0.39 0.023
77 K 0.09 4.37
4.2 K 0.08 5.1
1260 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

Fig. 14. Plastic strain threshold (pn) and intensity parameter (A) as a function of temperature.

Fig. 15. Simulated tensile curve for 316L stainless steel compared to an experimental curve measured at the same
temperature (77 K).

Table 7
Set of material parameters identified for 316L stainless steel at 77 K (thickness 0.25 mm)
E (GPa) m ry (MPa) C0 (MPa) h A pn Dv
206 0.3 630 1680 2.7 4.37 0.09 0.05
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1261

1200 0.7

Martensite content
1000
0.6
Stress [MPa]

0.5
800
0.4
600
0.3
400 0.2
200 0.1
0 0
0 0.25 0 0.25
50 0.2 50 0.2
Tem
100 0.15 Tem100 0.15
p erat
150 per 150 n
0.1
ain atur 0.1 ai
ure 200
[K] 250 0.05 Str e 200
[K] 250 0.05 Str
300 0 300 0

Fig. 16. Evolution of stress and martensite content as a function of strain for isothermal processes.

5. Discussion

The present paper is aimed at showing the way to measure the evolution of martensite
content in fine gauge stainless steel sheets subjected to plastic strains. Moreover, an elasto-
plastic constitutive model containing ten parameters easily identified, is proposed for cryo-
genic applications. The special AISI 316L SS, chosen as the material for hydroforming the
cryogenic bellows expansion joints, shows good stability against spontaneous transforma-
tion. For both heats and the corresponding thicknesses (0.15, 0.25 mm), there is no spon-
taneous transformation. On the other hand, as the Md30 is of 265–266 K, the steel
undergoes the plastic strain-induced martensitic transformation. The amount of martens-
ite in the austenitic matrix (volume fraction of martensite) can efficiently be quantified by
using magnetic methods. Both the vibrating sample magnetometer and the extraction
magnetometer are suitable to perform the measurements with a good precision (108
and 5 · 107 A m2). The VSM method shows a better sensitivity. However, the range of
magnetic field is much smaller when compared to the extraction method. It is worth point-
ing out, that it is rather the magnetic susceptibility that is measured and then the magnetic
permeability is calculated. The permeability is directly linked to the volume fraction of
martensite by a linear relationship shown in Fig. 5. The precision of evaluation of the vol-
ume fraction of martensite is estimated to 2%. As expected, the martensite content corre-
sponding to plastic straining at RT does not exceed some 4–4.5%, under total strain
reaching around 70%. This process is much more intensive at 77 K where the maximum
martensite content reached 15% for a total strain not exceeding 10%. Finally, at 4.2 K a
maximum of 20% volume fraction of martensite was estimated for a total deformation
of 8%. This indicates clearly that the process of plastic strain induced martensitic transfor-
mation intensifies when the temperature falls to liquid helium temperatures. In general, the
volume fraction of martensite does not exceed 25% for a total strain staying below 10%.
The curves of martensite content against total strain, serve as a basis for identification of
parameters of the constitutive model. The parameters were identified in the framework of
the kinetic law of evolution of the martensite content (Eq. (12)): pn and A. The third
important parameter nL was not identified since the volume fraction of martensite did
not reach its saturation level, even for a total strain exceeding some 40%. However, this
parameter is less important since the validity of the constitutive model is limited to around
20% of plastic strain. It is worth pointing out, that the threshold drops from 0.4 at room
1262 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

temperature to below 0.1 at 77 K and at 4.2 K. Thus, the material is more susceptible to
plastic strain induced phase transformation at cold. The intensity parameter A increases
from 0 at RT (plastic strain range limited to 20%) to 4 at 77 K and 5 at 4.2 K. This again
reflects an increasing intensity of the martensitic transformation at cryogenic tempera-
tures. For the considered grade 316L, the transformation intensity at 4.2 K is higher than
at 77 K. The other parameters of the constitutive model, like ry, C0 and h were identified
by using the tensile curves at each temperature level. The hardening parameter of the
austenitic c phase is identified at the beginning of the plastic flow, when the martensite
content is still negligible and does not affect the strain hardening process. The parameter
h reflects an increase in kinematic hardening corresponding to the restricted mobility of
dislocations in the c phase due to the presence of a 0 martensite sites (creating local stress
fields, associated with the bain strain). This parameter, identified as the best fit to the
experimentally obtained tensile curves, is equal to 2.7 at 77 K. Thus, for the martensite
content equal to 0.25 the increase in kinematic hardening modulus (Eq. (29)) amounts
to 0.68 (68% more with respect to the pure austenite phase). It is worth mentioning that
the convergence between the numerical simulation and the experimental results is very
good, even beyond 20% of plastic strain. This tends to confirm validity of the constitutive
model. Clearly, the model needs further validation under a three dimensional stress state
and under monotonic or cyclic loads. Nevertheless, it can easily be applied to the analysis
of a two-dimensional stress state in thin-walled, axisymmetric structures like cryogenic bel-
lows expansion joints. It allows the evolution of stiffness of a bellows and the increase in
magnetic permeability at the root and crest of its convolutions, resulting from local mar-
tensitic transformation, to be predicted with a good precision.

6. Conclusions

The results of magnetic susceptibility measurements performed on a 316L stainless


steel, transformed at 77 and 4.2 K from nearly pure austenitic c phase to a two-phase
material consisting of a c matrix and a 0 martensite lathes, lead to the following
conclusions:

 The grade is perfectly stable against spontaneous martensitic transformation due to


cool down.
 The strain induced martensitic transformation (Md30 = 265–266 K) occurs already at
room temperature, however for large strains exceeding 40%.
 The process of strain induced martensitic transformation is much more intensive at
cryogenic temperatures with the threshold pn of around 5–10% (4.2 and 77 K) and
the transformation intensity parameter A of around 4–5 (77 and 4.2 K). The martensite
content in fine gauge stainless steel sheets (0.15 and 0.25 mm) does not exceed 25% for a
plastic strain smaller than 10%.

Identification of the parameters of the constitutive model highlights the increase in


kinematic hardening parameter C(n) of around 68% for 25% of martensite in the austenitic
phase. The second portion of kinematic hardening G(n), based on the Mori–Tanaka
homogenization, amounts to around 44%. Thus local tangent modulus corresponding
to the kinematic hardening and to 25% volume fraction of martensite equals to
3540 MPa and is much larger than the initial tangent modulus, for the pure c austenite
C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264 1263

equal to 1680 MPa. The constitutive model offers a good prediction of stress-strain behav-
iour, when implemented numerically. The calculated tensile curve corresponds well to the
experimental curve, even beyond 20% of plastic strain. The model (formulated as a general
three-dimensional model) can easily be implemented to compute the response to plastic
strain of thin-walled shells (for example bellows expansion joints) at cryogenic tempe-
ratures.

References

Angel, T., 1954. Formation of martensite in austenitic stainless steel. J. Iron Steel Inst. 138, 165–174.
Baudry, G., Pineau, A., 1977. Influence of strain induced martensitic transformation on the low-cycle fatigue
behavior of a stainless steel. Mater. Sci. Eng. 28 (2), 229–242.
Bayerlein, M., Mughrabi, H., Kesten, M., Meier, B., 1992. Improvement of the strength of a metastable austenitic
stainless steel by cyclic deformation-induced martensitic transformation at 103 K. Mater. Sci. Eng. 159 (1),
35–41.
Chen, D.L., Wang, Z.G., Jiang, X.X., Ai, S.H., Shih, C.H., 1989. The dependence of near-threshold fatigue crack
growth on microstructure and environment in dual-phase steels. Mater. Sci. Eng. A 108, 141–151.
Cherkaoui, M., Berveiller, M., Lemoine, X., 2000. Couplings between plasticity and martensitic phase
transformation: overall behavior of polycrystalline TRIP steels. Int. J. Plast. 16, 1215–1241.
Dachowski, S., Boehm, M., 2004. Finite thermoplasticity with phase changes based on isomorphisms. Int. J.
Plast. 20 (2), 323–334.
Diani, J.M., Parks, D.M., 1998. Effects of strain state on the kinetics of strain induced martensite in steels. J.
Mech. Phys. Solids 46 (9), 1613–1635.
Fischer, F.D., 1997. Modeling and simulation. In: Berveiller, M., Fischer, F.D. (Eds.), Mechanics of Solids with
Phase Changes, No. 368. Springer, Wien, New York, pp. 189–237.
Fisher, F.D., Oberaigner, E.R., Tanaka, K., Nishimura, F., 1998. Constitutive model for transformation
plasticity accompanying strain-induced martensitic transformations in metastable austenitic steels. Acta
Metall. Mater. 40 (7), 1703–1716.
Fisher, F.D., Reisner, G., Werner, E., Tanaka, K., Cailletaud, G., Antretter, T., 2000. A new view on
transformation induced plasticity (TRIP). Int. J. Plast. 16, 723–748.
Garion, C., Skoczeń, B., 2001. Plastic strain induced damage evolution and martensitic transformation in ductile
materials at cryogenic temperatures. In: Balachandran, B., Gibser, D., Hartwig, T. (Eds.), Advances in
Cryogenic Engineering, Proceedings of the International Cryogenic Materials Conference-ICMC, Madison,
2001, vol. 48A. American Institute of Physics, New York, pp. 170–177.
Garion, C., Skoczeń, B., 2002. Modelling of plastic strain induced martensitic transformation for cryogenic
applications. J. Appl. Mech. 69 (6), 755–762.
Garion, C., Skoczeń, B., 2003. Combined model of strain induced phase transformation and orthotropic damage
in ductile materials at cryogenic temperatures. Int. J. Damage Mech. 12 (4), 331–356.
Grosse, M., Niffenegger, M., Kalkhof, D., 2001. Monitoring of low-cycle fatigue degradation in x6crniti18-10
austenitic steel. J. Nucl. Mater. 296 (1-3), 305–311.
Hecker, S.S., Stout, M.G., Staudhammer, K.P., Smith, J.L., 1982. Effects of strain state and strain rate on
deformation-induced transformation in 304 stainless steel: Part I. Magnetic measurements and mechanical
behavior. Metall. Trans. A 13A, 619–626.
Heung Nam, H., Chang Gil, L., Chang-Seok, O., Sung-Joon, K., 2004. A model for deformation behavior and
mechanically induced martensitic transformation of metastable austenitic steel. Acta Mater. 52 (17), 5203–5214.
Iwamoto, T., 2004. Multiscale computational simulation of deformation behavior of trip steel with growth of
martensitic particles in unit cell by asymptotic homogenization method. Int. J. Plast. 20 (4–5), 841–869.
Iwamoto, T., Tsuta, T., 2002. Computational simulation on deformation behavior of ct specimens of trip steel
under mode i loading for evaluation of fracture toughness. Int. J. Plast. 18 (11), 1583–1606.
Iwamoto, T., Tsuta, T., Tomita, Y., 1998. Investigation on deformation mode dependence of strain-induced
martensitic transformation in TRIP steels and modelling of transformation kinetics. Int. J. Mech. Sci. 40 (2-3),
173–182.
Larbalestier, D.C., King, H.W., 1973. Austenitic stainless steels at cryogenic temperatures, structural stability and
magnetic properties. Cryogenics 13, 160–167.
1264 C. Garion et al. / International Journal of Plasticity 22 (2006) 1234–1264

Lebedev, A.A., Kosarchuk, V.V., 2000. Influence of phase transformation on the mechanical properties of
austenitic stainless steels. Int. J. Plast. 16, 749–767.
Levitas, V.I., Idesman, A.V., Olson, G.B., 1992. Condinuum modeling of strain-induced martensitic transfor-
mation at shear band intersections. Acta Mater. 47 (1), 219–233.
Mori, T., Tanaka, K., 1973. Average stress in matrix and average energy of materials with mistfitting inclusions.
Acta Metall. 21, 571–574.
Morris, J.W., Chan, J.W., Mei, Z., 1992. The influence of deformation-induced martensite on the cryogenic
behavior of 300-series stainless steels. Cryogenics ICMC (Suppl. 32), pp. 78–85.
Narutani, T., Olson, G.B., Cohen, M., 1982. Constituitve flow relations for austenitic steels during strain-induced
martensitic transformation. J. Phys. 43 (12), 429–434.
Oberste-Brandenburg, C., Bruhns, O.T., 2004. A tensorial description of the transformation kinetics of the
martensitic phase transformation. Int. J. Plast. 20, 2083–2109.
Olson, G.B., Cohen, M., 1975. Kinetics of strain-induced martensitic nucleation. Metall. Trans. A (6), 791–795.
Ortiz, M., Simo, J.C., 1986. An analysis of a new class of integration algorithms for elastoplastic constitutive
equations. Int. J. Numer. Meth. Eng. 23, 353–366.
Reed, R.P., 1983. Material at Low Temperatures. American Society for Testing and Materials.
Skoczeń, B., 2004. Compensation Systems for Low Temperature Applications. Springer, Berlin, Heidelberg, New
York.
Stringfellow, R.G., Parks, D.M., Olson, G.B., 1992. Constitutive model for tranformation plasticity accompa-
nying strain-induced martensitic transformations in metastable austenitic steels. Acta Metall. 40 (7), 1703–
1716.
Suzuki, K., Fukakura, J., Kashiwaya, H., 1988. Cryogenic Fatigue Properties of 304L and 316L Stainless Steels
Compared to Mechanical Strength and Increasing Magnetic Permeability. American Society for Testing and
Materials, pp. 190–197.
Takaya, S., Miya, K., 2005. Application of magnetic phenomena to analysis of stress corrosion cracking in
welded part of stainless steel. J. Mater. Process. Technol. 161, 66–74.
Tomita, Y., Iwamoto, T., 1995. Constitutive modeling of TRIP steel and its application to the improvement of
mechanical properties. Int. J. Mech. Sci. 37, 1295–1305.
Tomita, Y., Iwamoto, T., 2001. Computational prediction of deformation behavior of TRIP steels under cyclic
loading. Int. J. Mech. Sci. 43, 2017–2034.
Tomita, Y., Shibutani, Y., 2000. Estimation of deformation behavior of TRIP steels – smooth/ringed-notched
specimens under monotonic and cyclic loading. Int. J. Plast. 16, 769–789.
Trémolet de Lacheisserie, E., 2000. Magnétisme 1: Fondements. EDP Sciences.
Watanabe, Y., Sakai, H., 2003. Control of magnetic gradient in magnetically graded materials fabricated by
martensitic transformation technique. Mater. Sci. Forum 423-425, 435–440.
Wechsler, M.S., Lieberman, D.S., Read, T.A., 1953. On the theory of the formation of martensite. AIME Trans.
J. Metals 197, 1503–1515.
Zhang, L., Kamada, Y., Kikuchi, H., Mumtaz, K., Ara, K., Takahashi, S., Sato, M., Tsukada, T., 2004.
Magnetic transition temperatures of some model alloys for simulating radiation induced segregation in
austenitic stainless steel. J. Magnet. Magnet. Mater. 271, 402–408.

You might also like