Startseite Corrosion performance of cold deformed austenitic stainless steels for biomedical applications
Artikel Öffentlich zugänglich

Corrosion performance of cold deformed austenitic stainless steels for biomedical applications

  • Mohd Talha , Yucong Ma , Yuanhua Lin EMAIL logo , Yong Pan , Xiangwei Kong , O.P. Sinha und C.K. Behera
Veröffentlicht/Copyright: 26. April 2019

Abstract

Austenitic stainless steels possess an excellent balance of strength and ductility along with the high ability to further raise their strength during cold deformation (CD). Corrosion resistance of austenitic stainless steels (SSs) is affected by cold deformation because passive films on the surface of steels are expected to be modified. A low level of CD enhances the surface diffusion, which results in the formation of a stable passive film leading to an increase in the corrosion resistance in neutral chloride solutions. The chromium content in the passive film on a deformed steel surface is usually richer, with a higher Cr/Fe ratio than that formed on annealed steels. A higher chromium content makes surface films more stable, which improves the corrosion resistance. However, severe CD results in the formation of strain-induced martensite phase and deformation twins, which decreases the localized corrosion resistance by increasing the number of active anodic sites on the surface. The corrosion resistance, especially the pitting resistance, in SSs is diminished with increasing volume fraction of the martensite. In this review, we highlighted the failure modes of corrosion of stainless steel implants, factors affecting corrosion, and effect of CD on mechanical properties and emphatically on the corrosion resistance of SSs for biomedical applications.

1 Introduction

A large number of materials such as metals, alloys, polymers, and composites have been widely used to devise bio-implants. Metallic materials are most extensively used as biomaterials because, when compared to polymeric and ceramic materials, they have superior mechanical properties (Sumita et al., 2004). Typical metallic biomaterials include stainless steels (SSs), Co-based alloys, commercially cp-Ti, and Ti-based alloys (Pound, 2014). Cp-Ti and Ti alloys are normally preferred because of their lower modulus, superior biocompatibility, and greater corrosion resistance (Woodman et al., 1984; Liu et al., 2016). However, austenitic SS (e.g. 316L) is still the broadly used material for many applications because of its promising combination of mechanical properties, adequate biocompatibility, and cost effectiveness compared to other metallic implant materials (Disegi & Eschbach, 2000; Schmidt et al., 2001). Although metals exhibit high strength and toughness, they are at risk of chemical and electrochemical degradation. The implant materials may corrode or wear, resulting in the generation of particulate debris, which may, in turn, deteriorate the body environment (Talha et al., 2013). The implants bear aggressive corrosive environment including blood and other constituents of the body fluid (Williams, 1987). The two physical factors that determine the implant corrosion are thermodynamic forces, which cause corrosion either by oxidation or reduction reactions and the kinetic barrier such as the surface oxide layer, which physically prevents corrosion reactions (Chu et al., 2002; Shahryari et al., 2008). However, the surface passive layer is highly prone to localized corrosion such as pitting, crevice, etc., in some environments. The key problem associated with SS is the negative effect of metal ions or fretting debris, which can be released from the implant devices because of corrosion and wear (Eliades et al., 2004).

Mechanical properties and corrosion resistance of SSs is notably affected by metallurgical parameters such as cold and hot deformation, alloy composition, inclusions, heat treatment, grain size, sensitization, and secondary precipitates (Szklarska-Smialowska, 1971; Sedriks, 1985). Cold deformation (CD) of austenitic SSs is unavoidable because these are exposed to different levels of cold working during the final manufacturing stages of components for biomedical applications. CD modifies the surface passive films; thus, affecting corrosion resistance of SSs (Phadnis et al., 2003). It is known that chloride ions (present in the human body), lead to depassivation of the SS surface. Pits can nucleate at carbides, grain boundaries, and other material heterogeneities on the metal surface, which arise due to deformation (Solomon & Solomon, 2017). It was also predicted that on CD, strain-induced martensite and residual stresses are introduced on the surface in significant amount, which affects the localized corrosion resistance by increasing the number of active anodic sites (Elayaperumal et al., 1972). However, the change in the corrosion-resistant property of SS depends on the solution condition and degree of deformation, which shall be discussed later on in this review.

Sufficient research was done in the past to evaluate the effect of CD on the corrosion resistance of SS for medical applications. Hence, it is necessary to review and summarize the corrosion performance of cold-deformed SSs for bio-implants. This review, in particular, assesses the failure modes of corrosion of implant materials, role of passive film in corrosion, and factors affecting corrosion of SSs for medical applications. Furthermore, we review the effect of CD on mechanical properties and particularly on corrosion resistance of SSs.

2 Failure modes of corrosion of stainless steel implants

Surgical implants are exposed to strict working conditions, including mechanical loading (cyclic and shock), chemical aggression (body fluids), and the static or dynamic contact between the implant material and the state of implant-to-bone contact (Azevedo & Hippert, 2002; Roffey, 2012). Even though there are many forms of corrosion damage, the rate of attack of general corrosion is very low due to the presence of passive surface films on most of the metallic implants used (Geetha et al., 2009). An ample range of failure mechanisms may occur, including pitting corrosion, crevice corrosion, wear, fretting, fatigue, and combinations of them. The mechanical failures of metallic implants may also be influenced by numerous other factors, including the design, material selection, manufacturing practice, improper installation, post-operative complications, and misuse (Azevedo & Hippert, 2002). The various failure modes of implants are as follows.

2.1 Localized corrosion (pitting and crevice corrosion)

Pitting corrosion is one of the most occurring localized corrosion in metallic implants. Studies reported that more than 90% of the failure of implants of 316L SS are due to pitting and crevice corrosion attack (Sivakumar et al., 1995). Pitting corrosion is caused by local dissolution of passive film and the formation of cavities surrounded by an intact passive surface (Zsklarska-Smialowska, 1986). It was suggested that the absorption of the damaging ions on the surface of the passive film of SSs produces mutual repulsion, which lowers the interfacial surface tension, and when the repulsive force is sufficient, the passive film cracks (Zaya, 1984). Pitting usually occurs in solutions containing halide ions in which chloride ion is the most aggressive. Pits almost initiate due to chemical or physical heterogeneity at the surface such as inclusions, second-phase particles, solute-segregated grain boundaries, flaws, mechanical damage, and dislocations (Zsklarska-Smialowska, 1986). Most engineering alloys have many such defects, and a pit tends to initiate at the most susceptible sites first. For example, pits in SSs are often associated with manganese sulfide (MnS) inclusions, which are found in most commercial steels. The role of MnS inclusions in promoting the breakdown and localized corrosion of SSs were recognized (Frankel, 1998; Sidharth, 2009). The initial stage of pitting is generally agreed to be of very short duration, and pit initiation is strongly dependent on the surface conditions of the material (Melchers, 1994).

Crevice corrosion is also a type of localized corrosion similar to pitting, which occurs preferentially in regions on the metal surface where mass transfer is limited, e.g. in narrow crevices or other shielded areas. In the case of bone plates and screws, crevice corrosion can occur at the interface between the tightly contacted bone plate and the screw on the countersink (Buhagiar et al., 2012). To function as a corrosion site, a crevice must be wide enough to permit liquid entry but suitably narrow to maintain the stagnant zone. This type of attack occurs in numerous mediums, although it is usually most intense in ones containing chloride ions. Crevice corrosion can be viewed as a less severe form of localized corrosion when compared with pitting. During crevice corrosion, the attack is localized within shielded areas, while the remaining surface suffers slight or no damage. The presence of crevice corrosion in the countersink portion of the bone plate can induce crack propagation.

2.2 Tribocorrosion/fretting corrosion

Tribocorrosion is defined as the conjoint action of mechanical wear and corrosive attack on a material surface. A special mode of tribocorrosion highly relevant to implants such as hip, knee, and shoulder replacements is fretting corrosion. Fretting corrosion is a form of damage that occurs at the interface of two closely fitted surfaces when they are subjected to slight oscillatory slip and joint corrosion action. Fretting corrosion is very common in all load-bearing metallic orthopedic implants. Fretting corrosion can radically alter the corrosion behavior by mechanically destroying the passive film. Fretting corrosion, alone or in combination with crevice corrosion, was identified as one of the important factors in implant corrosion (Brown et al., 1995; Khan et al., 1999). Fretting can occur even in the absence of a corrosive medium. The medical significance of fretting attack lies in its intensity that may give rise to a high amount of corrosion products in adjacent tissues, or it may be a major factor in crack initiation and fracture failure (fatigue) of an implant (Syrett & Wing, 1978).

2.3 Stress corrosion cracking

The presence of mechanical loading may create further complication of sudden fracture of implants, due to the phenomenon of stress corrosion cracking (SCC), which is believed to be the most dangerous form of corrosion-assisted failures. SCC may be a particular concern for devices having sharp contours such as pins and screws that are used to secure orthopedic implants. Stress corrosion cracks may propagate unobserved to a sudden disastrous failure. Such sudden failures caused by SCC of an implant may have serious results such as difficult removal of the failed device and painful irritation or inflammation of the surrounding tissues. In the past, numerous instances of fracture due to SCC of metallic implants of SSs, titanium alloys, and Co-Cr alloys were reported (Bombara & Cavallini, 1977; Sivakumar & Rajeswari, 1992). Pitting is one of the common SCC initiators, while the localized crack-tip dissolution facilitates crack propagation (Raman & Siew, 2014). Crack initiation takes place due to the localized electrochemical dissolution of the metal. The protective films that form at the crack tip break with sustained tensile stresses, resulting in exposure of the fresh anodic metal to the corrosive medium continuously, and hence, the SCC propagates progressively over a period of time (Panda et al., 2014).

2.4 Fatigue

Biomedical materials that are subjected to cyclic loading and enough stresses in the presence of an aggressive environment fail due to fatigue (Kamachi & Baldev, 2008). The fatigue failure occurs in two stages: (i) the initiation of a fatigue crack and (ii) crack propagation to failure with repeated cyclic loading. Mechanical failure of orthopedic implants is the most common due to fatigue. In some cases, however, the mechanism responsible for crack initiation and crack propagation may be different (Hu et al., 1993). Concerning the fatigue properties, the surface condition is crucial for the development of a crack under cyclic loads. Nucleation of fatigue cracks has been associated to the presence of pits on the surface of metallic materials (Sudarshan et al., 1990). Under fatigue conditions, the aqueous environment can accelerate the initiation of a surface flaw and propagate it to a critical size, leading to fracture. This process is known as corrosion fatigue, indicating the failure of a material under the simultaneous action of cyclic load and chemical attack (Suresh, 2004). The reduction in fatigue life of SS implants under corrosion fatigue was well documented (Giordani et al., 2004).

The fatigue process is found to increase further due to the formation of wear debris leading to fatigue wear. During fatigue, there is disruption of the oxide layer and the inability of the material to repassivate immediately, exposing some regions of the metal to the environment leading to corrosion. The initiation of crack due to fatigue was examined during the measurement of the corrosion potential of CD 316L SS, and it was also observed that the fatigue strength dropped radically when the repassivation was suppressed. This confirms that the oxide layer formation plays a vital role in the determination of the fatigue life of the materials exposed to an aggressive corrosive environment (Park, 1984). In addition, fretting that occurs between an implant and the bone is also found to accelerate the fatigue as the repassivation becomes more difficult in the presence of fretting (Geetha et al., 2010).

2.5 Metal release in a biological environment and their effect

The rupture of the protective surface oxide layer subsequently results in exposure of the nascent surface to corrosive attack and the potential release of metal ions (Shaw & Kelly, 2006). A huge amount of released metal ions could be harmful to human health and may eventually lead to severe complications and failure of the implant system. Metal ion release was frequently recognized as a cause of clinical failure or cutaneous allergic reactions. The correlation between metal ion release and hypersensitivity reaction to Ni, Cr, and Co released from SS and Co-Cr-Mo alloy was observed (Summer et al., 2007). Dermatitis related to Ni toxicity was reported by many researchers. Numerous studies on animals also showed the carcinogenicity of Ni through inhalation, and intravenous and intramuscular parenteral administration of ground Ni complexes. Cr toxicity appears to be closely related to the Cr valence state. Cr(+6), is hemotoxic, genotoxic, and carcinogenic (Barceloux, 1999). When Cr(+6) enters the bloodstream, it damages blood cells by causing oxidation reactions. This oxidative damage can lead to hemolysis and, ultimately, failure of the kidneys and liver (Dayan & Paine, 2001).

The surface oxide of SS is present in complex environments like the body, exposed to different ligands (complexing agents) such as citrate and proteins that influence metal ion release because they are the so-called electrolytes (Hedberg et al., 2017). This induces ligand adsorption, complexation with a surface oxide/hydroxide metal atom, and the possible detachment of the ligand-metal complexes from the surface oxide. Ligand-induced metal release is, thus, adsorption dependent (Carbonaro et al., 2008) but also dependent on pH, stability constants of surface complexes, and other parameters such as temperature (Milosev & Strehblow, 2000; Carbonaro et al., 2008; Hedberg et al., 2011). Depending on the environment, the metal released from steel is also influenced by other processes including active corrosion and protonation (Hedberg et al., 2011, 2013). It was reported that the metallic materials with strong passive films exhibit lower quantities of metal ion release from implants. Therefore, the amount of a metal ion released changes depending on the nature and strength of the metal-oxide bond, structure, role of alloying elements, composition, and thickness of oxide films (Browne & Gregson, 1994). The role of alloying elements in passive films formed on 316L SS and Co-Cr-Mo alloy in a calf serum solution was observed by X-ray photoelectron spectroscopy (XPS). The XPS spectra of the passive films demonstrated the peaks of Cr2O3, CoO, Fe oxide, and Mo oxide (Okazaki et al., 1997). Further, XPS examination made clear that metal release from various grades of SS into synthetic body fluids changes depending on the Cr content in the oxide film. The rate of metal release from SSs into artificial lysosomal fluid decreases with the increase in Cr/(Cr + Fe) ratio in the oxide film, which clearly illustrates that Cr is attributed to passive film formation (Herting et al., 2007).

3 Human body environment

The water content of the human body ranges from 40% to 60% of its total weight. On the basis of function, the total body water can be subdivided into two major fluid compartments, namely, the extracellular and the intracellular fluids. Extracellular fluids (ECFs) consist of the plasma found in the blood vessels. The interstitial fluid surrounds the cells, the lymph, and transcellular fluids such as cerebrospinal fluid and joint fluids. Intracellular fluid (ICF) refers to water inside the cells. Both the amount and the distribution of body fluids and electrolytes are kept normal and constant by a mechanism known as homeostasis (Eliaz, 2009). Blood plasma has high concentrations of sodium, chloride, bicarbonate, and proteins. In contrast, the ICF has elevated amounts of potassium, phosphate, magnesium, and proteins. Overall, the ICF contains a high concentration of potassium and phosphate, while both plasma and the ECF contain high concentrations of sodium and chloride ions. Implants remain in contact with blood when inserted into the body except the dental implants. There is a slight difference between the surrounding environment of implant placed as stents and implant for orthopedic use. Table 1 shows the chemical composition of human blood plasma (Oyane et al., 2003). The chemical environment of blood plasma is highly aggressive for many metals and alloys, especially due to the presence of a high concentration of chloride ions and their ability to induce localized corrosion. Other important considerations of body fluids are the dissolved oxygen and pH levels. The body temperature of 37°C can accelerate electrochemical reactions. However, the dissolved oxygen level in the blood is lower than in artificial solutions exposed to air atmosphere due to the combination of oxygen with hemoglobin, which is the main component of red blood cells. Other ions may also contribute to the corrosion process, either as accelerators or inhibitors (Virtanen et al., 2008).

Table 1:

A typical chemical composition of normal human blood plasma (Oyane et al., 2003).

Ion Na+ K+ Mg+2 Ca+2 Cl HCO3 HPO4−2 SO4−2
Conc. (mmol l−1) 142.0 5.0 1.5 2.5 103.0 27.0 1.0 0.5

4 Various biomedical implants used and their corrosion behavior

4.1 Cardiovascular implants

The use of cardiovascular implants is subjected to its blood compatibility and its integration with the neighboring environment where it is implanted (Agrawal, 1998). Cardiovascular implants should possess unique blood biocompatibility because one of the leading causes of cardiac biomaterial failure is the initiation of thrombosis formation by the blood-contacting devices (Weng et al., 2012). The use of cardiovascular implant applications includes heart valves, endovascular stents, and stent-graft combinations (Jaganathan et al., 2014). One of the important applications of the cardiovascular implant is stents. The 316 SS, Ti and its alloys, and Co-Cr alloys are frequently used metallic materials for stents. The 316L SS is most commonly used as a metal for stents either with or without a coating material. The Co-Cr alloys, which conform to ASTM standards F562 and F90, are also being used for making stents because of their high elastic modulus (210 GPa) (Geetha et al., 2010).

Shape-memory alloys were proposed for cardiovascular stents due to their self-expansion ability. In the medical field, the family of Ni-Ti alloys (nitinol) is the most popular one that has been widely used because of its good biocompatibility, significant resistance to corrosion and fatigue, and the fact that its elastic modulus is close to that of human bone (approximately 35 GPa) (Dai and Ning, 2004). Nitinol constituting 49.5–57.5 at% nickel is used for fabricating self-expanding stents. Self-expanding stents have a smaller diameter at room temperature and are capable of expanding up to their preset diameter at body temperature. After implantation, it regains its original shape and conforms to the vessel wall because of the increase in temperature inside the body (Kurella, 2006). In comparison with other alloys, nitinol exhibits excellent fatigue properties at high strain levels. It forms a passive TiO2 layer that acts as both a physical barrier to nickel oxidation and protects the bulk material from corrosion. Its corrosion resistance is greater than that of SSs (Ryhanen, 1999). In some cases, Ni or Ti (only a few percent) in Ni-Ti alloys can be partially replaced by Cu, Co, Fe, Nb, or Mo to improve the hysteresis (stress and/or temperature hysteresis), corrosion performance, control of transformation temperatures, fatigue behavior, etc. (Humbeeck et al., 1998).

In comparison to other alloys, nitinol shows excellent fatigue properties at high strain levels.

4.2 Orthopedic implants

One of the most prominent applications of biomaterials is as orthopedic implant devices. The materials used for orthopedic implants, particularly for load-bearing applications, should possess superior corrosion resistance in the body environment, excellent combination of high strength and low modulus, high fatigue and wear resistance, and high ductility. Orthopedic implants include both temporary implants such as plates and screws and permanent implants that are used to replace the hip, knee, spine, shoulder, etc. The corrosion mechanisms that occur in temporary implants are crevice corrosion at shielded sites in the screw/plate interface and beneath the heads of fixing screws and pitting corrosion of the implants made of SS (Yu et al., 1993; Jones, 1996). One of the major causes of the failure of the orthopedic implants is wear, which, in turn, is found to accelerate the corrosion. Hence, high wear-resistant materials such as Co-Cr alloys are often preferred to fabricate orthopedic implants. The femoral components are sometimes coated with cement to have good fixation. Total joint replacement is widely regarded as the major achievement in orthopedic surgery.

4.3 Dental implants

These implants face a more aggressive environment in the mouth because the pH of saliva varies from 5.2 to 7.8. Thus, the major reasons for the corrosion of metallic dental implants and fillings are temperature, quantity and quality of saliva, plaque, pH, protein, and the physical and chemical properties of food and liquids as well as oral health conditions. Chaturvedi (2009) reviewed broadly the corrosion aspect of dental implants. Sometimes, as two metallic components are used together in making dental implants, galvanic corrosion occurs very frequently in dental implants. Titanium and its alloys are resistant enough to pitting corrosion in different in vivo conditions encountered; however, they undergo corrosion in high fluoride solutions in dental cleaning procedures (Probster et al., 1992).

4.4 Other implants

Evolution of cochlear implant technology provides a sense of sound to a person with severe to profound sensorineural hearing loss. Binaural cochlear implantation was used in children as well as in adults. The implant consists of an external portion that sits behind the ear and a second portion that is surgically placed under the skin. Because of their development, perimodiolar electrodes, implantable microphones, and rechargeable batteries, etc., are used for fully implanted devices (Balkany et al., 2002). Solid biocompatible implantable devices for sustained or controlled intravitreal drug delivery to the posterior segment of the eye were developed employing diverse approaches and including osmotic mini-pumps, nonbioerodible and bioerodible drug-loaded pellets, configured capillary fibers, biodegradable scleral plugs, scleral discs, polymeric matrices, and scaffolds of a variety of geometries providing unique mechanisms of drug release for the delivery of drugs (Choonara et al., 2010).

In all types of corrosion of metallic implants, the surface passive film plays a crucial role in corrosion resistance, and the film determines in vivo metal ion release (Hanawa, 2004). The function of the passive film on the surface of SSs shall be discussed in the coming section of this review. The special mode of corrosion of SSs used as implants is pitting corrosion in the presence of body fluids due to the presence of a higher extent of chloride ions. Therefore, this review will focus especially on pitting corrosion in a chloride-containing solution.

5 Role of passive film in corrosion resistance of stainless steels

It is generally recognized that the susceptibility of metals and the rate at which the corrosion processes take place are closely associated to the quality of the passive film, which typically occurs on the metal surface (Zsklarska-Smialowska, 1986). Although the nature of the passive films is not totally understood, many researchers accept that corrosion susceptibility is related to local imperfections or discontinuities (Moayed et al., 2003). The structure and composition of the passive film formed may depend on many factors such as pretreatment, composition of the metal, electrode potential, polarization time, environmental chemistry, temperature, etc. (Drogowska et al., 1996). The barrier properties of passive films greatly reduce in the presence of chloride ions (Metikos-Hukovic et al., 2011). Passive films formed on SSs in aqueous environments are typically very thin (about 3.6 nm), contain higher chromium, and are protective in aggressive environments (Hanawa et al., 2002; Metikos-Hukovic et al., 2011). This film consists of a mixture of oxides and hydroxides of iron and chromium, and there is considerable evidence of a dual structure encompassing an inner oxide and an outer hydroxide layer (Clayton & Lu, 1986; Olefjord & Wegrelius, 1990). In this context, it is valuable to point out that the electronic structure of passive films plays an important role in the corrosion resistance of the alloy.

When describing the electronic-semiconducting properties of passive films, they are assumed as amorphous semi-crystalline structures with localized states within the band gap (Dean & Stimming, 1987). The localized states become ionized when the Fermi level cross them inducing the space charge layer at the solid/liquid interface, which results in an n- or p-type semi-conductivity of the passive layer (Metikos-Hukovic et al., 2011). For autenitic SS and Fe-Cr alloys, it is expected that the semiconducting behavior reflects the duplex structure of their passive films. It was considered that the inner barrier layer is essentially formed of chromium(III) oxide, which behaves as a p-type semiconductor, whereas the outer region of the passive film is composed mainly of iron(III) oxide, which acts as an n-type semiconductor (Gaben et al., 2004). The point defect model assumed that oxygen vacancies and cation interstitials in passive film impart n-type semiconductor properties, while cation vacancies impart p-type semiconductor properties (Sikora et al., 1996). Jinlong and Tongxiang (2015) investigated that the chromium oxides are stable in acidic solution, while iron oxides can dissolve easily in acidic solution; therefore, less chromium vacancies in the inner layer and more oxygen vacancies in the outer layer are formed. The oxide film develops by the passage of metal cations at the matrix from the oxide/metal interface to the metal/solution interface. At the same time, vacancies would be left at the oxide/metal interface and finally leading to a local accumulation, which results in stresses in the passive film and its subsequent breakdown (Chao et al., 1981; Lin et al., 1981).

6 Factors affecting corrosion of stainless steels in body fluid

The corrosion behavior of an implant is influenced by a various factors, including the material itself, the surroundings, as well as the construction. Changes in these variables can have an additional influence on the mode and rate of metal ion release. These factors can be divided into two categories as discussed below.

6.1 Material parameters

This includes the factors that are directly related to the material itself. The following factors are under this category.

6.1.1 Alloying elements

The pit initiation sufficiently depends on the alloy composition and, hence, affects the corrosion rate (Zsklarska-Smialowska, 1986). Interstitial elements such as C and N increase the strength of austenitic SSs, and N is more effective than any other element. As an alloying element, N is identified to enhance the pitting corrosion resistance of SSs, increase the local pH due to the formation of NH4+ ions within the active sites, and enhance the repassivation of the passive film (Olefjord & Wegrelius, 1996). The solubility of nitrogen increases by alloying elements such as Cr, Mn, Mo, Nb, and V and decreases by Ni and Si (Stein et al., 1988; Feichtinger & Zheng, 1990; Feichtinger, 1993). Carbon is also a strong austenite stabilizer similar to nitrogen. Speidel (2006) reported that carbon improves the resistance against pitting and crevice corrosion of high-nitrogen steel (HNS). The pitting resistance equivalent number (PREN) is a predictive measurement of resistance to localized pitting corrosion of SS based on its chemical composition. A higher PREN is generally in agreement with better pitting corrosion resistance in most SSs. There are several PREN formulae. The most common is as follows (Bernhardsson, 1991):

(1) PREN = 1 [ Cr ] + 3.3 [ Mo ] + 16 [ N ]  [content in wt . % ] )

The passive film analysis of high interstitial alloys recognized that the concentration of Cr in the passive film increases with the addition of C and N, thereby, making the passive film more protective (Ha et al., 2009a),b). However, the development of SSs followed the path of trying to reduce the carbon content in the alloys to very low in view of the fact that the added carbon to SSs is prone to form chromium carbide, which reduces the resistance to localized corrosion. Another alloying element used in SSs is Mn, which was usually considered as an austenite former and is frequently added to increase the solubility of nitrogen (Davis, 1994). Mn is used to replace Ni to obtain moderately lower cost than those of common austenitic SSs. However, the addition of Mn is usually accompanied by a reduction of pitting corrosion resistance associated with the formation of MnS inclusions, which are known to be precursor sites for pitting attack. The effect of MnS on the pit initiation process was examined by Stewart and Williams (1992) and confirmed that these inclusions dominated as pit nucleation sites. The presence of Mn drastically reduces the critical pitting temperature (CPT) values, transfers the pitting potential (Epit) toward a less noble site, and increases the corrosion current density of SSs in chloride-containing solutions (Pardo et al., 2008). Mo is another beneficial alloying element for SSs, the positive effect assigned to the presence of Mo6+ within the passive film, which blocks the active surface sites due to the formation of Mo oxy hydroxide or molybdate of chromium and iron (Hashimoto et al., 1979a,b) or by the adsorption of mixed molybdenum compounds in the aggressive pit environment facilitating the pit repassivation (Kolotyrkin et al., 1994). According to Pardo et al. (2008), Mo increases CPT and Epit values for 316L and 304 SS, thereby, reducing progressively the corrosion rate.

6.1.2 Surface topography

Most engineering failures due to fatigue, wear, corrosion, etc., are recognized to initiate at the surface. Despite the variety of demands, today, many metallic medical parts aim at low surface roughness and polished surface with a glossy appearance. The state of the metal surface is known to affect pitting susceptibility (Zamaletdinov, 2007). The more homogenous the surface is, both chemically and physically, the higher is the potential for pitting resistance (Malik et al., 1992). It is well established that at temperatures above the CPT, the pitting potentials tend to decrease as the sample surface roughness increases (Moayed et al., 2003). The higher chance of stable pitting in rough surfaces is attributed to the longer length of diffusion and bigger micro-crevices near the inclusions (Moayed et al., 2003). Burstein and Pistorius (1995), Burstein and Vines (2001) studied the effect of roughness on the pitting of SS in chloride solution. They reported that a smoother surface reduced the frequency of metastable pitting and prohibited the propagation of nucleated sites.

Surface finish is an important section in any specification of SS regardless of the intended use. A rough surface finish can effectively lower the corrosion resistance as mentioned above. SS is available in a wide variety of standard and special finishes. The majority of finishes can be divided into three categories: (i) mill finishes, (ii) mechanically polished finishes, and (iii) special finishes. In each case, the finishes are described under either their appropriate standard or the name by which they are commonly known. Electropolishing of metallic surfaces is used to obtain a high-purity surface for medical equipment or for a decorative and gleaming finish. Electropolishing can also increase the chemical stability of many metal materials (Richardsa et al., 2012). This technology is a dedicated and precise polishing process, which involves electrochemical anodic dissolution. In electropolishing, a low voltage is applied to the anodic work piece, which is immersed in an electrolyte bath, and the surface peaks are chemically smoothened (Kuhn, 2004). It will not only remove the deformed layer and improve surface roughness but also form a thin passive film on the surface. This passive film, because of its metallurgical composition, will increase corrosion resistance (Hryniewicz, 1994). Except these, other significant efforts were also made to develop methods for modification of the SSs surface and/or its passive films in order to improve the corrosion resistance, and its biocompatibility. Such surface treatments include low-temperature plasma nitriding (Gontijo et al., 2006), magnetron sputtering (Saker et al., 1991), and plasma immersion ion implantation (Williams et al., 1991; Wei et al., 1996) and various biocompatible coatings (Nagarajan et al., 2012). The details of such surface treatment techniques are beyond the scope of this review.

6.2 Environmental parameters

This includes the surrounding environment of the implants. The following factors are under this category.

6.2.1 High chloride ion concentration

The presence of chloride ions largely affects corrosion resistance of SS as described above. Because of the high diffusivity of chloride ions, these can easily penetrate the film and accumulate at the metal/film interface (Pardo et al., 2000; Saadi et al., 2016). According to the model proposed by Burstein et al. (2004), chloride ions could further react with a metal substrate at the metal/film interface to form metal chloride. The surface film may rupture consequently owing to the volumetric expansion because the molar volume of metal chloride is larger than that of metal oxide in the surface film. As a result, pit nucleation occurs. During the pitting process, the pH value of the electrolyte in the pit tends to decrease. Once the pH value within the pits drops below the critical value, the repassivation behavior of passive film is inhibited, and pits could further develop into stable pitting (Olefjord & Wegrelius, 1996).

6.2.2 pH and dissolved oxygen

The corrosion behavior of an implant is also influenced by pH and oxygen content present. The body temperature of 37°C can accelerate electrochemical reactions and may change the mechanism of corrosion from that taking place at room temperature. The pH affects the metal release in different ways, either directly such as dissolution by protonation (Okazaki & Gotoh, 2005) or indirectly by changing the extent of adsorption of complex agents or proteins (Merritt & Brown, 1988). The pH of body at the control point is found to be 7.4±0.06 (Virtanen et al., 2008). Normally, changes in the pH value in the body fluids are comparatively small as the fluids are buffered. On implantation, the pH of the surrounding tissue may diminish to values about 5 and, then, recover to 7.4 within a few weeks (Hench & Ethridge, 1975). Changes in the pH value were found in periprosthetic tissue of revised hip implants (Willert et al., 1996). The oxygen content in the surroundings of the implant can vary depending on the particular application. Highly oxidizing conditions for SS can lead to dissolution of the passive film by the formation of soluble Cr (VI) species. In addition to dissolved molecular oxygen, more active oxygen species such as peroxides can be formed in biological reactions (Virtanen et al., 2008). The stability of the passive film is dependent on the accessibility of oxygen. The adsorption of proteins and cells onto the surface of materials could limit the diffusion of oxygen to specific regions of the surface, causing preferential corrosion of the oxygen-deficient areas resulting to the breakdown of the passive layer (Ratner et al., 2004).

6.2.3 Presence of biomolecules

Implants are in contact with blood, which contains a variety of inorganic species and organic compounds of high molecular weight, especially proteins. Implants are covered with a layer of adsorbed protein immediately after insertion into the body and affect the surface reactions (Rudee & Price, 1985). The development of a protein-containing biofilm on the metal surface was shown to increase the corrosion process of the base alloy (Yan et al., 2007). However, it has not yet been clarified unambiguously whether the biomolecules increase or inhibit electrochemical reactions (Yang & Black, 1994; Omanovic & Roscoe, 1999; Milosev, 2002; Minovic et al., 2003). Possibly, the effect is specific for a particular metal/biomolecule combination. The enhancement of the dissolution rate of implants in the presence of proteins can be explained by the formation of complexes between metal ions and proteins (Steinemann, 1996). These complexes can be transported away from the instant surrounding area. To maintain the equilibrium, the dissolution rate of a base metal increases and, consequently, suppresses the formation of the passive layer (Okazaki & Gotoh, 2005). However, the effect of biomolecules on corrosion of SSs is still controversial, and more work is required in this area for better understanding.

6.2.4 Presence of biological cells

Cells may also influence the corrosion of metallic materials (Hanawa, 2004). In the early stage of implantation of biomaterials in the human body, immunocytes, such as macrophages and lymphocytes, accumulate around the materials, and the adhesion of cells, such as fibroblasts and osteoblasts, follows thereafter (Matsumoto, 1978). Cells adhere on the surface through the points of contact with cell-adhesive proteins (Hayashi, 1995). Cells generate various chemicals, such as cytokines (growth factor) and inorganic ions (Alberts et al., 2002). In addition, the cells generate various kinds of extracellular matrices (ECM) (Hayashi, 1995; Lanza et al., 1997). In the ECM, polysaccharides form a gel-like basement containing a lot of water as well as fibrous proteins and cell-adhesive proteins. Water in the gel-like basement governs the diffusion of molecules (Alberts et al., 1990). Those cell adhesion mechanisms and cell products seem to influence the chemical properties of a solution over the material. Gilbert et al. (1998) suggested the retardation of dissolved oxygen near osteoblast cells on Ti because, after the consumption of dissolved oxygen on the surface by cathodic polarization, the recovery of dissolved oxygen concentration is delayed near the cells. Hiromoto and Hanawa (2004) found a decrease in pH from 7.5 to 5.2 to 6.8 near fibroblast L929 cells on 316L steel and Ti. The results indicated that the corrosion behavior of metallic biomaterials should be affected by the presence of cells, such as osteoblasts and fibroblasts, and their ECM (Hiromoto & Hanawa, 2006). Aoyagi et al. (2004) analyzed the composition of the protein layers on 316L steel and Ti formed with and without L929 cells with time of flight-secondary ion mass spectroscopy and verified the difference in composition of the protein layer with and without cells. Hiromoto et al. (2002) found that the cathodic current density on Ti decreases with L929 cells in the polarization test, and according to the fitting calculation of parameters in an equivalent circuit based on alternating current impedance test, the diffusion resistance at the interface increases with L929 cells. Messer et al. (2001) found that the corrosion rate of the metallic material increases with human gingival fibroblasts. Mustafa et al. (2002) performed electrochemical impedance measurement of Ti with culturing osteoblast-like cells and found that the osteoblast-like cell does not significantly influence the corrosion resistance of Ti. Hiromoto and Hanawa (2006) demonstrated that for 316L, the pitting potential decreased with L929 fibroblast cells, indicating a decrease in the pitting corrosion resistance. In addition, collagen coating would provide an environment for anodic reaction similar to that with culturing cells. There are many studies about the effect of cells on corrosion, but the available literature shows an uncertain role of cells on the corrosion of metallic implants. Therefore, more systematic study is required to understand this phenomenon well.

Besides these factors, one important factor that modifies the corrosion behavior of metallic implants is CD given to alloy during the manufacturing of various components. It also greatly affects the mechanical properties of metallic implants. CD has strong effect on austenitic stability, alters the surface passive films, changes the surface free energy, and hence greatly affects corrosion and, hence, biocompatibility. These effects due to CD will be discussed in coming sections of this review.

7 Changes in the properties of austenitic stainless steels due to cold deformation

Austenitic SSs are the most common and familiar types of SS and are widely used as implants. Austenitic SSs even in annealed state have numerous advantages, i.e.: (i) Austenitic SSs have a face-centered cubic (fcc) structure and are characterized by very low yield strength-to-tensile strength ratio and high formability. (ii) To increase strength, cold working and successive strain aging treatment can be applied. (iii) Austenitic SSs have satisfactory corrosion-resistant property. (iv) Austenitic SSs are essentially nonmagnetic. In terms of disadvantages, austenitic SSs generally have higher sensitivity toward pitting corrosion, and have lower crevice corrosion and stress corrosion cracking (SCC) resistance (Shimohira, 1995; IARC, 1999). They can be cold worked to improve hardness, strength, and stress resistance. Austenitic steels are non-magnetic in the annealed condition, although they can become slightly magnetic when cold worked. CD can alter the mechanical properties as well as the corrosion behavior of SSs as discussed below.

7.1 Change in mechanical properties

The material that will be selected for a particular biomedical application is decided by mechanical properties. Hence, the mechanical behavior of an implant material is a very essential factor. Some of the properties that are of primary concern are hardness, tensile strength, modulus, and elongation. It is well known that austenitic SSs cannot be hardened by heat treatments. On the other hand, cold or warm working (drawing, rolling, forging, etc.) can make such SSs harden. A number of studies were performed on the effect of CD on the structural and mechanical behavior of SSs. Bhav Singh et al. (2009) reported that increasing the percentage of cold reduction increases the strength and hardness with loss of ductility for nitrogen-alloyed steels. It is known that austenitic SSs are metastable, which means that they can undergo a diffusionless phase transformation from austenite to martensite by deep cooling or by plastic deformation or by cyclic loading (Takaki et al., 2004; Hou et al., 2011). Strain-induced α′-martensite produced in metastable austenitic SSs during cold rolling led to significant increase in their strength. Formation and the amount of strain-induced martensite depend on the austenite stability (chemical composition and initial austenite grain size) and the rolling conditions (extent of deformation given, temperature, and rolling speed). The amount of deformation-induced martensite also depends upon numerous other factors such as plastic strain, strain rate, stress state, deformation mode, grain size, and grain orientation (Sato et al., 1989; Nakada et al., 2010). The martensite formation is also largely dependent on the steel composition and stacking fault energy (SFE) (Opiela et al., 2009).

When the austenite stability and the deformation temperature are low, or the degree of deformation is high, the martensite content increases (Murata et al., 1993). The increase in α′ martensite formed by plastic deformation causes an alteration in the physical properties of austenitic SSs (Huang et al., 2007). CD can modify the physical properties even when the martensite phase is not formed. The hardness was found to increase with an increase in the percent reduction in thickness, which could be attributed to the work-hardening effect. The Vickers hardness number of 304SS was found to double at the initial stages of cold work (up to 25%) (Roger, 1985). Further, cold working of this type of steel (up to 50%) raised the hardness value only by 7%. Milad et al. (2008) reported that the tensile strength, yield strength, and hardness of 304SS were found to increase with the increase in cold rolling up to 45%. It was found that the ratio of tensile strength to Vickers hardness is constant and is approximately three times up to 50% of cold rolling. Figure 1 demonstrates the engineering stress-strain curves for annealed and cold-worked 304 steel. It may be seen that yield strength (0.2%) and ultimate tensile strength increased, and elongation decreased with an increase in the degree of cold working (Milad et al., 2008). It was demonstrated for 316L and HNS that the annealed materials exhibit large strain before fracture and reveal the largest area under the stress-strain curve and, therefore, possess a high capacity for energy absorption (Talha et al., 2015). The ability for energy absorption decreased with an increase in cold reduction. Lula (1986) also reported that the CD of 304SS caused a substantial increase in the yield and tensile strengths. The ductility was found to decrease sharply at the beginning of rolling (up to 30%); after that, the rate of decrease was observed to be lower. A detailed discussion related to the effect of CD on the mechanical properties of austenitic SSs is beyond the scope of this review. CD has significant impact on the corrosion behavior of austenitic SS, which shall be discussed in the next section.

Figure 1: 
						The stress-strain curves of the deformed and the undeformed 304 SS (Milad et al., 2008), reproduced with permission from Elsevier).
Figure 1:

The stress-strain curves of the deformed and the undeformed 304 SS (Milad et al., 2008), reproduced with permission from Elsevier).

7.2 Change in corrosion resistance

SSs are exposed to different levels of cold work during the final manufacturing stages of components for biomedical applications. Moreover, work hardening was a common process to further enhance the strength of austenitic SSs, and in many cases, they are used at a cold-worked state. CD is known to modify the corrosion resistance properties of SSs by increasing the dislocation density, which may help in increasing the diffusion of certain species, phase transformation, nucleation, etc. (Hamdy et al., 2007; Dwivedi et al., 2017). Both the thickness and composition of passive films are likely to be altered in many ways by CD (Navai and Debbouz, 1999; Phadnis et al., 2003). CD results in the formation of martensite and residual stresses on the surface, which affects the localized corrosion resistance (Elayaperumal et al., 1972).

As discussed above, the austenite phase in SSs exists as a metastable phase. It can be transformed into a different phase such as α′ or ε martensite, during the CD process (Nagy et al., 2004; Shirdel et al., 2015). Figure 2 shows the image quality (IQ) maps and kernel average misorientation (KAM) maps obtained from electron backscatter diffraction (EBSD) for non-deformed and deformed 304 SS samples (Luo et al., 2017). As revealed in Figure 2A, the samples without deformation have irregular austenite grains with only annealing twins. After 5% deformation, the steel possessed the same polygonal microstructure as the non-deformed samples. Slight CD did not alter the grain size or angles of the grain boundaries. After 20% deformation reduction, slip bands and deformation twins were found in some grains. As shown in KAM maps, the non-deformed samples exhibited defect-free grains. It is clear from the KAM maps in Figure 2C2 that the changes in each grain were not uniform after 20% deformation, and the degree of non-uniformity is increased with the increase in the deformation reduction. Numerous low-angle boundaries and CD-induced martensite phases were observed in the austenite matrix with an increase in the degree of CD, and the presence of martensite also affected the uniformity of the grains. Figure 3 reveals he IQ maps, IQ maps with grain boundaries, and KAM maps after 40%, 60%, and 80% CD of 304 SS (Luo et al., 2017). The majority of the twin boundaries disappeared during the deformation process, and numerous low-angle boundaries instead of high-angle boundaries were observed after 40% deformation. Martensite phases were generated at the slip lines or stacking faults. The allocation of the martensite phase was heterogeneous, and some of the remaining austenite phase that did not transform was evidently observed in the microstructure. When the CD increased from 60% to 80%, as shown in Figure 3B, B1, C, and C1, the content of the martensite phase increased with the level of deformation, and the originally large-sized grains were replaced by small-sized grains. The KAM maps (Figure 3A2–C2) exhibit that the grains were non-uniformly deformed, and there was more rigorous distortion in the martensite phase grains. As the deformation reduction increased to 80%, the grain boundaries disappear, and the average grain size could not be accurately measured. It was demonstrated that the grain boundary diffusion and corrosion resistance strongly depend on the crystallographic structure of the grain boundary (Shimada et al., 2002; Kobayashi et al., 2016).

Figure 2: 
						EBSD results of image quality maps, image quality maps with grain boundaries (red lines represent low-angle boundary (2° 5°), green lines represent low-angle boundary (5°–15°), blue lines represent high-angle boundary above 15°), and the KAM maps of 304L SS with different deformations (A), (A1), (A2) 0% deformation reduction; (B), (B1), (B2) 5% deformation reduction; (C), (C1), (C2) 20% deformation reduction (Luo et al., 2017), reproduced with permission from Elsevier).
Figure 2:

EBSD results of image quality maps, image quality maps with grain boundaries (red lines represent low-angle boundary (2° 5°), green lines represent low-angle boundary (5°–15°), blue lines represent high-angle boundary above 15°), and the KAM maps of 304L SS with different deformations (A), (A1), (A2) 0% deformation reduction; (B), (B1), (B2) 5% deformation reduction; (C), (C1), (C2) 20% deformation reduction (Luo et al., 2017), reproduced with permission from Elsevier).

Figure 3: 
						EBSD results of image quality maps, image quality maps with grain boundaries (red lines represent low-angle boundary (2°–5°), green lines represent low-angle boundary (5°–15°), blue lines represent high-angle boundary above 15°), and the KAM maps of 304L SS with different deformations (A), (A1), (A2) 40% deformation reduction; (B), (B1), (B2) 60% deformation reduction; (C), (C1), (C2) 80% deformation reduction (Luo et al., 2017), reproduced with permission from Elsevier).
Figure 3:

EBSD results of image quality maps, image quality maps with grain boundaries (red lines represent low-angle boundary (2°–5°), green lines represent low-angle boundary (5°–15°), blue lines represent high-angle boundary above 15°), and the KAM maps of 304L SS with different deformations (A), (A1), (A2) 40% deformation reduction; (B), (B1), (B2) 60% deformation reduction; (C), (C1), (C2) 80% deformation reduction (Luo et al., 2017), reproduced with permission from Elsevier).

Figure 4 reveals XRD patterns and EBSD phase maps with different deformation reductions of 304 SS (Luo et al., 2017). A single fcc-phase was attained without CD and no phase change was observed after 5% CD. After 20% CD, a weak peak representing the α′-martensite (110) phase was detected in the XRD patterns. However, there is no evidence of formation of martensite up to 20% CD for 316LVM and HNSs (Talha et al., 2014a). As shown in Figure 4, the intensity of the austenite phase peaks decreased, and the martensite phase peaks increased, with an increase in the deformation reduction. The amount of martensite formed was associated with the degree of CD, and the amount of the martensite phase increases with the CD level as shown in the phase maps of the EBSD (Figure 2). Thus, the SSs are more susceptible to martensite formation when deformed at room temperature. It was also reported that during the plastic deformation process, two types of martensite could be formed: α′-martensite (bcc) and ε-martensite (hcp), depending on the chemical compositions, deformation variables, and the stacking-fault energies (Tavares et al., 2003).

Figure 4: 
						XRD patterns and EBSD of phase maps of 304L SS with different CD reductions (blue in the phase map stands for the austenite phase, and the red stands for the martensite phase) (Luo et al., 2017), reproduced with permission from Elsevier).
Figure 4:

XRD patterns and EBSD of phase maps of 304L SS with different CD reductions (blue in the phase map stands for the austenite phase, and the red stands for the martensite phase) (Luo et al., 2017), reproduced with permission from Elsevier).

Many studies have reported the influence of deformation level on the passive behavior of SS in different solutions (Phadnis et al., 2003; Feng et al., 2016; Ren et al., 2016). The influence of CD, in particular, the amount of α′ martensite formation, of AISI 304 SS in 3.5 wt.% NaCl was studied by Xu and Hu (2004). They observed that when the content of martensite was less than 6% and more than 22%, the pitting sensitivity increased. However, when the martensite content was between 6% and 22%, the sensitivity decreased with increasing degree of deformation. Fang et al. (1997) reported that the corrosion potential of the martensite phase was more negative than that of austenite, and this is the primary reason for the selective corrosion of martensite. In the case of high chloride concentration, the degree of pitting increases linearly with increasing volume fraction of martensite (Fang et al., 1997). It was observed that the intergranular SCC of 316L SS and a nickel-based alloy were associated with the increase in the fraction of the low coincidence site lattice (CSL) boundaries (West and Was, 2009). It was demonstrated by Luo et al. (2017) that when the deformation increases, the fraction of low CSL decreases, which increase corrosion in acid solution. The related passive current in acidic media is found to increase with deformation by several orders of magnitude (Barbucci et al., 2002), but an opposite result was reported in neutral chloride-containing medium (Phadnis et al., 2003). For cold-worked 18Cr-10Ni-2Mo steel, the number of pits generally increases with an increase in deformation in NaCl solution, except in the 15% deformation where low pits count and large average pit sizes were reported (Szklarska-Smialowska, 1971). Forchhammer and Engell (1969) studied the effect of CD on the pitting corrosion behavior of different austenitic SSs, i.e. 18Cr-10Ni, 25Cr-10Ni, 17Cr-10Ni-2.4Mo, and 16Cr-14Ni-Mo. They established that in 30% NaCl solution, pitting potential was not greatly affected by CD, but for the deformed specimens, the number of pits was higher, and the pits were smaller. According to Randak and Trautes (1970) the quasi-martensite produced in 18Cr-8Ni steels by cold working does not change the pitting potential value in chloride solutions, but pit density increases with increasing CD level.

It is well established that the main factor that is responsible for these degradations in austenitic SS is the martensitic transformation of γ (fcc) to α (bcc) by deformation or irradiation (Niffenegger & Leber, 2008; Xu et al., 2017). In the past decades, the effect of CD on pitting corrosion resistance of SSs was extensively studied. Almost all the researchers suggested that severe CD (beyond 20% cold work) would result in a sharp decrease in pitting corrosion resistance for conventional SSs. Additionally, it can be concluded that the deformation microstructures including deformation bands (or dislocation pile-ups) (Mudali et al., 2002; Peguet et al., 2007), martensite formation (Barbucci et al., 2002; Hamdy et al., 2007), or strain-induced residual stress (Barbucci et al., 2002; Hamdy et al., 2007; Kumar et al., 2007) would weaken the pitting corrosion resistance.

It was established that pitting and crevice corrosion resistances of ferritic SSs were not changed by cold working up to 20% in 0.5-m NaCl solution irrespective of the defect structure containing deformation bands (Mudali et al., 2002). Alternatively, it was also found that corrosion resistance properties of 316L and 316LVM steels increased with an increase in CD up to 20% deformation in simulated body fluid (Talha et al., 2014b). The calculated corrosion rates obtained from the weight loss study of these steels after 24 weeks of immersion in Hank’s and 3.5 wt.% NaCl solutions demonstrated that corrosion rate decreased with increasing degree of CD in both the solutions. Figure 5 shows the polarization curves for annealed and deformed 316L SSs in Hank’s solution at 37°C. It was established that the values of corrosion current decreased with increasing CD up to 20%, indicative of greater protection of the passive films as a result of CD (Talha et al., 2014b). One important requirement for selecting an implant for biomedical application is breakdown potential (Ebd), i.e. the potential at which the passive film breaks and the anodic current significantly increases. The Ebd was also found higher for CD samples than for annealed samples and increased with the degree of deformation. Figure 6 shows the scanning electron microscopy (SEM) micrographs of the different corroded surfaces of SSs (annealed and cold work) after the polarization tests in Hank’s solution (Talha et al., 2014b). Pits were found on annealed as well as on cold deformed steel surfaces, but shallower and smaller pits were associated with deformed steels (up to 20%) compared to annealed steels. Similar results were also observed for HNSs (Navai, 1995). Figure 7 illustrates the change in critical pitting potential with the degree of CD for different austenitic SSs in neutral chloride medium at room temperature (Mudali et al., 2002). It is clear from the figure that the critical pitting potential increased up to 20% CD, beyond which the values decreased for 30% and 40% cold working for the 316L and two more nitrogen-containing steels. A potentiodynamic anodic polarization study by Mudali et al. (2002) also revealed that CD up to 20% enhanced the pitting resistance, and thereafter, a sudden decrease in pitting resistance was noticed for these steels. Pit initiation is closely related to passive film stability and to non-metallic inclusions acting as triggers for future pits. The XPS studies by Talha et al. (2014b) indicated that the Cr oxide:Cr hydroxide ratio and the Fe oxide:Fe hydroxide ratio were higher for deformed (up to 20%) 316LVM SS than for the annealed steel, indicating that the passive films on deformed surfaces are more stable and protective, which improves the corrosion resistance. Phadnis et al. (2003) also attributed the enhanced passivation characteristics of rolled 304 SS to the occurrence of a thicker passive film with a higher Cr/Fe ratio.

Figure 5: 
						Polarization curves for stainless steels in Hank’s solution at 37°C for 316L SS: (A) annealed, (B) 10% C.W., (C) 20% C.W. (Talha et al., 2014b).
Figure 5:

Polarization curves for stainless steels in Hank’s solution at 37°C for 316L SS: (A) annealed, (B) 10% C.W., (C) 20% C.W. (Talha et al., 2014b).

Figure 6: 
						SEM micrographs of SSs after the polarization tests in Hank’s solution: (A) 316L (annealed), (B) 316L (10% C.W.), (C) 316L (20% C.W.), (D) 316LVM (annealed), (E) 316LVM (10% C.W.), (F) 316LVM (20% C.W.) (Talha et al., 2014b).
Figure 6:

SEM micrographs of SSs after the polarization tests in Hank’s solution: (A) 316L (annealed), (B) 316L (10% C.W.), (C) 316L (20% C.W.), (D) 316LVM (annealed), (E) 316LVM (10% C.W.), (F) 316LVM (20% C.W.) (Talha et al., 2014b).

Figure 7: 
						Critical pitting potential values for the alloys (A, B, and C) against cold working levels: (A) N=0 in 316L, (B) 0.1 wt.% in 316LN, and (C) 0.22 wt.% in 36LHN (Mudali et al., 2002), reproduced with permission from Elsevier).
Figure 7:

Critical pitting potential values for the alloys (A, B, and C) against cold working levels: (A) N=0 in 316L, (B) 0.1 wt.% in 316LN, and (C) 0.22 wt.% in 36LHN (Mudali et al., 2002), reproduced with permission from Elsevier).

The effect of deformation on the corrosion of SS is not similar in acid solution than in neutral simulated body fluid or neutral chloride solution. It is prominent that rolling results in a preferential texture of high-index, low-density planes along the rolling direction of austenitic SSs. The effect of this texture results reduced the binding energy of the atoms in this less closed-packed plane, which can result in enhanced dissolution in an acidic chloride medium, where the metal dissolution step plays a main role in the rate control mechanism. This is not the case in a neutral chloride solution, where oxygen diffusion is the rate-controlling step. Enhanced surface diffusion could have contributed to the stable passive film formation at a low degree of cold working, leading to enhancement in pitting resistance (Talha et al., 2014b). Additionally, the chromium content in the passive film formed on a deformed steel surface is richer than that formed on heat-treated samples, and the higher chromium content improved the corrosion resistance of passive films (Phadnis et al., 2003). The results up to now demonstrated that a small amount of CD is significant to the corrosion resistance, but a high deformation is damaging in a neutral solution containing chloride ions (Fu et al., 2009). This is attributed to the fact that the formation of a significant amount of deformation twins because of a higher level of CD in the alloys, which show a different potential than the matrix (Glavatskaya, 1999) and the formation of strain-induced α′ martensite phase as discussed above. A higher degree of CD introduced a high defect density into the matrix, resulting in a less protective passive film as a result of reduced corrosion resistance for heavily cold-worked SS (Fu et al., 2009). This can be illustrated as shown in Figure 8.

Figure 8: 
						Schematic illustration of the stability of passive film on surface and martensite formation in SSs during CD.
Figure 8:

Schematic illustration of the stability of passive film on surface and martensite formation in SSs during CD.

For HNSs, the findings are even more interesting. It is well known that HNSs are more corrosion resistant due to the presence of nitrogen. Actually, rigorous CD significantly weakened the pitting corrosion resistance of HNS with relatively low nitrogen content (Fu et al., 2008). However, for HNS with high nitrogen content, the detrimental effect was found to be mitigated (Ren et al., 2016). Recently, Wang et al. (2017) surprisingly found that the detrimental effect of cold working (up to 60%) could be completely eliminated for HNS with very high nitrogen content (0.92 wt.%). This means that HNS could acquire the ability to heal the damage of passive film caused by cold working. Nitrogen enrichment at the oxide/metal interface was found by Olsson (1995). Although nitrogen is believed to play a significant role for the exceptional corrosion resistance of HNS, the form of nitrogen in passive film remains controversial (Baba et al., 2002; Ningshen et al., 2007). Wang et al. also found that severe CD significantly weakened the stability of passive film on the surface of HNS with 0.76 wt.% N, but had almost no detrimental effect on steels having 0.92 wt.% N. After severe CD, both passive films on steels having 0.76 and 0.92 still retained their special structures. Severe CD had a very strong influence on the enrichment layers of passive film on steel having 0.76 wt.% nitrogen, but almost had no influence on that of passive film on steels with 0.92 wt.% N. According to Wang et al. (2017), the doped nitrogen concentration in the passive film on 0.92 wt.% N was higher than that on 0.76 wt.% N. Severe CD significantly decreased the doped nitrogen concentration in the passive film on steel with 0.76 wt.% N, but this doped nitrogen concentration was not influenced in passive film on steel with 0.92 wt.% N. In other words, the doping effect of nitrogen enrichment in 0.92 N containing steel was high enough to compensate the detrimental effect of the severely deformed microstructure on passive film and could maintain its excellent pitting corrosion resistance at severely cold-worked state. Twins and deformation bands are their main structures, and no martensite was detected from X-ray diffraction (XRD) and EBSD studies in cold-worked microstructures of steels having 0.76 N and 0.92 N. Also, no difference in the behavior of inclusions was found between these two steels at cold-worked state. They conclude on the basis of previous study that the deformation bands (Mudali et al., 2002; Peguet et al., 2007) and residual stress (Barbucci et al., 2002; Hamdy et al., 2007) might be the main reason for the decay of pitting corrosion resistance for severely CD steel with 0.76 wt.% N as discussed above.

One can conclude from the above discussion that CD has no detrimental effect up to 20% CD, and it would result in a sharp decrease in pitting corrosion resistance for conventional SSs beyond 20% cold work in neutral chloride solutions, but it decreases regularly in acid solution even at slight cold work state. This effect is not obtained in steels having a very high content of nitrogen in neutral chloride solution. The main reason of the decreasing corrosion resistance is the formation of martensite phase, deformation bands, and residual stress after severe cold working. The work done by previous researchers and their outputs associated to the effect of cold working on the corrosion behavior of SSs in solutions containing chloride ions is summarized in Table 2.

Table 2:

Summary of the available experimental work on the effect of CD on corrosion behavior of austenitic stainless steels in solution containing chloride ions.

Alloys Experimental conditions
Methods Findings
Reference
Medium C.W. (%) pH/temp (°C) Mech. Prop. Corr. res. Others
HNS 3.5% NaCl 10–60 25 Electrochemical, XPS, SIMS analysis Steel with 0.92 wt.% N has the self-healing property for pitting corrosion resistance at higher degree of deformation. A self-healing reaction involving the nitrogen doping in passive film (Wang et al., 2018)
HNS 3.5% NaCl 10–60 RT Potentiodynamic polarization, SEM, TEM, SIMS analysis ↑ up to 20% cold work Steels with higher nitrogen content are almost not influenced by cold working (Wang et al., 2017)
00Cr18Mn15Mo2N0.86 SS 0.9% saline solution 10–50 Electrochemical methods Unchanged Higher nitrogen content can reduce the negative effect of CD on pitting corrosion (Ren et al., 2016)
AISI 316L, 316LVM SS, Ni-free HNS SBF 10, 20 7/37 Weight loss method, EIS, potentiodynamic polarization No evidence of martensite formation or any other secondary phases (Talha et al., 2014a)
AISI 316L, 316LVM SS SBF 10, 20 7/37 Weight loss method, EIS, potentiodynamic polarization, XPS (Talha et al., 2014b)
AISI 316L SS PBS 17–47 RT Electrochemical methods Sandblasting enhances the surface microhardness but decreases the corrosion resistance (Arifvianto et al., 2012)
Ni-free HNS 0.5 m H2 SO4+0.5 m NaCl, 3.5% NaCl and 0.5 m NaOH+0.5 m NaCl 8, 30, 49, 60 0.4, 5.8, 13.1 EIS, potentiodynamic polarization, TEM, XPS Reduction in corrosion resistance by cold work was not obvious in a 0.5 m H2SO4+0.5 m NaCl. Sensitized HNSSs showed more pits on the corroded surfaces in comparison with non-sensitized ones (Fu et al., 2009)
Ni-free HNS 3.5% NaCl, 10% FeCl3 8–60 5.8/RT Immersion test, EIS, potentiodynamic polarization, TEM, XPS The specimen surfaces after polarization tests showed no obvious changes on the pit size and distribution with increasing cold work level (Fu et al., 2008)
Nb containing austenitic SS 3.5 wt.% NaCl 23, 40, 50 RT EIS, polarization, SEM, EDS ↑ up to 23% C.W. Corrosion resistance increased by increasing the Nb content (Hamdy et al. 2007)
AISI 304, 430 SS 0.5 m NaCl, 10, 20, 30, 70 6.6/23 Electrochemical method, TEM The pit initiation frequency shows a maximum after 20% cold-rolling reduction or 10% tensile deformation (Peguet et al. 2007)
AISI 304 Na2SO4 0.3%+Cl (103 to 5×104) 35, 58 Electrochemical method ↑ (low Cl)

↓ (high Cl)
(Barbucci et al., 2002)
316L 0.5 m NaCl 5, 10, 15, 20, 30, 40 7/RT Electrochemical, SEM, TEM ↑ up to 20% C.W. than ↓ (Mudali et al. 2002)
301 SS 1 m H2SO4 with different Cl- ions 27, 45, 60 RT Electrochemical measurements, TEM (Barbucci et al., 2001)
AISI 304 3%, 5% NaCl 66 Electrochemical methods (Dean and Stimming, 1987)
  1. C.W., Cold work; Mech. Prop., mechanical properties; Corr. Res., corrosion resistance; SBF, simulated body fluid; EIS, electrochemical impedance spectroscopy; TEM, transmission electron microscopy; EDS, energy-dispersive X-ray spectroscopy; SIMS, secondary ion mass spectroscopy; PBS, phosphate buffer saline; RT, room temperature; ↑, increase; ↓, decrease.

8 Conclusions

The austenitic alloys possess an excellent balance of strength and ductility as well as the high ability to again increase their strength during CD. There is no detrimental effect up to 20% CD, and it would result in a sharp decrease in pitting corrosion resistance for austenitic SSs beyond 20% cold work in body fluids or in neutral chloride solutions. A low degree of CD generally do not result in the formation of α′ martensite, but it enhances the surface diffusion, which results in the formation of a stable passive film leading to the improvement in pitting corrosion resistance in neutral solutions. Furthermore, the chromium content in the passive film formed on a low deformed austenitic SS surface is richer in Cr, and chromium content makes surface films more stable, which improves the corrosion resistance performance.

However, severe CD results in the formation of strain-induced martensite, which decreases the localized corrosion resistance by increasing the number of active anodic sites. The formation of deformation twins is also significant for highly deformed samples, which have a different potential than the matrix. The corrosion potential of the martensite phase is more negative than that of austenite, and this is the chief reason for the selective corrosion of martensite. Additionally, a higher degree of CD introduced a high defect density into the matrix, resulting in a less protective passive film as well as a reduced corrosion resistance for the heavily deformed SS. This effect is not obtained in steels having a very high content of nitrogen, and the detrimental effect was found to be mitigated in these steels. However, more work is required on the effect of CD on fatigue and SCC behavior of SSs especially for high-nitrogen steels and other metallic implants for long-term biomedical applications.

  1. Funding: Mohd Talha, is pleased to acknowledge the financial assistance from China Post Doctoral Science Foundation (grant no. 2017M623058).

  2. Conflict of interest: There are no conflicts to declare.

References

Agrawal CM. Reconstructing the human body using biomaterials. J Miner Met Mater Soc 1998; 50: 31–35.10.1007/s11837-998-0064-5Suche in Google Scholar

Alberts B, Bray D, Lewis J, Raff M, Roberts K, Watson JD. Molecular biology of the cell. Tokyo: Kyoiku-sha, 1990.Suche in Google Scholar

Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P. The cell cycle and programmed cell death. In: Anderson M, Granum S, editors. Molecular biology of the cell, 4th ed. New York: Garland Science/Taylor & Francis Group, 2002: 1015.Suche in Google Scholar

Arifvianto SB, Widodo TD, Mahardika M, Dewo P, Salim UA. Effect of cold working and sandblasting on the microhardness, tensile strength and corrosion resistance of AISI 316L stainless steel. Int J Miner Metall Mater 2012; 19: 1093–1099.10.1007/s12613-012-0676-1Suche in Google Scholar

Aoyagi S, Hiromoto S, Hanawa T, Kudo M. TOF–SIMS investigation of metallic material surface after culturing cells. Appl Surf Sci 2004; 231–232: 470–474.10.1016/j.apsusc.2004.03.181Suche in Google Scholar

Azevedo CRF, Hippert Jr E. Failure analysis of surgical implants in Brazil. Eng Fail Anal 2002; 9: 621–633.10.1016/S1350-6307(02)00026-2Suche in Google Scholar

Baba H, Kodama T, Katada Y. Role of nitrogen on the corrosion behavior of austenitic stainless steels. Corros Sci 2002; 44: 2393–2407.10.1016/S0010-938X(02)00040-9Suche in Google Scholar

Balkany TJ, Hodges AV, Eshraghi AA, Butts S, Bricker K, Lingvai J, Polak M, King J. Cochlear implants in children – a review. Acta Otolaryngol 2002; 122: 356–362.10.1080/00016480260000012Suche in Google Scholar PubMed

Barbucci A, Delucchi M, Panizza M, Sacco M, Cerisola G. Electrochemical and corrosion behaviour of cold rolled AISI 301 in 1M H2SO4. J Alloys Compd 2001; 317–318: 607–611.10.1016/S0925-8388(00)01396-7Suche in Google Scholar

Barbucci A, Cerisola G, Cabot PL. Effect of cold working in the passive behavior of 304 stainless steel in sulfate media. J Electrochem Soc 2002; 149: B534–B542.10.1149/1.1516774Suche in Google Scholar

Barceloux DG. Chromium. J Toxicol Clin Toxicol 1999; 37: 173–194.10.1081/CLT-100102418Suche in Google Scholar

Bernhardsson S. The corrosion resistance of duplex stainless steels. In: Charles J, Bernhardsson S, editors. Proceedings of Conference on Duplex Stainless Steels 91, Les Editions de Physique, Les Vlis, France, 1991: 185–210.Suche in Google Scholar

Bhav Singh B, Sivakumar K, Balakrishna Bhat T. Effect of cold rolling on mechanical properties and ballistic performance of nitrogen-alloyed austenitic steels. Int J Impact Eng 2009; 36: 611–620.10.1016/j.ijimpeng.2008.07.082Suche in Google Scholar

Bombara G, Cavallini M. Stress corrosion cracking of bone implants. Corros Sci 1977; 17: 77–85.10.1016/0010-938X(77)90010-5Suche in Google Scholar

Brown SA, Flemming CA, Kawalec JS, Placko HE, Vassaux C, Merritt K, Payer JH, Kraay MJ. Fretting corrosion accelerates crevice corrosion of modular hip tapers. J Appl Biomater 1995; 6: 19–26.10.1002/jab.770060104Suche in Google Scholar PubMed

Browne M, Gregson PJ. Surface modification of titanium alloy implants. Biomaterials 1994; 15: 894–898.10.1016/0142-9612(94)90113-9Suche in Google Scholar PubMed

Buhagiar J, Spiteri A, Sacco M, Sinagra E, Dong H. Augmentation of crevice corrosion resistance of medical grade 316LVM stainless steel by plasma carburizing. Corros Sci 2012; 59: 169–178.10.1016/j.corsci.2012.02.023Suche in Google Scholar

Burstein GT, Pistorius PC. Surface roughness and the metastable pitting of stainless steel in chloride solutions. Corrosion 1995; 51: 380–385.10.5006/1.3293603Suche in Google Scholar

Burstein GT, Vines SP. Repetitive nucleation of corrosion pits on stainless steel and the effects of surface roughness. J Electrochem Soc 2001; 148: B504–B516.10.1149/1.1416503Suche in Google Scholar

Burstein GT, Liu C, Souto RM, Vines SP. Origins of pitting corrosion. Corros Eng Sci Technol 2004; 39: 25–30.10.1179/147842204225016859Suche in Google Scholar

Carbonaro RF, Gray BN, Whitehead CF, Stone AT. Carboxylate-containing chelating agent interactions with amorphous chromium hydroxide: adsorption and dissolution. Geochim Cosmochim Acta 2008; 72: 3241–3257.10.1016/j.gca.2008.04.010Suche in Google Scholar

Chao CY, Lin LF, Macdonald DD. A point defect model for anodic passive films I. Film growth kinetics. J Electrochem Soc 1981; 128: 1187–1194.10.1149/1.2127591Suche in Google Scholar

Chaturvedi TP. An overview of the corrosion aspect of dental implants (titanium and its alloys). Ind J Dent Res 2009; 20: 91–98.10.4103/0970-9290.49068Suche in Google Scholar PubMed

Choonara YE, Pillay V, Danckwerts MP, Carmichael TR, du Toit LC. A review of implantable intravitreal drug delivery technologies for the treatment of posterior segment eye diseases. J Pharm Sci. 2010; 99: 2219–2239.10.1002/jps.21987Suche in Google Scholar PubMed

Chu PK, Chen JY, Wang LP, Huang N. Plasma-surface modification of biomaterials. Mater Sci Eng Rep 2002; 36: 143–206.10.1016/S0927-796X(02)00004-9Suche in Google Scholar

Clayton CR, Lu YC. A bipolar model of the passivity of stainless steel: the role of Mo addition. J Electrochem Soc 1986; 133: 2465–2473.10.1149/1.2108451Suche in Google Scholar

Dai K, Ning CQ. Shape memory alloys and their medical application. In: Poitout DG, editor. Biomechanics and biomaterials in orthopedics. Springer: London, 2004: 179–184.10.1007/978-1-4471-3774-0_17Suche in Google Scholar

Davis JR. ASM Speciality Handbook, Stainless Steel. Materials Park, OH: ASM International, 1994: 22.Suche in Google Scholar

Dayan AD, Paine AJ. Mechanisms of chromium toxicity, carcinogenicity and allergenicity: review of the literature from 1985 to 2000. Hum Exp Toxicol 2001; 20: 439–451.10.1191/096032701682693062Suche in Google Scholar PubMed

Dean MH, Stimming V. Capacity of semiconductor electrodes with multiple bulk electronic states. Part I. Model and calculations for discrete states. J Electroanal Chem 1987; 228: 135–151.10.1016/0022-0728(87)80103-1Suche in Google Scholar

Disegi JA, Eschbach L. Stainless steel in bone surgery. Injury 2000; 31: 4–6.10.1016/S0020-1383(00)80015-7Suche in Google Scholar PubMed

Drogowska M, Menard H, Lasia A, Brossard L. Impedance study of the passive film on stainless steel 304 in pH 8 carbonate solution. J Appl Electrochem 1996; 26: 1169–1177.10.1007/BF00243742Suche in Google Scholar

Dwivedi D, Lepkova K, Becker T. Carbon steel corrosion: a review of key surface properties and characterization methods. RSC Adv 2017; 7: 4580–4610.10.1039/C6RA25094GSuche in Google Scholar

Elayaperumal K, De PK, Balachandra J. Passivity of type 304 stainless steel-effect of plastic deformation. Corrosion 1972; 28: 269–273.10.5006/0010-9312-28.7.269Suche in Google Scholar

Eliades T, Pratsinis H, Kletsas D, Eliades G, Makou M. Characterization and cytotoxicity of ions released from stainless steel and nickel-titanium orthodontic alloys. Am J Orthod Dentofacial Orthop 2004; 125: 24–29.10.1016/j.ajodo.2003.09.009Suche in Google Scholar

Eliaz N. Biomaterials and corrosion. In: Mudali UK, Raj B, editors. Corrosion science and technology: mechanism, mitigation and monitoring. New Delhi: Narosa Publishing House, 2008: 356–397; USA: CRC Press and Europe: Alpha Science International, 2009.Suche in Google Scholar

Fang Z, Wu YS, Zhang L, Li J. Effect of deformation induced martensite on electrochemical behaviors of type 304 stainless steel in the active state. Corros Sci Prot Technol 1997; 9: 75–83.Suche in Google Scholar

Feichtinger HK. Concepts of nitrogen solubility. In: Kiew, editor. High nitrogen steels, HNS 93, proceeding of third international conference. Institute of Metal Physics, Kiev, Academy of Science of Ukraine, 1993: 45–54.Suche in Google Scholar

Feichtinger HK, Zheng X. Powder metallurgy of high nitrogen steels. Powder Metall Int 1990; 22: 7–13.Suche in Google Scholar

Feng X, Lu X, Zuo Y, Zhuang N, Chen D. The effect of deformation on metastable pitting of 304 stainless steel in chloride contaminated concrete pore solution. Corros Sci 2016; 103: 223–229.10.1016/j.corsci.2015.11.022Suche in Google Scholar

Forchhammer P, Engell HJ. Investigation into the pitting corrosion of passive austenitic Cr-Ni steels in neutral chloride solutions. Mater Corros 1969; 20: 1–12.10.1002/maco.19690200103Suche in Google Scholar

Frankel G. Pitting corrosion of metals a review of the critical factors. J Electrochem Soc 1998; 145: 2186–2198.10.1149/1.1838615Suche in Google Scholar

Fu Y, Wu XQ, Han EH, Ke W, Yang K, Jiang ZH. Influence of cold work on pitting corrosion behavior of a high nitrogen stainless steel. J Electrochem Soc 2008; 155: C455–C463.10.1149/1.2939213Suche in Google Scholar

Fu Y, Wu XQ, Han EH, Ke W, Yang K, Jiang ZH. Effects of cold work and sensitization treatment on the corrosion resistance of high nitrogen stainless steel in chloride solutions. Electrochim Acta 2009; 54: 1618–1629.10.1016/j.electacta.2008.09.053Suche in Google Scholar

Gaben F, Vuillemin B, Oltra R. Influence of the chemical composition and electronic structure of passive films grown on 316L SS on their transient electrochemical behavior. J Electrochem Soc 2004; 151: B595–B604.10.1149/1.1803562Suche in Google Scholar

Geetha M, Singh AK, Asokamani R, Gogia AK. Ti based biomaterials, the ultimate choice for orthopaedic implants – a review. Prog Mater Sci 2009; 54: 397–425.10.1016/j.pmatsci.2008.06.004Suche in Google Scholar

Geetha M, Durgalakshmi D, Asokamani R. Biomedical implants: corrosion and its prevention – a review. Recent Pat Corros Sci 2010; 2: 40–54.10.2174/1877610801002010040Suche in Google Scholar

Gilbert JL, Zarka L, Chang E, Thomas CH. The reduction half cell in biomaterials corrosion: oxygen diffusion profiles near and cell response to polarized titanium surfaces. J Biomed Mater Res 1998; 42: 321–330.10.1002/(SICI)1097-4636(199811)42:2<321::AID-JBM18>3.3.CO;2-PSuche in Google Scholar

Giordani EJ, Guimarães VA, Pinto TB, Ferreira I. Effect of precipitates on the corrosion fatigue crack initiation of ISSO 5832–9 stainless steel biomaterial. Int J Fatigue 2004; 26: 1129–1136.10.1016/j.ijfatigue.2004.03.002Suche in Google Scholar

Glavatskaya NI. Effect of structure-texture evolution on cold work hardening caused by rolling of nitrogen Cr-Ni-Mn austenitic steels. Mater Sci Forum 1999; 318–320: 799–804.10.4028/www.scientific.net/MSF.318-320.799Suche in Google Scholar

Gontijo LC, Machado R, Kuri SE, Casteletti LC, Nascente PAP. Corrosion resistance of the layers formed on the surface of plasma-nitrided AISI 304L steel. Thin Solid Films 2006; 515: 1093–1096.10.1016/j.tsf.2006.07.075Suche in Google Scholar

Ha HY, Lee TH, Oh CS, Kim SJ. Effects of combined addition of carbon and nitrogen on pitting corrosion behavior of Fe–18Cr–10Mn alloys. Scr Mater 2009a; 61: 121–124.10.1016/j.scriptamat.2009.03.018Suche in Google Scholar

Ha HY, Lee TH, Oh CS, Kim SJ. Effects of carbon on the corrosion behavior in Fe–18Cr–10Mn–N–C stainless steels. Steel Res Int 2009b; 80: 488–492.Suche in Google Scholar

Hamdy AS, El-Shenawy E, El-Bitar T. The corrosion behavior of niobium bearing cold deformed austenitic stainless steels in 3.5% NaCl solution. Mater Lett 2007; 61: 2827–2832.10.1016/j.matlet.2006.10.043Suche in Google Scholar

Hanawa T. Metal ion release from metal implants. Mater Sci Eng C 2004; 24: 745–752.10.1016/j.msec.2004.08.018Suche in Google Scholar

Hanawa T, Hiromoto S, Yamamoto A, Kuroda D, Asami K. XPS characterization of the surface oxide film of 316L stainless samples that were located in quasi-biological environments. Mater Trans 2002; 43: 3088–3092.10.2320/matertrans.43.3088Suche in Google Scholar

Hashimoto K, Asami K, Teramoto K. An X-ray photo-electron spectroscopic study on the role of molybdenum in increasing the corrosion resistance of ferritic stainless steels in HCl. Corros Sci 1979a; 19: 3–14.10.1016/0010-938X(79)90003-9Suche in Google Scholar

Hashimoto K, Naka M, Asami K. An X-ray photo-electron spectroscopy study of the passivity of amorphous Fe–Mo alloys. Corros Sci 1979b; 19: 165–170.10.1016/0010-938X(79)90014-3Suche in Google Scholar

Hayashi M. Cell-adhesive molecules (Saibo-Secchaku-Bunshi-no-Sekai). Tokyo: Yodo-sha, 1995: 11–21.Suche in Google Scholar

Hedberg Y, Hedberg J, Liu Y, Wallinder IO. Complexation- and ligand-induced metal release from 316L particles: importance of particle size and crystallographic structure. Biometals 2011; 24: 1099–1114.10.1007/s10534-011-9469-7Suche in Google Scholar PubMed

Hedberg Y, Wang X, Hedberg J, Lundin M, Blomberg E, Wallinder IO. Surface-protein interactions on different stainless steel grades: effects of protein adsorption, surface changes and metal release. J Mater Sci Mater Med 2013; 24: 1015–1033.10.1007/s10856-013-4859-8Suche in Google Scholar PubMed PubMed Central

Hedberg Y, Karlsson ME, Wei Z, Znidarsic M, Wallinder IO, Hedberg J. Interaction of albumin and fibrinogen with stainless steel: influence of sequential exposure and protein aggregation on metal release and corrosion resistance. Corrosion 2017; 73: 1423–1436.10.5006/2504Suche in Google Scholar

Hench LL, Ethridge EC. Biomaterials: the interfacial problem. Adv Biomed Eng 1975; 5: 35–150.10.1016/B978-0-12-004905-9.50007-4Suche in Google Scholar

Herting G, Wallinder IO, Leygraf C. Metal release from various grades of stainless steel exposed to synthetic body fluids. Corros Sci 2007; 49: 103–111.10.1016/j.corsci.2006.05.008Suche in Google Scholar

Hiromoto S, Hanawa T. pH near cells on stainless steel and titanium. Electrochem Solid State Lett 2004; 7: B9–B11.10.1149/1.1645353Suche in Google Scholar

Hiromoto S, Hanawa T. Electrochemical properties of 316L stainless steel with culturing L929 fibroblasts. J R Soc Interface 2006; 3: 495–505.10.1098/rsif.2005.0108Suche in Google Scholar PubMed PubMed Central

Hiromoto S, Noda K, Hanawa T. Development of electrolytic cell with cell-culture for metallic biomaterials. Corros Sci 2002; 44: 955–965.10.1016/S0010-938X(01)00110-XSuche in Google Scholar

Hryniewicz T. Concept of microsmoothing in the electropolishing process. Surf Coat Technol 1994; 64: 75–80.10.1016/S0257-8972(09)90006-8Suche in Google Scholar

Hou J, Peng QJ, Shoji T, Wang JQ, Han EH, Ke W. Effects of cold working path on strain concentration, grain boundary microstructure and stress corrosion cracking in Alloy 600. Corros Sci 2011; 53: 2956–2962.10.1016/j.corsci.2011.05.037Suche in Google Scholar

Hu J, Zhano ZJ, Li LX. Corrosion fatigue resistance of surgical implant stainless steels and titanium alloys. Corros Sci 1993; 35: 587–597.Suche in Google Scholar

Huang CX, Yang G, Gao YL, Wu SD, Li SX. Investigation on the nucleation mechanism of deformation-induced martensite in an austenitic stainless steel under severe plastic deformation. J Mater Res 2007; 22: 724–729.10.1557/jmr.2007.0094Suche in Google Scholar

Humbeeck JV, Stalmans R, Besselink PA, Shape memory alloys. In: Helsen JA, Brème HJ, editors. Metals as biomaterials, Ed. 1. Chichester: Wiley & Sons, 1998: 73–100.Suche in Google Scholar

IARC monographs on the evaluation of carcinogenic risks to humans. Surgical implants and other foreign bodies. vol. 74. Lyon: World Health Organization (WHO), 1999: 65.Suche in Google Scholar

Jaganathan SK, Supriyanto E, Murugesan S, Balaji A, Asokan MK. Biomaterials in cardiovascular research: applications and clinical implications. BioMed Res Int 2014: 1–11, Article ID: 459465.10.1155/2014/459465Suche in Google Scholar PubMed PubMed Central

Jinlong L, Tongxiang L. Improved corrosion resistance of 316L stainless steel by nanocrystalline and electrochemical nitridation in artificial saliva solution. Appl Surf Sci 2015; 359: 158–165.10.1016/j.apsusc.2015.09.267Suche in Google Scholar

Jones DA. Principles and prevention of corrosion, 2nd ed. New Jersey: Prentice Hall, 1996; 5: 75–115.Suche in Google Scholar

Kamachi MU, Baldev R. Corrosion science and technology: mechanism, mitigation and monitoring. UK: Taylor & Francis, 2008: 283–356.Suche in Google Scholar

Khan MA, Williams RL, Williams DF. Conjoint corrosion and wear in titanium alloys. Biomaterials 1999; 20: 765–772.10.1016/S0142-9612(98)00229-4Suche in Google Scholar PubMed

Kobayashi S, Kobayashi R, Watanabe T. Control of grain boundary connectivity based on fractal analysis for improvement of intergranular corrosion resistance in SUS316L austenitic stainless steel. Acta Mater 2016; 102: 397–405.10.1016/j.actamat.2015.08.075Suche in Google Scholar

Kolotyrkin YM, Freiman LI, Reformatskaya II, Panshin EA. On the mechanism for increasing pitting resistance of stainless steels due to molybdenum adding. Prot Metals Zaschita Metallov 1994; 30: 453–462.Suche in Google Scholar

Kuhn A. The electropolishing of titanium and its alloys. Metal Finishing 2004; 102: 80–86.10.1016/S0026-0576(04)82510-8Suche in Google Scholar

Kumar BR, Mahato B, Singh R. Influence of cold-worked structure on electrochemical properties of austenitic stainless steels. Metall Mater Trans A 2007; 38: 2085–2094.10.1007/s11661-007-9224-4Suche in Google Scholar

Kurella SRA. Laser surface modification of Ti-6Al-4V: wear and corrosion characterization in simulated biofluid. J Biomater Appl 2006; 21: 49–73.10.1177/0885328206055998Suche in Google Scholar PubMed

Lanza RP, Langer R, Vacanti J. Principles of tissue engineering, 2nd ed. California, USA: Academic Press, 1997: 57–71.Suche in Google Scholar

Lin LF, Chao CY, Macdonald DD. A point defect model for anodic passive films II. Chemical breakdown and pit initiation. J Electrochem Soc 1981; 128: 1194–1198.10.1149/1.2127592Suche in Google Scholar

Liu LL, Xu J, Lu X, Munroe P, Xie ZH. Electrochemical corrosion behavior of nanocrystalline β-Ta coating for biomedical applications. ACS Biomater Sci Eng 2016; 2: 579–594.10.1021/acsbiomaterials.5b00552Suche in Google Scholar PubMed

Lula RA. Stainless steel. Metals Park, Ohio: American Society for Metals, 1986: 68.Suche in Google Scholar

Luo H, Su H, Ying G, Dong C, Li X. Effect of cold deformation on the electrochemical behaviour of 304L stainless steel in contaminated sulfuric acid environment. Appl Surf Sci 2017; 425: 628–638.10.1016/j.apsusc.2017.07.057Suche in Google Scholar

Malik A, Kutty PM, Siddiqi NA, Andijani IN, Ahmed S. The influence of pH and chloride concentration on the corrosion behaviour of AISI 316L steel in aqueous solutions. Corros Sci 1992; 33: 1809–1827.10.1016/0010-938X(92)90011-QSuche in Google Scholar

Matsumoto H. Interface phenomena on materials with good biocompatibility. Chem Rev 1978; 21: 35–54.10.2307/412997Suche in Google Scholar

Melchers RE. Pitting corrosion in marine environments: a review. Australia: Department of Civil Engineering and Surveying, University of Newcastle, 1994.Suche in Google Scholar

Merritt K, Brown SA. Effect of proteins and pH on fretting corrosion and metal ion release. J Biomed Mater Res 1988; 22: 111–120.10.1002/jbm.820220204Suche in Google Scholar PubMed

Messer R, Kimbrough D, Austin G. Venugopalan R. In vitro test system combining cell culture and corrosion techniques. In: Proc. Society for Biomaterials 27th Annual Meeting Transactions, Saint Paul, 2001: 221.Suche in Google Scholar

Metikos-Hukovic M, Babic R, Grubac Z, Petrovic Z, Lajçi N. High corrosion resistance of austenitic stainless steel alloyed with nitrogen in an acid solution. Corros Sci 2011; 53: 2176–2183.10.1016/j.corsci.2011.02.039Suche in Google Scholar

Milad M, Zreiba N, Elhalouani F, Baradai C. The effect of cold work on structure and properties of AISI 304 stainless steel. J Mater Process Technol 2008; 203: 80–85.10.1016/j.jmatprotec.2007.09.080Suche in Google Scholar

Milosev I. Effect of complexing agents on the electrochemical behaviour of orthopaedic stainless steel in physiological solution. J Appl Electrochem 2002; 32: 311–320.10.1023/A:1015595420097Suche in Google Scholar

Milosev I, Strehblow HH. The behavior of stainless steels in physiological solution containing complexing agent studied by X-ray photoelectron spectroscopy. J Biomed Mater Res 2000; 52: 404–412.10.1002/1097-4636(200011)52:2<404::AID-JBM22>3.0.CO;2-ZSuche in Google Scholar

Minovic A, Milosev I, Pihlar B. The influence of complexing agent and proteins on the corrosion of stainless steels and their metallic components. J Mater Sci Mater Med 2003; 14: 69–77.10.1023/A:1021505621388Suche in Google Scholar

Moayed MH, Laycock N, Newman R. Dependence of the critical pitting temperature on surface roughness. Corros Sci 2003; 45: 1203–1216.10.1016/S0010-938X(02)00215-9Suche in Google Scholar

Mudali UK, Shankar P, Ningshen S, Dayal RK, Khatak HS, Raj B. On the pitting corrosion resistance of nitrogen alloyed cold worked austenitic stainless steels. Corros Sci 2002; 44: 2183–2198.10.1016/S0010-938X(02)00035-5Suche in Google Scholar

Murata Y, Ohashi S, Uematsu Y. Recent trends in the production and use of high strength stainless steels. ISIJ Int 1993; 33: 711–720.10.2355/isijinternational.33.711Suche in Google Scholar

Mustafa K, Pan J, Wroblewski J, Leygraf C, Arvidson K. Electrochemical impedance spectroscopy and X-ray photoelectron spectroscopy analysis of titanium surfaces cultured with osteoblast-like cells derived from human mandibular bone. J Biomed Mater Res 2002; 59: 655–664.10.1002/jbm.1275Suche in Google Scholar PubMed

Nagarajan S, Mohana M, Sudhagar P, Raman V, Nishimura T, Kim S, Kang YS, Rajendran N. Nanocomposite coatings on biomedical grade stainless steel for improved corrosion resistance and biocompatibility. ACS Appl Mater Interfaces 2012; 4: 5134–5141.10.1021/am301559rSuche in Google Scholar PubMed

Nagy E, Mertinger V, Tranta F, Solyom J. Deformation induced martensitic transformation in stainless steels. Mater Sci Eng A 2004; 378: 308–313.10.1016/j.msea.2003.11.074Suche in Google Scholar

Nakada N, Ito H, Matsuoka Y, Tsuchiyama T, Takaki S. Deformation-induced martensitic transformation behavior in cold-rolled and cold-drawn type 316 stainless steels. Acta Mater 2010; 58: 895–903.10.1016/j.actamat.2009.10.004Suche in Google Scholar

Navai F. Effects of tensile and compressive stresses on the passive layers formed on a type 302 stainless steel in a normal sulphuric acid bath. J Mater Sci 1995; 30: 1166–1172.10.1007/BF00356115Suche in Google Scholar

Navai F, Debbouz O. AES study of passive films formed on a type 316 austenitic stainless-steels in a stress field. J Mater Sci 1999; 34: 1073–1079.10.1023/A:1004556314364Suche in Google Scholar

Niffenegger M, Leber HJ. Sensitivity of the magnetization curves of different austenitic stainless tube and pipe steels to mechanical fatigue. J Nucl Mater 2008; 377: 325–330.10.1016/j.jnucmat.2008.03.007Suche in Google Scholar

Ningshen S, Mudali UK, Mittal VK, Khatak HS. Semiconducting and passive film properties of nitrogen-containing type 316LN stainless steels. Corros Sci 2007; 49: 481–496.10.1016/j.corsci.2006.05.041Suche in Google Scholar

Okazaki Y, Gotoh E. Comparison of metal release from various metallic biomaterials in vitro. Biomaterials 2005; 26: 11–21.10.1016/j.biomaterials.2004.02.005Suche in Google Scholar PubMed

Okazaki Y, Tateishi T, Ito Y. Corrosion resistance of implant alloys in Pseudo physiological solution and role of alloying elements in passive films. Mater Trans JIM 1997; 38: 78–84.10.2320/matertrans1989.38.78Suche in Google Scholar

Olefjord I, Wegrelius L. Surface analysis of passive state. Corros Sci 1990; 31: 89–98.10.1016/0010-938X(90)90095-MSuche in Google Scholar

Olefjord I, Wegrelius L. The influence of nitrogen on the passivation of stainless steels. Corros Sci 1996; 38: 1203–1220.10.1016/0010-938X(96)00018-2Suche in Google Scholar

Olsson COA. The influence of nitrogen and molybdenum on passive films formed on the austenoferritic stainless steel 2205 studied by AES and XPS. Corros Sci 1995; 37: 467–479.10.1016/0010-938X(94)00148-YSuche in Google Scholar

Omanovic S, Roscoe SG. Electrochemical studies of the adsorption behavior of bovine serum albumin on stainless steel. Langmuir 1999; 15: 8315–8321.10.1021/la990474fSuche in Google Scholar

Opiela M, Grajcar A, Krukiewicz W. Corrosion behaviour of Fe-Mn-Si-Al austenitic steel in chloride solution. JAMME 2009; 33: 159–165.Suche in Google Scholar

Oyane A, Kim HM, Furya T, Kokubo T, Miyazaki T, Nakamura T. Preparation and assessment of revised simulated body fluids. J Biomed Mater Res 2003; 65A: 188–195.10.1002/jbm.a.10482Suche in Google Scholar PubMed

Panda B, Sujata M, Madan M, Bhaumik SK. Stress corrosion cracking in 316L stainless steel bellows of a pressure safety valve. Eng Fail Anal 2014; 36: 379–389.10.1016/j.engfailanal.2013.11.007Suche in Google Scholar

Pardo A, Otero E, Merino M, Lopez M, Utrilla M, Moreno F. Influence of pH and chloride concentration on the pitting and crevice corrosion behavior of high-alloy stainless steels. Corrosion 2000; 56: 411–418.10.5006/1.3280545Suche in Google Scholar

Pardo A, Merino MC, Coy AE, Viejo F, Arrabal R, Matykina E. Pitting corrosion behaviour of austenitic stainless steels – combining effects of Mn and Mo additions. Corros Sci 2008; 50: 1796–1806.10.1016/j.corsci.2008.04.005Suche in Google Scholar

Park JB. Biomaterials science and engineering. Plenum, New York: Wiley-Liss, 1984: 193–233.10.1007/978-1-4613-2769-1_8Suche in Google Scholar

Peguet L, Malki B, Baroux B. Influence of cold working on the pitting corrosion resistance of stainless steels. Corros Sci 2007; 49: 1933–1948.10.1016/j.corsci.2006.08.021Suche in Google Scholar

Phadnis SV, Satpati AK, Muthe KP, Vyas JC, Sundaresan RI. Comparison of rolled and heat treated SS 304 in chloride solution using electrochemical and XPS techniques. Corros Sci 2003; 45: 2467–2483.10.1016/S0010-938X(03)00099-4Suche in Google Scholar

Pound BG. Corrosion behavior of metallic materials in biomedical applications. II. Stainless steels and Co-Cr alloys. Corr Rev 2014; 32: 21–41.10.1515/corrrev-2014-0008Suche in Google Scholar

Probster L, Dent M, Lin WL, Huttenmann H. Effect of fluoride prophylactic agents on titanium surfaces. Int J Oral Max Impl 1992; 7: 390–394.Suche in Google Scholar

Raman RKS, Siew WH. Stress corrosion cracking of an austenitic stainless steel in nitrite-containing chloride solutions. Materials 2014; 7: 7799–7808.10.3390/ma7127799Suche in Google Scholar PubMed PubMed Central

Randak VA, Trautes FW. On the influence of the austenite stability of 18/8-chromium-nickel steels on the deformation properties and on the corrosion behavior of these steels. Mater Corros 1970; 21: 97–109.10.1002/maco.19700210206Suche in Google Scholar

Ratner BD, Hoffman AS, Schoen FJ, Lemons JE. Biomaterials science: an introduction to materials in medicine. San Diego: Elsevier/Academic Press, 2004.Suche in Google Scholar

Ren Y, Zhao H, Liu W, Yang K. Effect of cold deformation on pitting corrosion of 00Cr18Mn15Mo2N0.86 stainless steel for coronary stent application. Mater Sci Eng C 2016; 60: 293–297.10.1016/j.msec.2015.11.048Suche in Google Scholar PubMed

Richardsa J, Cremers C, Fischer P, Schmidt K. Corrosion studies on electro polished stainless steels for the use as metallic bipolar plates in PEMFC applications. Energy Procedia 2012; 20: 324–333.10.1016/j.egypro.2012.03.032Suche in Google Scholar

Roffey P. Case study: failure of high nitrogen stainless steel femoral stem. Eng Fail Anal 2012; 20: 173–182.10.1016/j.engfailanal.2011.11.010Suche in Google Scholar

Roger N. Metals Handbook, 26, American Society for Metals. Metals Park, Ohio, 1985: 38.Suche in Google Scholar

Rudee M, Price TM. The initial stages of adsorption of plasma derived proteins on artificial surfaces in a controlled flow environment. J Biomed Mater Res 1985; 19: 57–66.10.1002/jbm.820190106Suche in Google Scholar PubMed

Ryhanen J. Biocompatibility evolution of nickel–titanium shape memory alloy, Academic Dissertation, Faculty of Medicine, University of Oulu, Finland, 1999.Suche in Google Scholar

Saadi SA, Yi YS, Cho PY, Jang CH, Beeley P. Passivity breakdown of 316Lstainless steel during potentiodynamic polarization in NaCl solution. Corros Sci 2016; 111: 720–727.10.1016/j.corsci.2016.06.011Suche in Google Scholar

Saker A, Leroy Ch, Michel H, Frantz C. Properties of sputtered stainless steel nitrogen coatings and structural analogy with low temperature plasma nitrided layers of austenitic steels. Mater Sci Eng A 1991; 140: 702–708.10.1016/0921-5093(91)90500-MSuche in Google Scholar

Sato K, Ichinose M, Hirotsu Y, Inoue Y. Effects of deformation induced phase transformation and twinning on the mechanical properties of austenitic Fe–Mn–Al alloys. ISIJ Int 1989; 29: 868–877.10.2355/isijinternational.29.868Suche in Google Scholar

Schmidt C, Ignatius A, Claes LE. Proliferation and differentiation parameters of human osteoblasts on titanium and steel surfaces. J Biomed Mater Res 2001; 54: 209–215.10.1002/1097-4636(200102)54:2<209::AID-JBM7>3.0.CO;2-7Suche in Google Scholar

Sedriks AJ. In: Proceedings of the International Conference on Stainless Steels 85. London: The Institute of Metals, 1985: 125.Suche in Google Scholar

Shahryari A, Omanovic S, Szpunar JA. Electrochemical formation of highly pitting resistant passive films on a biomedical grade 316LVM stainless steel surface. Mater Sci Eng C 2008; 28: 94–106.10.1016/j.msec.2007.09.002Suche in Google Scholar

Shaw BA, Kelly RG. What is corrosion? Electrochem Soc Interface 2006; 15: 24–26.10.1149/2.F06061IFSuche in Google Scholar

Shimada M, Kokawa H, Wang Z, Sato Y, Karibe I. Optimization of grain boundary character distribution for intergranular corrosion resistant 304 stainless steel by twin-induced grain boundary engineering. Acta Mater 2002; 50: 2331–2341.10.1016/S1359-6454(02)00064-2Suche in Google Scholar

Shimohira S. Materials science of corrosion and corrosion protection. Fusyoku-Bousyoku-no-Zairyou-Kagaku, Agune Gijyutu Center, Tokyo, 1995: 213.Suche in Google Scholar

Shirdel M, Mirzadeh H, Parsa MH. Nano/ultrafine grained austenitic stainless steel through the formation and reversion of deformation-induced martensite: mechanisms, microstructures, mechanical properties, and TRIP effect. Mater Charact 2015; 103: 150–161.10.1016/j.matchar.2015.03.031Suche in Google Scholar

Sidharth AAP. Effect of pitting corrosion on ultimate strength and buckling strength of plate – a review. Digest J Nanomater Biostruct 2009; 4: 783–788.Suche in Google Scholar

Sikora E, Sikora J, Macdonald DD. A new method for estimating the diffusivities of vacancies in passive films. Electrochim Acta 1996; 41: 783–789.10.1016/0013-4686(95)00312-6Suche in Google Scholar

Sivakumar M, Rajeswari S. Investigation of failures in stainless steel orthopaedic implant devices: pit-induced stress corrosion cracking. J Mater Sci Lett 1992; 11: 1039–1042.10.1007/BF00729754Suche in Google Scholar

Sivakumar M, Kumar S, Dhanadurai K, Rajeswari S, Thulasiraman V. Failures in stainless steel orthopaedic implant devices: a survey. J Mater Sci Lett 1995; 14: 351–354.10.1007/BF00592147Suche in Google Scholar

Solomon N, Solomon I. Effect of deformation-induced phase transformation on AISI 316 stainless steel corrosion resistance. Eng Fail Anal 2017; 79: 865–875.10.1016/j.engfailanal.2017.05.031Suche in Google Scholar

Speidel MO. Nitrogen containing austenitic stainless steels. Mat-wiss. u. Werkstoiftech 2006; 37: 875–880.10.1002/mawe.200600068Suche in Google Scholar

Stein G, Menzel J, Choudhary A. Industrial manufacture of massive nitrogen alloyed steels. In: Special melting and processing technologies, San Diego, New Jersey: Noyes Publications, 1988.Suche in Google Scholar

Steinemann SG. Metal implants and surface reactions. Injury 1996; 27: SC16–SC22.10.1016/0020-1383(96)89027-9Suche in Google Scholar PubMed

Stewart J, Williams DE. The initiation of pitting corrosion on austenitic stainless steel: on the role and importance of sulphide inclusions. Corros Sci 1992; 33: 457–474.10.1016/0010-938X(92)90074-DSuche in Google Scholar

Sudarshan TS, Srivatsan TS, Harvey IIDP. Fatigue processes in metals – role of aqueous environments. Eng Fract Mech 1990; 36: 827–852.10.1016/0013-7944(90)90262-FSuche in Google Scholar

Summer B, Fink U, Zeller R, Rueff F, Maier S, Roider G, Thomas P. Patch test reactivity to a cobalt–chromium–molybdenum alloy and stainless steel in metal-allergic patients in correlation to the metal ion release. Contact Dermatitis 2007; 57: 35–39.10.1111/j.1600-0536.2007.01139.xSuche in Google Scholar PubMed

Sumita M, Hanawa T, Teoh SH. Development of nitrogen-containing nickel-free austenitic stainless steels for metallic biomaterials – review. Mater Sci Eng C 2004; 24: 753–760.10.1016/j.msec.2004.08.030Suche in Google Scholar

Suresh S. Fatigue of materials, 2nd ed., Cambridge: Cambridge University Press, 2004.Suche in Google Scholar

Syrett BC, Wing SS. An electrochemical investigation of fretting corrosion of surgical implant materials. Corrosion 1978; 11: 379–386.10.5006/0010-9312-34.11.378Suche in Google Scholar

Szklarska-Smialowska Z. Review of literature on pitting corrosion. Corrosion 1971; 27: 223–233.10.5006/0010-9312-27.6.223Suche in Google Scholar

Takaki S, Fukunaga K, Syarif J, Tsuchiyama T. Effect of grain refinement on thermal stability of metastable austenitic steel. Mater Trans 2004; 45: 2245–2251.10.2320/matertrans.45.2245Suche in Google Scholar

Talha M, Behera CK, Sinha OP. A review on nickel-free nitrogen containing austenitic stainless steels for biomedical applications. Mater Sci Eng C 2013; 33: 3563–3575.10.1016/j.msec.2013.06.002Suche in Google Scholar PubMed

Talha M, Behera CK, Kumar S, Pal O, Singh G, Sinha OP. Long term and electrochemical corrosion investigation of cold worked AISI 316L and 316LVM stainless steels in simulated body fluid. RSC Adv 2014a; 4: 13340–13349.10.1039/c3ra47881eSuche in Google Scholar

Talha M, Behera CK, Sinha OP. In-vitro long term and electrochemical corrosion resistance of cold deformed nitrogen containing austenitic stainless steels in simulated body fluid. Mater Sci Eng C 2014b; 40: 455–466.10.1016/j.msec.2014.04.006Suche in Google Scholar PubMed

Talha M, Behera CK, Sinha OP. Effect of nitrogen and cold working on structural and mechanical behavior of Ni-free nitrogen containing austenitic stainless steels for biomedical applications. Mater Sci Eng C 2015; 47: 196–203.10.1016/j.msec.2014.10.078Suche in Google Scholar PubMed

Tavares SSM, Gunderov D, Stolyarov V, Neto JM. Phase transformation induced by severe plastic deformation in the AISI 304L stainless steel. Mater Sci Eng A 2003; 358: 32–36.10.1016/S0921-5093(03)00263-6Suche in Google Scholar

Virtanen S, Milosev I, Gomez-Barrena E, Trebse R, Salo J, Konttinen YT. Special modes of corrosion under physiological and simulated physiological conditions. Acta Biomaterialia 2008; 4: 468–476.10.1016/j.actbio.2007.12.003Suche in Google Scholar PubMed

Wang Q, Zhang B, Ren Y, Yang K. Eliminating detrimental effect of cold working on pitting corrosion resistance in high nitrogen austenitic stainless steels. Corros Sci 2017; 123: 351–355.10.1016/j.corsci.2017.04.006Suche in Google Scholar

Wang Q, Zhang B, Ren Y, Yang K. A self-healing stainless steel: role of nitrogen in eliminating detrimental effect of cold working on pitting corrosion resistance. Corros Sci 2018; 145: 55–66.10.1016/j.corsci.2018.09.013Suche in Google Scholar

Wei R, Vajo JJ, Matossian JN, Wilbur PJ, Davis JA, Williamson DL, Collins GA. A comparative study of beam ion implantation, plasma ion implantation and nitriding of AISI 304 stainless steel. Surf Coat Technol 1996; 83: 235–242.10.1016/0257-8972(95)02825-0Suche in Google Scholar

Weng Y, Chen J, Tu Q, Li Q, Maitz MF, Huang N. Biomimetic modification of metallic cardiovascular biomaterials: from function mimicking to endothelialization in vivo. Interface Focus 2012; 2: 356–365.10.1098/rsfs.2011.0126Suche in Google Scholar PubMed PubMed Central

West EA, Was GS. IGSCC of grain boundary engineered 316L and 690 in supercritical water. J Nucl Mater 2009; 392: 264–271.10.1016/j.jnucmat.2009.03.008Suche in Google Scholar

Willert HG, Broback LG, Buchhorn GH, Jensen PH, Koster G, Lang I, Ochsner P, Schenk R. Crevice corrosion of cemented titanium alloy stems in total hip replacements. Clin Orthop Related Res 1996; 333: 51–75.10.1097/00003086-199612000-00006Suche in Google Scholar

Williams DF. Review–tissue-biomaterial interactions. J Mater Sci 1987; 22: 3421–3445.10.1007/BF01161439Suche in Google Scholar

Williams DL, Ozturk O, Glick S, Wei R, Wilbur PJ. Microstructure of ultrahigh dose nitrogen-implanted iron and stainless steel. Nucl Inst Methods Phys Res B 1991; 59–60: 737–741.10.1016/0168-583X(91)95693-8Suche in Google Scholar

Woodman JL, Jacobs JJ, Galante JO, Urban RM. Metal ion release from titanium-based prosthetic segmental replacements of long bones in baboons: a long-term study. J Orthop Res 1984; 1: 421–430.10.1002/jor.1100010411Suche in Google Scholar PubMed

Xu C, Hu G. Effect of deformation-induced martensite on pit propagation behavior of 304 stainless steel. Anti-Corrosion Methods and Materials 2004; 51: 381–388.10.1108/00035590410560921Suche in Google Scholar

Xu C, Liu X, Qian W, Li Y. Magnetic properties of a stainless steel irradiated with 6 MeV Xe ions. J Nucl Mater 2017; 495: 421–424.10.1016/j.jnucmat.2017.08.028Suche in Google Scholar

Yan Y, Neville A, Dowson D. Bio-tribocorrosion of Co-Cr-Mo orthopaedic implant materials: assessing the formation and effect of the biofilm. Tribol Int 2007; 40: 1492–1499.10.1016/j.triboint.2007.02.019Suche in Google Scholar

Yang J, Black J. Competitive binding to chromium, cobalt and nickel to serum proteins. Biomaterials 1994; 15: 262–268.10.1016/0142-9612(94)90049-3Suche in Google Scholar PubMed

Yu J, Zhao ZJ, Li LX. Corrosion fatigue resistances of surgical implant stainless steels and titanium alloy. Corros Sci 1993; 35: 587–597.10.1016/0010-938X(93)90193-KSuche in Google Scholar

Zamaletdinov I. Pitting on passive metals. Prot Met 2007; 43: 470–475.10.1134/S0033173207050098Suche in Google Scholar

Zaya PGR. In: Evaluation of theories for the initial stages of pitting corrosion, McMaster University, McMaster, 1984.Suche in Google Scholar

Zsklarska-Smialowska Z. Pitting corrosion of metals. Houston, TX: National Association of Corrosion Engineers, 1986.Suche in Google Scholar

Received: 2019-01-22
Accepted: 2019-03-20
Published Online: 2019-04-26
Published in Print: 2019-08-27

© 2019 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 20.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2019-0004/html
Button zum nach oben scrollen