Sie sind auf Seite 1von 9

Available online at www.sciencedirect.

com
Procedia
Engineering
Procedia Engineering 00 (2009) 000000
Procedia Engineering 2 (2010) 22292237
www.elsevier.com/locate/procedia
www.elsevier.com/locate/procedia

Fatigue 2010

Strain heterogeneities between phases in a duplex stainless steel.


*
Comparison between measures and simulation.
A. El Bartalia,b,c, P. Evrarda,b,c, V. Aubina,b,c, S. Herend
I. Alvarez-Armasd, A.F. Armas and S. Degallaix-Moreuila,b,c,*
a
Univ Lille Nord de France, F-59000 Lille, France
b
ECLille, LML, F-59650 Villeneuve d'Ascq, France
c
CNRS, UMR 8107, F-59650 Villeneuve d'Ascq, France
d
Instituto de Fsica Rosario, CONICET - Universidad Nacional de Rosario,
Bv. 27 de Febrero 210 bis, 2000 Rosario, Argentina

Received 19 February 2010; revised 11 March 2010; accepted 15 March 2010

Abstract

Many studies aimed at understanding the strain and damage micro-mechanisms in low-cycle fatigue in duplex stainless steels.
Different methods allow the evaluation of the strains on the microstructural scale, some are surface techniques (AFM, EBSD,
DIC) and others are bulk techniques (TEM, XRD). Recently, some works have analyzed the plastic strain sharing between
austenitic and ferritic phases using one or several of these experimental methods (TEM, XRD, AFM, EBSD, etc.). The present
work proposes to analyze the strain fields measured by digital image correlation (DIC) on the microstructural scale at the surface
of a specimen cyclically strained in low-cycle fatigue, and to compare it with that obtained by a polycrystalline bi-phased
microstructure numerical calculation.
A cyclic mechanical test was carried out at room temperature on a forged duplex stainless steel. Before the test, a surface
crystallographic orientation measurement was performed by EBSD analysis in the central part of the specimen. A specific in-situ
optical microscopic device equipped with a CCD camera was used in order to observe and take digital images of the specimen
surface on the microstructural scale, at different number of cycles. The displacement and strain fields of a representative surface
zone were obtained by DIC using the specific software CORRELIQ4. The strain distribution and the strain sharing between
austenitic and ferritic phases were analyzed from mechanical fields obtained between images taken at different number of cycles.
In parallel, a quasi-2D microstructure finite element calculation of a small zone experimentally analyzed was performed.
Numerical microstructure and grain orientations were obtained from EBSD measures. Two crystal plasticity constitutive laws
were used to model the behaviors of austenitic (FCC) and ferritic (BCC) grains, respectively. The mechanical fields and the strain
distributions per phase were compared to those obtained by DIC on the first hysteresis loop.
Keywords: Duplex stainless steel; Cyclic elastic-plastic behaviour; EBSD analysis; Digital image correlation; Kinematic fields; Microstructure
numerical simulation.

c 2010 Published by Elsevier Ltd.

*
* Suzanne Degallaix-Moreuil. Tel.: 00 33 3 29 33 53 77; fax: 00 33 3 20 33 54 99.
E-mail address: suzanne.degallaix@ec-lille.fr

1877-7058 c 2010 Published by Elsevier Ltd.


doi:10.1016/j.proeng.2010.03.239
2230 A. El Bartali et al. / Procedia Engineering 2 (2010) 22292237

2 E. Bartali et al./ Procedia Engineering 00 (2010) 000000

1. Introduction

Duplex stainless steels (DSSs) are austenite/ferrite two-phased metallic alloys developed and progressively
improved since several decades. The DSSs forged contain usually approximately equal proportions of ferrite (BCC)
and austenite (FCC), i.e. from 60/40 to 40/60 percent volume fractions of austenite/ferrite. DSSs combine the good
mechanical properties of both austenitic and ferritic phases (high YS, UTS, work-hardening rate and ductility), some
of them (YS, UTS) still increased by the small grain size, and progressively improved during recent years thanks to
an increasing addition of nitrogen [1-6]. Moreover, DSSs possess a very good corrosion resistance [1-4], also
improved thanks to nitrogen alloying [2, 4]. Today, they are used as structural materials in large fields of industries,
such as oil and gas, petrochemical, paper, and nuclear industries, also replacing even progressively the more costly
300 series austenitic stainless steels. Finally, their high mechanical properties permit thickness reductions,
particularly appreciated in transportation industries for instance.
In low-cycle fatigue (LCF), plastic deformations take place progressively during cycling, damage at room
temperature proceeds by appearance of slip bands at the surface, changing progressively in intrusions/extrusions,
and then microcracks initiate in the intrusions and then propagate up to fracture (see in particular the review recently
published [7]). Given the bi-phased character of the DSSs, there are strain incompatibilities between austenitic and
ferritic phases, due to their different crystallographic structures, morphologies, yield stresses and other mechanical
properties. In addition, internal stresses develop during elaboration process and can change, even disappear, during
further mechanical loading [19]. Up to recently, many studies have aimed to understand the strain and damage
physical mechanisms during low-cycle fatigue in the duplex stainless steels (see for instance [8-20]) using various
physical analysis methods (OM, XRD, SEM, TEM, EBSD, AFM). In order to evaluate the strain distribution
between the two phases, some works analyzed the plastic strain using and eventually combining several
experimental methods such as [10, 16, 17, 19- 22]. If TEM is a bulk analysis technique, the others are surface, thus
non-destructive analysis techniques, and some of them can be used in situ during cycling, using specific equipments.
Lillbacka et al. [23] studied experimentally the influence of the composition and microhardness of each phase on the
load sharing in several DSSs during low-cycle fatigue using SEM surface observations TEM bulk observations.
Moreover, they used a multi-scale material modeling (i.e. polycrystalline bi-phased modeling), to evaluate the load
sharing between both phases. They show that the model describes relatively well the actual evolution of the stress-
strain behavior of individual phases. The work presented in this paper has two objectives. Firstly, in order to meet
experimentally knowledge of cyclic plastic strain sharing between the two phases of a DSS, surface displacement
and strain field measures by digital image correlation during cycling, and surface damage observations by SEM after
cycling, both on the grain scale, were performed. Results obtained for a tension-compression LCF test at t/2 = +/-
0.5% [22, 24] are presented here. Secondly, a previous work [25, 26] developed a quasi-2D microstructure
numerical calculation in order to evaluate the strains in grains of each phase of the same DSS, after a tension up to
t = + 0.5% and a tension up to + 0.5% followed by a compression up to t = - 0.5%. The results of the first study are
presented here and compared with those of the second one.

2. Experimental procedure

The material studied is a X2 Cr Ni Mo N 25-07 duplex stainless steel with the chemical composition (in wt %)
C:0.024, Cr:24.7, Ni:6.54, Mo:2.84, Mn:0.79, Si:0.62, N:0.17. It was forged as a 70 mm diameter bar. It was solid
solution treated for one hour at 1060C and then water quenched. Typical micrographs are shown in Fig. 1. The
volume fractions (in percents) were 40/60 and the mean grain sizes were 10 m and 30 m for austenite and ferrite,
respectively. Despite the morphological texture, none crystallographic texture was measured by X-ray diffraction.
A LCF test was performed at room temperature under total strain control, under fully reversed
tension/compression with a triangular form signal at the constant strain rate 10-3 s-1 and the macroscopic total strain
amplitude t/2=+/-5x10-3. The test specimen was taken in the longitudinal direction of the bar. It was cylindrical
(=10 mm, l0=32 mm) and button-headed, with a slight notch in the central part, where the further observations
would be made during the mechanical test. A specific in-situ optical microscopic device equipped with a CCD
camera was used in order to observe and take digital images of the specimen surface on the microstructural scale at
different times during the test, corresponding to the zero macro-strain on the stress-strain hysteresis loops. The
radius of the notch (radius = 40 mm, depth = 0.4 mm) was sufficiently large to slightly concentrate the deformation
A. El Bartali et al. / Procedia Engineering 2 (2010) 22292237 2231

E. Bartali et al./ Procedia Engineering 00 (2010) 000000 3

in the zone observed (Kt = 1.14), and to be compatible with the low field depth of the optical microscopic device.
The specimen geometry and the experimental equipment are shown in Fig. 2. Before the mechanical test, a
crystallographic orientation measurement was performed by EBSD analysis in the central part of the notch. During
the test, 4x6 images covering about 500x400 m2 were taken, each image covering a surface of 120x90 m2,
corresponding to 1368x1024 pixels. Five zones (respectively noted A, B, C, D, E) corresponding to five images
among the 24 ones were more precisely studied by digital image correlation (DIC). The micro-displacement fields
of these representative zones were measured by DIC and the strain fields were calculated by derivation of the
displacement fields measured, using the specific software CORRELIQ4 [27]. These fields were obtained from
couples of images taken at two successive times during the test (images taken at cycles 0, , , 1, 4, 20, 50, 100,
200, 350, 500, 750, 1000). For the spatial resolution of 32 pixels, uncertainties calculated in terms of displacements
and strains were 0.02 pixel and 8x10-4, respectively [22]. The surface grey level texture obtained by a slight
chemical etching was used for DIC measurements. Except for the first cycle, the images were always taken at zero
macro-strain, i.e. at the end of one strain cycle. The additional strains (t) were obtained by DIC between two
successive images. In order to obtain the additional cumulated strains (comparison between two any images) and
the cumulated strains (comparison between the images at cycle n and cycle 0), additional strains must be added. It
must be noted that the cumulated strains defined here are different from the cumulated plastic strains usually defined
in LCF. Indeed usually in LCF, p,cumul=2xNxp, where N is the number of cycles.

(a) axial section (b) transverse section

Fig. 1. Microstructure of the forged duplex stainless steel studied: (a) longitudinal section, (b) transverse section.

(a) (b) (c)

Fig. 2. Geometries of the specimen (a) and of the notch (b), and experimental equipment (c).
2232 A. El Bartali et al. / Procedia Engineering 2 (2010) 22292237

4 E. Bartali et al./ Procedia Engineering 00 (2010) 000000

3. Numerical microstructure calculation

An internal reduced part (63x60 m2, Fig. 3) of zone B studied by DIC was chosen in order to develop a quasi-
2D microstructure finite element calculation. From SEM observations and EBSD crystallographic orientation
measurements, a numerical polycrystalline microstructure was generated at the surface and extruded in the thickness
for a one-grain-size dimension (columnar grains are thus obtained), using ABAQUS finite element software and
specific subroutines. Crystal plasticity laws were used to model the behavior of austenitic (FCC, 12 slip systems
{111} 110 ) and ferritic (BCC, 48 slip systems {110}{112} 111 ) grains, based on crystallographic slips and
dislocation densities. A linear cubic elasticity laws models the elastic behavior of each phase. More details can be
found elsewhere [25, 26]. In these works, two types of boundary conditions applied to the numerical microstructure
were tested, called, respectively, homogeneous and heterogeneous experimental boundary conditions. Firstly, the
face behind of numerical microstructure was blocked in the thickness direction. In-plane displacements measured by
DIC on the lateral faces of the zone analysed were used to define the two types of boundary conditions.
Homogeneous experimental boundary conditions were defined as the averages of displacements measured in the two
in-plane directions. Heterogeneous experimental boundary conditions were defined as the actual displacement
distributions measured in the two in-plane directions. It was shown that a better description of intraphase and
intragranular strain heterogeneities was obtained using heterogeneous, instead of homogeneous, experimental
boundary conditions. The results obtained with these boundary conditions are used here for the calculation of strain
fields by DIC.

(a) (b) (c)

Fig. 3. (a) 




4. Experimental results and microstructure calculation results

4.1. Strains in fatigue obtained by DIC

The spatial averages of the longitudinal and transversal strains on each of the five zones specifically studied were
calculated by DIC between the photographs taken at cycle 0 and cycle (t=5x10-3 in tension) and between those
taken at cycle and cycle 1 (t= -5x10-3). The results are presented in Fig. 4. These strains vary slightly from one
zone to the other, because the zones are morphologically different and too small to represent the equivalent
homogeneous medium (HEM). A standard deviation equal to 6x10-3 characterizes the spatial dispersion of
longitudinal strains in the five zones, much higher than measure uncertainties (equal to 8x10-4). The macroscopic
local total strain at the notch tip calculated by FE simulation was 6.4x10-3. The averages of the longitudinal and
transversal strains measured on the five zones are +6.12x10-3 and -2.72x10-3, respectively between cycles 0 and 1/4.
They are equal to -6.74x10-3 and +2.46x10-3, respectively between cycles 1/4 and 1. The ratios between these
A. El Bartali et al. / Procedia Engineering 2 (2010) 22292237 2233

E. Bartali et al./ Procedia Engineering 00 (2010) 000000 5

transversal and longitudinal strains are 0.444 and 0.365, respectively; they are intermediate between the elastic
macroscopic Poisson coefficient (0.3) and the plastic one (0.5).

 
 

 
  
   

^
^

 

 

 


  

 

    


 
         

(a)  11  22 (b)

Fig. 4. Spatial averages of the longitudinal (in dark) and transversal (in light) strains on the five zones studied, obtained by DIC between cycles 0
and (a) and cycles and 1 (b).

Figure 5-a gives the evolution during cycling of the additional cumulated spatial averages of the longitudinal
strains for each of the five zones. These cumulated strains increase during cycling, but the number of cycles in each
interval is also increasing. Figure 5-b gives the spatial average increments per cycle of the longitudinal strains for
each of the five zones. These increments are relatively high and scattered during the four first cycles: they stand in
the interval [-1.3x10-4; 5x10-4] per cycle. The scatter is due to the phase distributions and volume fractions, which
are very different from one zone to another one. Then, they decrease rapidly during cycling and remain scattered,
and finally stabilize after cycle 200, around 10-5 per cycle. This decreasing and this stabilization are explained by the
reversibility of the cyclic plastic straining, which increases progressively during cycling, up to stabilization, before a
macroscopic damage appears.

2,E-03 6,E-04

5,E-04
Spatial average increments per cycle

1 4 200 50 100 200 350 500 750 1000


0,E+00
Additional cumulated spatial averages

4,E-04
of the longitudinal strains.
of the longitudinal strains

3,E-04
-2,E-03
2,E-04

-4,E-03 1,E-04

0,E+00
-6,E-03 1 10 100 1000
-1,E-04

-2,E-04 Nomber of cycles


-8,E-03 Nomber of cycles

 Zone A  Zone B  Zone C  Zone D  Zone E


(a) (b)

Fig. 5. Results obtained by DIC on the five zones studied: additional cumulated spatial averages (a) and spatial average increments per cycle (b)
of the longitudinal strains.
2234 A. El Bartali et al. / Procedia Engineering 2 (2010) 22292237

6 E. Bartali et al./ Procedia Engineering 00 (2010) 000000

Then, the strain distribution between the austenitic and ferritic phases in zone B was studied on the first quarter
of cycle. As DIC software CORRELIQ4 uses a FE type mesh, each node of the mesh could be identified as
belonging to austenite or ferrite. The spatial averages of the longitudinal strains in austenite and ferrite were then
calculated by DIC between the photographs taken at cycles 0 and , i.e. for a macro-strain t=5x10-3 in tension.
These spatial average strains are given with the corresponding standard deviations in Table 1. The strain
distributions in both phases are relatively large, with standard deviations of the same order of magnitude, even
slightly higher than the strains. The strain distribution in the austenite is shifted at a clearly higher strain relative to
that in the ferrite (ratio 1.4 between the central values). That means that at the macro-strain t=5x10-3 the austenite
undergoes a higher plastic deformation than the ferrite. Indeed, the austenite has a lower yield stress than the ferrite.
Moreover, previous works [19, 20] showed that, due to the elaboration process, residual stresses exist before
mechanical testing in DSSs. These stresses are tension and compression stresses in the austenite and ferrite,
respectively, and add to the loading stresses. Therefore, the austenite deforms plastically earlier and much more than
the ferrite, during loading. Nevertheless, the ferrite participates also to the plastic deformation at this strain level, as
also observed for instance in [10, 21, 28] under cyclic and monotonous loadings, respectively. Furthermore, during
cyclic straining in LCF, the residual stresses will relax rapidly [20]. Nevertheless, during LCF at t/2=+/-5x10-3,
slip markings were observed at the surface earlier in the austenite (at cycle 5) than in the ferrite (at cycle 20) [10, 21,
22].

Table 1. Spatial averages and standard deviations of longitudinal strains in zone B, obtained by DIC between cycles 0 and .

Austenite Ferrite Total image


Spatial average 6.2x10-3 4.4x10-3 5.4x10-3
Standard deviation 7x10-3 6.3x10-3 6.5x10-3

4.2. Comparison of strain fields obtained by DIC and simulated

This analysis was performed on the internal reduced part of zone B, on which the numerical microstructure
calculation was performed (see Fig. 3). Given the calculation times, the simulation was performed for a loading
corresponding to the first cycle in fatigue. The spatial averages of the longitudinal strains in austenite, in ferrite
and in the whole reduced image, obtained by DIC and by numerical microstructure calculation at cycle , i.e. after a
tension up to t=5x10-3, and at cycle , i.e. after a tension up to t=5x10-3 followed by a compression up to t=-5x10-
3
, are given in Table 2; relative errors between experimental and simulated results are given.
The numerical microstructure simulation gives higher strains and lower standard deviations in the austenite than
in the ferrite at cycles and . On the contrary, DIC gives higher strains and lower standard deviations in the
austenite compared with the ferrite at cycle , while it is the reverse at cycle . The average strains after tension
and after tension followed by compression, obtained by DIC and by simulation, are of the same order of magnitude
(around 6x10-3 and around -10-2, respectively), while the macroscopic strains calculated by FEM are 6.33x10-3 and -
8.75x10-3, respectively. Spatial averages and standard deviations are of the same order of magnitude together, that
means that the distributions of intra-phase strains measured by DIC are relatively large. The simulation
overestimates both experimental spatial averages and standard deviations. Indeed, the various crystallographic
orientations and granular environments of the individual grains of austenite and ferrite explain the large strain
distributions obtained experimentally. Moreover, after tension followed by compression, the DSS exhibits
macroscopically a relatively high kinematic strain hardening, i.e. a strong Bauschinger effect, which is not taken into
account in the simulation by the crystal laws, as none kinematic strain hardening term is introduced in the crystal
laws, neither for the austenite, nor for the ferrite. An improvement of the FE bi-phased polycrystalline model used
for simulation would consist in adding such kinematic strain hardening terms in the crystal laws.
A. El Bartali et al. / Procedia Engineering 2 (2010) 22292237 2235

E. Bartali et al./ Procedia Engineering 00 (2010) 000000 7

Table 2. Spatial averages (SA) and standard deviations (SD) of longitudinal strains in the reduced zone obtained by DIC and by microstructure
calculation between cycles 0 and and between cycles 0 and 3/4, and relative errors (RE) between experimental and simulated result.

SA by DIC SA simul RE (%) SD by DIC SD simul RE (%)


Austenite between 6.8x10-3 6.8x10-3 <0.1 5.5x10-3 4.1x10-3 26
Ferrite cycles 5.6x10-3 6.6x10-3 14 6.0x10-3 5.3x10-3 6
-3 -3 -3
Zone 0 and 1/4 6.1x10 6.7x10 8.5 5.9x10 4.9x10-3 17
Austenite between -8.17x10-3 -1.02x10-2 22 6.0x10-3 7.0x10-3 13
Ferrite cycles -8.4x10-3 -8.7x10-3 4 5.6x10-3 7.3x10-3 23
Zone 0 and 3/4 -8.3x10-3 -9.4x10-3 12 5.8x10-3 7.2x10-3 20

Figure 6 presents the axial strain fields obtained by DIC and by numerical microstructure calculation, between
cycle 0 and cycles and . The scale ranges are common ( 2x10-2) for DIC and microstructure calculation results,
but the colors are reversed.
The strain field obtained by DIC between cycles 0 and (Fig. 6a) shows that, locally, the strains can achieve
2x10-2 in austenitic grains (white ovals) and in ferritic grains close to highly strained austenitic grains (black ovals).
A strain concentration band can be observed, tilted at about 45 relative to the loading axis. The strain field
calculated by numerical microstructure simulation between cycles 0 and (Fig. 6b) exhibits about the same strain
levels, 2x10-2, and a strain concentration band very close to that obtained by DIC. Moreover, sites with the highest
strains are very close between DIC and simulation results. Only the strain concentration zone observed on the
bottom left on simulated strain field is not observed on DIC results. As a conclusion, the intra-phase and intra-
granular strain heterogeneities agree relatively well between DIC and simulation results in terms of localization.
Moreover, the simulation underestimates the strain distribution width in each phase.

Fig. 6. Longitudinal strain fields obtained between cycle 0 and cycle (a and c) and cycle (b and d), by DIC (a and b) and by numerical
microstructure simulation (c and d).
2236 A. El Bartali et al. / Procedia Engineering 2 (2010) 22292237

8 E. Bartali et al./ Procedia Engineering 00 (2010) 000000

Strain fields obtained by DIC and by numerical microstructure simulation between cycles 0 and , in Fig. 6c and
6d, respectively, do not agree so well as after tension. Both experimental and simulated strain fields exhibit larger
zones highly strained than after tension, and several strain concentration bands, tilted at about 45 relative to the
loading axis. Nevertheless, these bands are shorter (one or two grains transversally) than those observed after
tension, their lengths and localizations differ between experimental and simulated results, and they are much more
marked on the simulated strain field. Finally, the simulation overestimates the strain distribution widths in each
phase.

5. Conclusions

A tension/compression low-cycle fatigue test was performed at room temperature on a duplex stainless steel
solid-solution treated, under fully reversed total strain control at the total strain amplitude t/2=+/-5x10-3. The
displacement and strain fields were measured by digital image correlation between images taken at different times
during cycling on the microstructural scale, using CORRELIQ4 specific software. A quasi-2D microstructure
numerical calculation was performed, of a small zone also studied by digital image correlation. The microstructure
was generated using identifying crystal orientations and grain and phase contours identified from EBSD
measurements. The crystal plasticity laws used for austenitic and ferritic grains were based on crystallographic slips
and dislocation densities.
The evolutions of the spatial average longitudinal and transversal strains and of the spatial average cumulated
longitudinal strains during cycling were studied on five surface zones of 120x90 m2. The results are the followings:
The longitudinal and transversal average strains measured by DIC on each of the five zones studied agree
relatively well with the macroscopic local strains calculated by finite element simulation and with the transversal
contraction moduli in the elactic-plastic domain.
The spatial average increments per cycle of the longitudinal strains in the five zones are scattered all along the
fatigue life, due to the microstructural configurations that are different for the five zones. They decrease during
accommodation phase (around 200 cycles), and are then stabilized at about 10-5 per cycle, showing the high
plastic strain reversibility.
The strains are higher in the austenite than in the ferrite, nevertheless, the ferrite participates significantly to the
deformation as early as the first quarter of cycle.
Due to the various crystallographic orientations and granular environments of the grains, the strain distributions
in austenitic and ferritic grains are relatively wide.
The strain fields measured by DIC and the strain fields obtained by microstructure numerical calculation at cycle
agree relatively well, in terms of strain levels as in terms of intra-granular and intra-phase heterogeneities. The
results are not so good at cycle . Probably, the microstructure numerical calculation results should be improved by
the introduction of kinematic strain hardening terms in the crystal plasticity laws.

Acknowledgements

This work was supported by the ECOS-Sud - MINCyT program, project number A06E01, of the Ministry of
National Education, Research and Technology and the Foreign Affairs Ministry of France, and the National
Ministry of Science and Technology of Argentina and the Consejo Nacional de Investigaciones Cientficas y
Tcnicas (CONICET).

References

[1] Charles J. Past, present and future of the duplex stainless steels. Int Conf & Expo, Duplex Stainless Steels2007, Grado, Italy, June 17-20.
[2] Gunn RN. Duplex stainless steels. ISBN 1-85573-318-8. 3rd ed. Abington Publishing: Cambridge England; 2003.
[3] Nilsson JO. Superduplex stainless steels. Mater Sci Technol 1992;8:685-700.
A. El Bartali et al. / Procedia Engineering 2 (2010) 22292237 2237

E. Bartali et al./ Procedia Engineering 00 (2010) 000000 9

[4] Foct J, Magnin T, Perrot P, Vogt JB. Nitrogen alloying of duplex stainless steels. Proceedings of Duplex stainless steels 91, Charles J and
Bernhardsson S Ed., Les Editions de Physiques 1991;1:49-66.
[5] Alvarez-Armas I, Degallaix-Moreuil S. Duplex stainless steels. ISBN 978-1-84821-137-7. ISTE Ltd, London, GB, and John Wiley &
Sons, Inc., Hoboken, USA;2009.
[6] Horvath W, Tabernig B, Werner E, Uggowitzer P. Microstructures and yield strength of nitrogen alloyed super duplex steels. Acta Mater
1997;45- 4:1645-1654.
[7] Man J, Obrtlik K, Polk J. Extrusions and intrusions in fatigue metals. Part 1. State of the art and history. Phil Mag 2009; 89:16:1295-
1336.
[8] Vogt J-B. Fatigue properties of high nitrogen steels. J Mater Process Technol 2001;117:364-369.
[9] Vogt J-B, Magnin T, Foct J. Fat Fract Eng Mater Struct 1993;16:555.
[10] Degallaix-Moreuil S, Seddouki A, Degallaix G, Kruml T, Polk. Fatigue damage in austenitic-ferritic stainless steels. Fat Fract Eng
Mater Struct 1995;18:65-77.
[11] Llanes L, Mateo A, Violan P, Mndez J, Anglada M. On the high cycle fatigue behavior of duplex stainless steels: Influence of thermal
aging. Mat Sci Eng A 1997;234-236:850-852.
[12] Mateo A, Llanes L, Akdut N, Anglada M. High cycle fatigue behaviour of a standard duplex stainless steel plate and bar. Mat Sci Eng A
2001; 319-321:516-520.
[13] Stolarz J, Foct J. Specific features of two-phase alloys response to cyclic deformation. Mat Sci Eng A 2001;319-321:501-505.
[14] Here S, Alvarez-Armas I, Armas AF, Girons A, Llanes L, Mateo A, Anglada M. Low cycle fatigue behaviour of duplex stainless
steels at high temperature. European Structural Intergrity Society 2002;29:15-23.
[15] Aubin V, Quaegebeur P, Degallaix-Moreuil S. Cyclic plasticity of a duplex stainless steel under non-proportional loading. Mat Sci Eng A
2004;346:1-2:208-215.
[16] Polk J, Petrenec M, Kruml T. Cyclic plastic response and fatigue life in superduplex 2507 stainless steel. Int J Fat 2010;32:279-287.
[17] Alvarez-Armas I, Marinelli MC, Malarra JA, Degallaix-Moreuil S, Armas AF. Microstructure associated with crack initiation during
low-cycle fatigue in a low nitrogen duplex stainless steels. Int J Fat 2007;29:758-764.
[18] Girons A, Villechaise P, Mateo A, Anglada M, Mndez J. EBSD studies on the influence of texture on the surface damage mechanisms
developed in cyclically loaded aged duplex stainless steels. Mat Sci Eng A 2004;387-389:516-521.
[19] Dakhlaoui R, Baczmanski A, Braham C, Wronski S, Wierzbanowski K, Oliver EC. Effect of residual stresses on individual phase
mechanical properties of austeno-ferritic duplex stainless steel. Acta Mater 2006;54:5027-5039.
[20] Jencus P, Polk J, Luks P, Murnsky O. In situ neutron difraction study of the low-cycle fatigue of the duplex stainless steel.
Physica B 2006;385-386:597-599.
[21] Salazar D. Etude du partage de la plasticit cyclique dun acier duplex par microscopie force atomique. PhD in Material Science,
Universit de Sciences et Technologies de Lille, 2008.
[22] El Bartali A. Apport des mesures de champs cinmatiques ltude des micromcanismes dendommagement en fatigue plastique dun
acier inoxydable duplex. PhD in Mechanics, Ecole Centrale de Lille, 2007.
[23] Lillbacka R, Chai G, Ekh M, Liu P, Johnson E, Runesson K. Cyclic stress-strain behavior and load sharing in duplex stainless steels:
Aspects of modeling and experiments. Acta Mater 2007, 55:5359-5368.
[24] El Bartali A, Aubin V, Degallaix-Moreuil S. Surface observation and measurement techniques to study the fatigue damage
micromechanisms in a duplex stainless steel. Int J Fat 2009;31:2049-2055.
[25] Evrard P. Modlisation polycristalline du comportement lasto-plastique dun acier inoxydable austno-ferritique. PhD in Mechanics,
Ecole Centrale de Lille, 2008.
[26] Evrard P, El Bartali A, Aubin V, Rey C, Degallaix S, Condo D. Influence of boundary conditions on bi-phased polycrystal
microstructure calculation. Accepted in Int Jnl Sol Struct 2010.
[27] Besnard G, Hild F, Roux S. Finite-element displacement fields analysisfrom digital images : Application to Portevin-Le Chatelier bands.
Experimental Mechanics 2006 ;46-6:789-804.
[28] Serre I, Salazar D, Vogt JB. Atomic force microscopy investigation of surface relief in individual phases of deformed duplex stainless
steel. Mat Sci Eng A 2008;492:428-433.

Das könnte Ihnen auch gefallen