You are on page 1of 8

International Journal of Fatigue 26 (2004) 1129–1136

www.elsevier.com/locate/ijfatigue

Technical Note
Effect of precipitates on the corrosion–fatigue crack initiation
of ISO 5832-9 stainless steel biomaterial
E.J. Giordani a,, V.A. Guimarães b, T.B. Pinto a, I. Ferreira a
a
Departamento de Engenharia de Materiais, Faculdade de Engenharia Mecânica, Universidade Estadual de Campinas, Caixa Postal 6122,
CEP 13083-970 Campinas, SP, Brazil
b
Faculdade de Engenharia e Arquitetura, Universidade de Passo Fundo, Caixa Postal 611, CEP 99001-970 Passo Fundo, RS, Brazil

Received 7 July 2003; received in revised form 21 January 2004; accepted 3 March 2004

Abstract

The correlation between the microstructure and fatigue crack initiation behavior (in air and 0.9% NaCl solution) of ISO 5832-9
austenitic stainless steel biomaterial was investigated. A combination of techniques such as extraction of precipitates, SEM, EDS
and X-ray diffraction was found to be very useful to characterize Z-phase precipitates, which are abundant in this steel. An analy-
sis of fatigue crack initiation, using SEM, revealed that fatigue and corrosion–fatigue cracking initiate preferentially in the elon-
gated Z-phase precipitates, which are parallel to the tensile direction, and in coarse precipitates usually associated with Al-rich
nonmetallic inclusions. Crack initiation is caused by particle rupture rather than by separation of the particle/matrix interface.
The findings indicate that coarse Z-phase precipitates and nonmetallic inclusions are detrimental to the fatigue properties of the
steel. These microstructural constituents accelerate the initiation of fatigue and especially corrosion–fatigue cracks.
# 2004 Elsevier Ltd. All rights reserved.

Keywords: Biomaterial; Stainless steel; Corrosion–fatigue; Crack initiation

1. Introduction in orthopedic applications, mainly when the implanted


device must remain in the human body for a relatively
Some metals and alloys that combine high strength long time (more than 12 months). The combination of
with reasonable corrosion resistance are favorite bio- such aspects favors the failure of orthopedic implants
materials for the fabrication of orthopedic implants,
by a synergy called as corrosion–fatigue. This synergy,
which are subjected to severe mechanical loading inside
which combines the electrochemical effects of corrosion
the human body. The metallic biomaterial most com-
and the mechanical effects of cyclic loading, has been
monly used for orthopedic applications today is the
austenitic stainless steel classified as ASTM F138 cited in specialized technical reports as the main cause
(a special type of 316L steel for medical applications) of fracture of metallic orthopedic implants [2–4], parti-
[1]. Its utilization is particularly justified by the combi- cularly those used in the lower extremities of the
nation of properties such as good acceptance by the human body, where implants are subjected to more
body; low cost; good machinability; good formability; severe cyclic loading. The fatigue or corrosion–fatigue
high strength, especially when cold worked, and life of a device is the time or number of cycles required
reasonable corrosion resistance. However, some aspects for cracking to initiate and to propagate until the final
such as low strength in the annealed condition and sus- fracture occurs. Fatigue life takes longer to produce a
ceptibility to localized corrosion (pit and crevice cor- crack measuring just a few micrometers than does the
rosion) often limit the wider use of this type of material propagation of that crack; thus, detailed studies of the
mechanisms which rule fatigue or corrosion–fatigue

Corresponding author. Tel.: +55-19-3788-3312; fax: +55-19-3289-
crack initiation are extremely important.
3722. Research aimed at developing new metallic materials
E-mail address: enricogiordani@uol.com.br (E.J. Giordani). to offset the limitations of austenitic stainless steel led
0142-1123/$ - see front matter # 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijfatigue.2004.03.002
1130 E.J. Giordani et al. / International Journal of Fatigue 26 (2004) 1129–1136

to the development of the chromium–cobalt alloys were determined using a computer-controlled servo-
in the 1930s and of titanium and its alloys in the 1960s. hydraulic MTS machine and the tests were carried out
However, high cost, low formability and poor machin- with sinusoidal loading control, with a stress ratio of
ability were some of the limitations preventing the 0.01 and a frequency of 10Hz. The test environments
unrestricted use of these metallic materials for orthope- were neutral (air) and aggressive (0.9%NaCl aqueous
v
dic applications. More recent studies have led to the solution at 37  1 C). An acrylic chamber was
development of a new type of high nitrogen austenitic developed to conduct the aggressive environment tests,
stainless steel, ISO 5832-9, which combines high where the specimen gage length was immersed in the
strength, even in the annealed condition, and high cor- aqueous solution while it was cyclically loaded (Fig. 2).
rosion resistance, making it a promising substitute for Electrolyte (1500 ml) was heated in a bath and pumped
F138 steel in the manufacture of orthopedic implants, through the chamber. The electrolyte was replaced for
particularly for applications involving heavier loading each new test. The specimens were used both to deter-
and longer periods inside the human body. mine the S–N curves and to carry out fatigue and cor-
Based on the constant need to gain a better under- rosion–fatigue tests which were interrupted after the
standing of the fatigue and especially the corrosion–fati- first cycle at 10, 20, 30, and 50% of the expected life
gue crack initiation mechanisms, which have been span for a cyclic load of 600 MPa.
reported as the main mechanisms of crack failures The fatigue crack initiation mechanisms were studied
in orthopedic implants, this paper reports on the micro- microscopically, using secondary electron imaging
structural characterization and study of fatigue crack (SEI) and backscattered electron imaging (BEI) in the
initiation mechanisms of an austenitic stainless steel, ISO SEM. Gage sections of the specimens tested in air and
5832-9, in both neutral and aggressive environments. in aqueous environments were extracted and their lat-
eral polished surfaces and fracture surfaces were ana-
2. Materials and methods lyzed. Some of the specimens were electrolytically
etched with HNO3 to reveal grain boundaries.
The material used in this study was an austenitic
stainless steel, ISO 5832-9 (ASTM F 1586), annealed at 3. Results and discussion
v
1030 C for 1 h, used for orthopedic implants. The
chemical composition and the main mechanical proper- 3.1. Microstructure
ties of this steel are listed in Table 1.
The study to characterize this steel was motivated by Fig. 3 illustrates the microstructure of the ISO 5832-
the significant amount of precipitated particles in its 9 steel in the longitudinal section, revealing the presence
microstructure. The study involved the precipitate of a large amount of precipitated particles dispersed
extraction technique with electrolytic dissolution of the in the austenitic matrix. Coarser particles were pre-
austenitic matrix in a methanol solution containing ferentially elongated and were mostly located along the
10% HCl. The extracted residue was analyzed using a grain boundary and parallel to the rolling direction.
JEOL JSM-5900LV scanning electron microscope EDS analyses revealed that these precipitates were rich
(SEM) to characterize the precipitate morphology, a in Nb and Cr. Fig. 4 shows an aspect of the steel’s
Noran Voyager 3050 energy dispersive spectroscope microstructure: the presence of coarse Nb and Cr-rich
(EDS) to identify the phase composition, and a Philips precipitates associated with nonmetallic Al-rich inclu-
PW-1700 X-ray diffractometer to determine its unit cell sions (probably alumina–slag resulting from the deoxi-
dimensions. dation process), revealed by EDS.
Axial fatigue tests were conducted based on the The technique for extracting precipitates by phase
ASTM E 466 standard. Fig. 1 shows the dimensions of dissolution associated with SEM helped to morphologi-
the specimens. The gage sections were mechanically cally characterize the precipitates present in the micro-
polished with diamond paste (1 lm). S–N curves structure of the ISO 5832-9 steel. Fig. 5a shows some

Table 1
Chemical composition and mechanical properties of ISO 5832-9 stainless steel

Chemical composition (mass %)


C Mn Ni Cr Mo N Nb Fe
0.0316 3.83 11.0 22.6 2.41 0.291 0.42 balance
Mechanical properties
Yield strength ðoffset ¼ 0:2%Þ (MPa) Tensile strength (MPa) Elongation (in 25 mm) (%) Vickers hardness
496 861 46 244
E.J. Giordani et al. / International Journal of Fatigue 26 (2004) 1129–1136 1131

Fig. 1. Shape and size of the axial fatigue specimen (in mm).

Fig. 2. Cell used in the axial fatigue test in the aggressive environ-
ment.

Fig. 5. Precipitates extracted from ISO 5832-9 steel: (a) coarse


plate-like precipitates (SEM/BEI); (b) varied morphology and size of
precipitates (SEM/SEI).

of the extracted precipitates on the filter’s membrane.


As can be observed, the larger precipitates show a
plate-like morphology. These particles appear to be
elongated and parallel to the rolling direction on the
longitudinal surface (Fig. 3). Fig. 5b shows a large
amount of precipitates of varied sizes and shapes. EDS
analyses revealed that the extracted precipitates had a
relative chemical composition of metallic elements close
to 63% Nb, 32% Cr and 5% Fe (wt%) or 49% Nb, 44%
Cr and 7% Fe (atomic%), irrespective of the particles’
size or morphology. Therefore, it is logical to suppose
Fig. 3. SEM/BEI micrograph of a longitudinal section of ISO 5832-
that all the analyzed precipitates were of the same phase.
9 steel. An X-ray diffraction analysis of the extracted residue
confirmed the existence of Z-phase precipitates with a
tetragonal structure and lattice parameters of a ¼ 3:035 G
and c ¼ 7:380 G, which were first identified by Hughes [5]
and later reformulated by Jack and Jack [6]. Z-phase is a
complex nitride of niobium and chromium with unit cells
containing Nb2Cr2N2. In the present work, the chemical
composition of the extracted precipitates showed a small
amount of Fe replacing the Cr atoms in the lattice, as did
Örnhagen et al. [7], who investigated two alloys in con-
formity with ISO 5832-9 standard, but unlike those
authors, no evidence of Mo replacing some of the Nb
atoms was found. The most probable composition of the
Z-phase in the present steel is Nb2(Cr, Fe)2N2.

3.2. Fatigue properties

Fig. 6 illustrates the S–N curves for the ISO 5832-9


Fig. 4. SEM/SEI micrograph of coarse Nb-rich precipitates associa- steel obtained from fatigue tests in air and in the
ted with Al-rich nonmetallic inclusions. aggressive environment.
1132 E.J. Giordani et al. / International Journal of Fatigue 26 (2004) 1129–1136

The study of the fatigue and corrosion–fatigue crack


initiation mechanisms aimed to identify the preferential
sites for crack initiation and the microstructural parti-
cularities and/or peculiarities associated to these sites
that are possibly responsible for precocious fatigue or
corrosion–fatigue crack initiation, thus contributing to
reducing the fatigue life of ISO 5832-9 steel.
A point worth noting is that no evidence of localized
corrosion (pits) was found on the surface of specimens
tested in the aggressive environment. Therefore, the
hypothesis that the crack initiation mechanism resulted
from the formation of corrosion pits, as verified by
Qian and Cahoon [10] in a commercial type 316 stain-
less steel in an aggressive environment (0.5 M NaCl at
pH ¼ 4:2), was discarded for the test condition used
here. However, some researchers have observed the
occurrence of the corrosion–fatigue synergy in environ-
ments where pit corrosion does not usually occur [11].
Another important aspect is the fact that no relevant
differences between fatigue crack initiation mechanisms
Fig. 6. S–N curves for ISO 5832-9 steel in air and in 0.9% NaCl sol-
v in air and in aggressive environments have been found.
ution at 37 C.
The analysis of secondary cracks on the lateral surface
of the specimens revealed that the preferential sites for
It is important to state that the levels of maximum
both fatigue and corrosion–fatigue crack initiation
cyclic stress used to determine the fatigue curves were,
were coarse Z-phase precipitates associated or unasso-
in every case, equal to or higher than the yield strength
ciated with Al-rich inclusions as shown in Fig. 7. An
of the steel (Table 1). Hence, the specimens underwent
analysis of the surface of a specimen loaded in air with
plastic deformation mainly in the first loading cycles.
a maximum cyclic stress of 600 MPa revealed the pres-
As can be observed, the aqueous environment exerts a
ence of geometric discontinuities, which already
marked influence on the steel’s fatigue life which is
occurred in the first loading cycle and were associated
further reduced with an increasing number of cycles to
with the fracture of elongated Z-phase precipitates
failure or even with longer testing time, since the fre-
(Fig. 8a) and nonmetallic inclusions surrounded by Z-
quency used is the same for all the specimens. This
phase precipitates (Fig. 8b). These geometric dis-
aspect is considered attributable to the corrosion–fati-
continuities act as stress and strain concentrators and
gue synergy. It is presumed that corrosive environ-
contribute to precocious fatigue crack initiation.
ments merely help the formation of geometrical
No fatigue cracking was found propagating into the
discontinuities on the previously smooth surface of the
matrix in specimens tested in air at 10% (17,500 cycles)
specimen, thereby favoring crack initiation [8]. Because
and 20% (35,000 cycles) of expected fatigue life. Crack-
this process is time-dependent, it is logical to assume
ing was observed only in tests interrupted at 30%
that the influence of the aggressive environment on the
(52,500 cycles) and 50% (87,500 cycles) of expected
material’s fatigue life increases as the metal is subjected
fatigue life, as illustrated in Fig. 9. Cracking in speci-
to prolonged exposure in a given environment.
mens tested in the aggressive environment was already
3.3. Fatigue crack initiation mechanisms visible at 10% (6000 cycles) of the expected fatigue life,
as illustrated in Fig. 10a. Both in air and in the
Understanding how a metal’s microstructure influ- aggressive environment, the larger and more precocious
ences fatigue and corrosion–fatigue endurance is parti- cracks were found to initiate in coarse elongated pre-
cularly important for the development of alloys and of cipitates or clusters of precipitates usually associated
thermal or thermomechanical treatments aimed at with Al-rich inclusions.
improving the performance of the material under cycli- The aggressive environment caused the crack
cal loading conditions in neutral or aggressive environ- initiation period to drop from approximately 50,000
ments through microstructural modifications. A good cycles to about 6000 cycles. In the aggressive environ-
understanding of the correlation between the micro- ment, the larger cracks displayed an average length of
structure and corrosion–fatigue behavior can also be approximately 100 lm (Fig. 10b) 24,000 cycles (40% of
useful to identify the mechanism whereby the aggress- the expected fatigue life under this condition) beyond
ive environment contributes to the initiation and the period required for the initiation of the first cracks
propagation of a fatigue crack [9]. (at 10% of the expected fatigue life). On the other
E.J. Giordani et al. / International Journal of Fatigue 26 (2004) 1129–1136 1133

Fig. 7. Sites of corrosion–fatigue crack initiation at a maximum


cyclic stress of 550 MPa: (a) elongated Z-Phase precipitate (SEM/ Fig. 9. Crack initiation sites in ISO 5832-9 steel in air for interrup-
BEI); (b) coarse Z-Phase precipitate associated with Al-rich inclu- ted tests at a maximum cyclic stress of 600 MPa: (a) at 30% (52,500
sions (SEM/SEI). cycles) of expected fatigue life; (b) at 50% (87,500 cycles) of expected
fatigue life (SEM/BEI).

Fig. 8. Geometric discontinuities created from the fracture of Z-


phase precipitates in air in the first cycle at a maximum cyclic stress Fig. 10. Crack initiation sites in ISO 5832-9 steel in the aggressive
of 600 MPa: (a) associated with an elongated Z-phase precipitate; (b) environment for interrupted tests at a maximum cyclic stress of 600
in a coarse Z-phase precipitate associated with Al-rich inclusions MPa: (a) at 10% (6000 cycles) of expected fatigue life; (b) at 50%
(SEM/BEI). (30,000 cycles) of expected fatigue life (SEM/BEI).
1134 E.J. Giordani et al. / International Journal of Fatigue 26 (2004) 1129–1136

hand, the average length of the larger cracks in the on the material’s surface may occur through the separ-
neutral environment reached approximately 20 lm ation of the matrix/particle interface, separation of
(Fig. 9a) 35,000 cycles (20% of the expected fatigue life) internal boundaries in complex particles, or even
beyond the time required for initiation of the first vis- through fracturing of the particle itself [13]. The latter
ible cracks (at 30% expected fatigue life). Hence, the usually occurs when the concentration of stress around
aggressive environment evidently exerts a strong influ- the particle exceeds its own strength but not the cohes-
ence on the initiation of fatigue cracking and on the ive strength in the matrix/particle interface [14]. In the
early stage of crack growth. present work, the main and secondary fatigue cracks in
The purpose of the SEM analyses of the fracture sur- both air and the aggressive environment were associa-
face was to define the fatigue and corrosion–fatigue ted with the fracture of Z-phase precipitates, which
crack initiation mechanism by identifying the main undoubtedly favored precocious fatigue and especially
crack initiation sites that led to the fracture of the spe- corrosion–fatigue crack initiation. According to Örnha-
cimens. These analyses demonstrated the participation gen et al. [7], Z-phase is an inherent phase in this type
of Z-phase precipitates in the initiation of the main of austenitic stainless steel. Fracturing of precipitates
cracks, confirming the relevance of the coarse pre- already in the first cycle of loading, where plastic
cipitates and clusters of precipitates, usually associated deformation occurs, introduces geometric dis-
with Al-rich inclusions, in the fatigue crack initiation continuities similar to microcracks in the austenitic
process (Fig. 11a). The fractographic analysis also matrix (Fig. 8). These geometric discontinuities cause
stress and strain concentration, helping fatigue crack
highlighted the participation of plate-like Z-phase pre-
initiation in these regions.
cipitates as preferential sites for fatigue crack initiation,
In the presence of an aggressive environment, apart
particularly in the aggressive environment (Fig. 11b).
from the mechanical effect associated with the stress
Most investigators agree that fatigue cracking
concentration in the fracture region of the Z-phase pre-
usually begins in or close to regions with particular
cipitates, the effect of electrochemical corrosion exerts a
characteristics, on or just under the surface of the
special influence. Already in the first cycle of loading,
metals. These characteristics may be inclusions, pre-
plastic deformation causes the rupture of many pre-
cipitates, brittle grain boundaries, severe scratches, pits
cipitates, particularly the coarser ones. Due to the brit-
or slip bands [12]. Fatigue crack initiation in particles tle nature of Z-phase particles, the rupture of particles
probably occurs by cleavage and the geometric dis-
continuity associated with this rupture propagates
instantly throughout the particle and is interrupted
only when it reaches the ductile austenitic matrix on
the other side of the particle. In this way, the aggress-
ive environment penetrates into the discontinuity
immediately following rupture of the precipitates. The
aggressive environment therefore immediately comes
into contact with the bare metal (devoid of a passive
film), which acts as an anodic site in relation to the rest
of the passive surface, promoting a process of anodic
dissolution of the metal until it becomes repassivated.
An occluded cell is then created inside this disconti-
nuity. It is known, however, that anodic dissolution
occurs through the release of metallic Mn+ ions and
that an increase in the concentration of these cations in
the electrolyte inside the discontinuity favors the elec-
trostatic attraction of aggressive Cl anions, which
abound in the used electrolyte. It is also known that
the hydrolysis of Mn+ ions leads to a great release of
H+ ions [15], resulting in a considerable drop in the pH
inside the discontinuity. With cyclic loading, the stress
and strain concentrations inside the discontinuity can
lead to an intermittent breakdown of the passivity in
this location. As a result, there is a continuous increase
Fig. 11. Sites of corrosion–fatigue crack initiation: (a) in coarse Z-
phase precipitates associated with Al-rich inclusions, at a maximum in the concentration of Cl- ions and an accentuated
cyclic stress of 650 MPa (SEM/BEI); (b) in coarse plate-like Z-phase drop of the pH of the electrolyte stagnated there. This
precipitates, at a maximum cyclic stress of 550 MPa (SEM/SEI). combination, allied to the shortage of oxygen in this
E.J. Giordani et al. / International Journal of Fatigue 26 (2004) 1129–1136 1135

region, consumed especially by the metal repassivation and high corrosion resistance of ISO 5832-9 steel [18],
reactions, creates a highly aggressive localized compo- attributed mainly to nitrogen in solid solution [15,19–
sition that lowers the stability of the passive film, con- 25], suggests that this steel is very promising for high
tributes to the acceleration of the anodic dissolution performance orthopedic applications. The integrity of
and renders repassivation of the damaged region diffi- these particles could be kept through hot deformation
cult, thus establishing an autocatalytic process by crev- of the steel. The results also suggest that refining and
ice corrosion. Additionally, anodic dissolution can be controlling the morphology of the Z-phase precipitates
accelerated by the highly deformed condition of the would improve the fatigue and corrosion–fatigue
matrix due to the concentration of strain in the vicinity properties of this steel. A possible alternative to aid
of the discontinuity. The atoms in this region are in a this refinement would be to reduce the amount of Al-
higher energy state, needing less activation energy to be rich inclusions, which apparently serve as a substrate
removed from the crystalline lattice than the less for the coarse Z-phase precipitation. However, control-
damaged regions of the material. To complement this ling the size and morphology of Z-phase precipitates,
extremely unfavorable condition, anodic dissolution associated or not with nonmetallic inclusions, would
inside the geometric discontinuity can unblock the slip certainly involve a more complex study of the steps of
interrupted by the accumulation or piling up of dis- the melting and remelting processes, as well as of the
locations present in this region [12,16], favoring sub- thermomechanical treatments to obtain the material.
sequent slipping and rendering the process
autocatalytic. According to Parkins [17], electrochemi-
cal factors associated with geometric discontinuities 4. Conclusions
may be as important as or even more significant than
the factors relating to stress concentration, particularly The study of fatigue crack initiation mechanisms
at the initiation and in the early stage of fatigue crack conducted with the help of a SEM contributed to the
propagation. elaboration of fatigue and corrosion–fatigue crack
The condition of anodic dissolution is more effective, initiation mechanisms, based on the development of
the more aggressive the environment inside the dis- stress and strain concentrations associated with geo-
continuity. Considering the principle that leads to the metric discontinuities, which were generated by the
phenomenon of crevice corrosion, the geometric dis- fracture of coarse particles characterized as Z-phase
continuity (crevice) that keeps the electrolyte more precipitates, associated or not with Al-rich nonmetallic
stagnated in its interior contributes to the development inclusions. In the presence of an aggressive environ-
of a more aggressive electrolyte. Hence, it is logical to ment, the electrochemical effects of corrosion take on a
assume that the discontinuities generated from the frac- special importance due to the establishment of very
ture of precipitates in fine plate-like form, such as the aggressive conditions of the stagnated electrolyte inside
one illustrated in Fig. 11b, are more effective as they these geometric discontinuities. This creates favorable
keep the electrolyte stagnated and, consequently, main- conditions for the crevice corrosion process, thereby
tain the aggressiveness in its interior. In these cases, the contributing considerably to reduce the fatigue crack
electrochemical factor of the localized corrosion is an initiation stage and to accelerate the early stage of fati-
important fatigue crack initiation factor. This may gue crack propagation.
explain the higher incidence of secondary cracks asso-
ciated with fine and elongated precipitates (probably in
Acknowledgements
plate-like form) in specimens cyclically loaded with a
lower stress in the aggressive environment when com- This research was partially performed at LME/
pared to the specimens tested in air at the same stress LNLS—National Synchrotron Light Laboratory, Bra-
levels. zil. The authors are also indebted to Baumer S/A for
The results obtained from the study of the fatigue supplying the stainless steel for this study, and to
crack initiation in ISO 5832-9 stainless steel, in air and CNPq for its financial support of this work.
in the aggressive environment, led to the supposition
that the rupture of Z-phase precipitates, exhaustively
verified in the present work, when the material is plasti- References
cally deformed—even with a very small plastic strain—
may limit the use of cold work processes in the manu- [1] Gotman I. Characteristics of metals used in implants. J
facture of orthopedic implants for applications involv- Endourol 1997;11(6):383–9.
ing severe and heavy long-term cyclic loading inside the [2] Lopez GD. Biodeterioration and corrosion of metallic implants
and prosthetic devices. Med Buenos Aires 1993;53(3):260–74.
human body, e.g. lower extremity prostheses. However, [3] Rimnac CM, Wright TM, Bartel DL, Klein RW, Petko AA.
provided the integrity of the Z-phase precipitate parti- Failure of orthopedic implants: three case histories. Mater Char-
cles is maintained, the combination of high strength act 1991;26(4):201–9.
1136 E.J. Giordani et al. / International Journal of Fatigue 26 (2004) 1129–1136

[4] Sivakumar M, Rajeswari S. Corrosion induced failure of stain- [17] Parkins RN. Significance of pits, crevices, and cracks in
less steel orthopaedic implant device. Steel Res 1995;66(1):35–8. environment-sensitive crack growth. Mater Sci Technol 1985;
[5] Hughes H. Complex nitride in Cr–Ni–Nb steels. J Iron Steel Inst 1(6):480–6.
1967;205:775–8. [18] Giordani EJ. Properties, microstructure and mechanisms of fati-
[6] Jack DH, Jack KH. Structure of Z-Phase, NbCrN. J Iron Steel gue crack initiation of two austenitic stainless steels used as bio-
Inst 1972;210:790–2. materials (in Brazil). Doctor Thesis, State University of
[7] Örnhagen C, Nilsson J-O, Vannevik H. Characterization of a Campinas; 2001.
nitrogen-rich austenitic stainless steel used for osteosynthesis [19] Pickering FB. Some beneficial effect of nitrogen in steels. In:
devices. J Biomed Mater Res 1996;31(1):97–103. Foct J, Hendry A, editors. Proceedings of the International Con-
[8] Sudarshan TS, Srivatsan TS, Harvey II DP. Fatigue processes in ference on Nitrogen Steels, Lille, France, 1988. London: The
metals—role of aqueous environments. Eng Fract Mech 1990; Institute of Metals; 1989, p. 10–31.
36(6):827–52. [20] Jargelius-Pettersson RFA. Electrochemical investigation of the
[9] Dickson JI, Shiqiong L, Baı̈lon J-P. Microstructural and fracto- influence of nitrogen alloying on pitting corrosion of austenitic
graphic aspects of corrosion fatigue. Mater Charact 1992; stainless steels. Corros Sci 1999;41(8):1639–64.
28(4):327–47. [21] Levey PR, VanBennekom A. A mechanistic study of the effects
[10] Qian YR, Cahoon JR. Crack initiation mechanisms for corrosion of nitrogen on the corrosion properties on stainless steels. Cor-
fatigue of austenitic stainless steel. Corrosion 1997;53(2):129–35. rosion 1995;51(12):911–21.
[11] Martin JW, Talbot DEJ. A study of crack initiation in corrosion [22] Rondelli G, Vicentini B, Cigada A. Localized corrosion tests on
fatigue of AISI type 316 stainless steel by dynamic measurement austenitic stainless steels for biomedical applications. Br Corros
of corrosion current transients. Nucl Technol 1981;55(2):499–504. J 1997;32(3):193–6.
[12] Srivatsan TS, Sudarshan TS. Mechanisms of fatigue crack [23] Azuma S, Miyuki H, Kudo T. Effect of alloying nitrogen on
initiation in metals: role of aqueous environments. J Mater Sci crevice corrosion of austenitic stainless steels. ISIJ Int
1988;23(5):1521–33. 1996;36(7):793–8.
[13] Shih T-Y, Araki T. The effect of non-metallic inclusions and [24] Sivakumar M, Mudali UK, Rajeswari S. In vitro electrochemical
microstructures on the fatigue crack initiation and propagation investigations of advanced stainless steels for applications as
in high strength carbon steels. Trans ISIJ 1973;13(1):11–9. orthopaedic implants. J Mater Eng Perform 1994;3(6):744–53.
[14] Husain Z. Initiation and early stages of growth of corrosion fati- [25] Mudali UK, Dayal RK, Gnanamoorthy JB, Rodriguez P.
gue cracks in a structural steel. Mater Sci Eng, A 1989;119:L1–4. Relationship between pitting and intergranular corrosion of
[15] Grabke HJ. The role of nitrogen in the corrosion of iron and nitrogen-bearing austenitic stainless steels. ISIJ Int 1996;36(7):
steels. ISIJ Int 1996;36(7):777–86. 799–806.
[16] Whitwham D, Evans UR. Corrosion fatigue—the influence of
disarrayed metal. J Iron Steel Inst 1950;165(1):72–9.

You might also like