The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review

The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review

Corrosion Science xxx (2014) xxx–xxx Contents lists available at ScienceDirect Corrosion Science journal homepage: www.elsevier.com/locate/corsci R...

1MB Sizes 6 Downloads 57 Views

Corrosion Science xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Corrosion Science journal homepage: www.elsevier.com/locate/corsci

Review

The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review R.K. Gupta a,b,⇑, N. Birbilis c a

Department of Chemical and Biomolecular Engineering, The University of Akron, Akron, OH 44325, USA Institute for Frontier Materials, Deakin University, VIC 3216, Australia c Department of Materials Engineering, Monash University, VIC 3800, Australia b

a r t i c l e

i n f o

Article history: Received 22 January 2014 Accepted 27 November 2014 Available online xxxx Keywords: A. Stainless steel B. Polarisation C. Passivity C. Pitting corrosion

a b s t r a c t Nanocrystalline materials with a grain size <100 nm have attracted significant attention over the past two decades. Various attempts have been made to prepare nanocrystalline stainless steel using various routes, along with the study of attendant corrosion properties. A nanocrystalline structure imparts an improvement in mechanical properties, coupled with distinct corrosion behaviour, not always leading to better corrosion resistance. This paper reviews the relevant works to date which have studied corrosion behaviour of nanocrystalline stainless steels, relating the performance to processing, along with attention given to mechanistic aspects which dictate corrosion of nanocrystalline stainless steel. Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction Stainless steel (SS), owing to its passivation ability and resistance to environmental degradation, has found wide applications ranging from kitchenware to critical components in nuclear reactors [1–6]. Many types of SS have been developed to meet various industrial demands, with the main alloying elements being chromium and nickel. Stainless steels are divided in four main groups based on their microstructure which is composition and processing dependent: ferritic, austenitic, duplex, and martensitic [4–6]. Corrosion behaviour of the various types of SS is well documented, along with the underlying mechanisms as discussed in the related monographs and journal literature [7–13]. Although stainless steels possess a higher density in comparison to the light metals (Al, Mg, Ti), their specific strength can be very high, which is in addition to high stiffness, fracture toughness, and excellent corrosion resistance [14–16]. Therefore, SS with improved strength and corrosion resistance could be next generation of metallic materials (due to high specific strength and durability) for the aerospace and several niche industries. Various strategies, e.g., thermochemical processing, nitriding, carburizing, alloying, etc. have been proved to be effective means of enhancing the surface as well as bulk strength [4,6,17–21]. Grain refinement ⇑ Corresponding author at: Department of Chemical and Biomolecular Engineering, The University of Akron, Akron, OH 44325, USA. Tel.: +1 330 972 7839; fax: +1 330 972 5856. E-mail address: [email protected] (R.K. Gupta).

has great influence on the mechanical strength and corrosion behaviour of SS [6,22–25]. Initial investigations have shown that grain refinement can improve both the corrosion and mechanical properties; however, grain refinement was limited to a few microns [6,21–32]. The finding that grain refinement to <100 nm induces distinctive properties [33–37] stimulated research in developing SS with a nanocrystalline structure. Various nanocrystalline SS have been prepared in recent years via various processing routes and their properties have been investigated. A nanocrystalline structure in SS has been reported to impart significantly higher oxidation resistance (owing to greater Cr diffusion and ease of formation of compact Cr-oxide layer) [38–43] and mechanical properties than [44– 49] conventional coarse grain counterparts of the same chemical composition. The unique properties of nanocrystalline materials are associated with a very fine grain size and a large number of structural defects, i.e., grain boundaries and triple points [33–35,50,51]. Such a high fraction of structural defects in nanocrystalline materials can lead to a significant increase in stored energy which may increase reactivity. This phenomenon is expected to have a dual effect on corrosion behaviour which depends upon material/environment system as reviewed in [52,53]. In passivating electrolytes (i.e., stainless steel in many aqueous environments) a nanocrystalline structure has been reported to lead to an improvement in corrosion resistance, whilst in depassivating electrolytes a decrease in corrosion resistance is reported [52,53].

http://dx.doi.org/10.1016/j.corsci.2014.11.041 0010-938X/Ó 2014 Elsevier Ltd. All rights reserved.

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

2

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

Diffusivity of alloying/impurity elements in nanocrystalline materials are reported to be significantly higher than that in conventional coarse grained materials due to considerably higher volume fraction of grain boundaries [54–58]. Detailed discussion of the diffusion processes in the nanocrystalline material is reported elsewhere [59–62]. The increase in the diffusivities of solute/impurity atoms in nanocrystalline alloys (i.e., P in Ni-P and Co-P alloys [63–67], and Cr in SS [53,68]) was reported to influence corrosion behaviour. However, recently it was proposed that the diffusion coefficient of Cr in SS at room temperature, even in nanocrystalline alloys, was too low to cause any significant influence on corrosion behaviour of SS [48,62]. These contradicting views related to the influence of diffusion on corrosion at room temperature require further investigations. The unique structure of nanocrystalline alloys, which is concomitant with increased free energy, alloying element diffusivity and chemical homogeneity, is expected to induce significantly distinct electrochemical/corrosion properties [62]. Initially, it was believed that nanocrystalline materials (including SS), owing to a high surface energy, may have displayed inferior corrosion resistance [69]. However, in recent years, the corrosion resistance of nanocrystalline SS produced by various routes has been investigated and the reported data presents a range of differing corrosion behaviour. Inconsistency in the reported data may be attributed to be a result of the various processing routes and processing parameters used to fabricate a nanocrystalline structure; in addition to the influence of the various electrolytes in which testing has been executed. Processing route used to fabricate nanocrystalline alloys will not only influence the grain size, but also several other metallurgical parameters, i.e., chemical homogeneity, dislocation density, phase transformations (i.e., resolutionising), and morphology of inclusions (size range and distribution), etc. The influence of all these parameters on corrosion behaviour is as important as grain size. The aim of the present review is to investigate works performed on the corrosion of nanocrystalline SS, as fabricated by various processing routes, leading to a timely discussion of the critical factors influencing the resultant corrosion as presently understood.

2. Passivation and pitting corrosion of stainless steels The corrosion resistance of SS results from the presence of a thin ‘‘passive film’’ upon the metal surface that is typically 1– 3 nm thick [11–13]. The phenomenon of formation of such a protective surface film is called passivation. Various theories of passivation and passive film growth kinetics have been proposed. One of the early models, known as the High Field Model (HFM) assumed that the rate determining step for film growth was the transfer of cations between adjacent lattice sites within the passive film. This model assumed that driving force for the transport of cations in the passive film was an electric field [70]. The Mott-Cabrera model was later proposed, which assumed that the rate-limiting step for the passive film growth is the emission of metal cations from the metal into film at the metal/passive film interface [71,72]. The Mott-Cabrara model [71,72] was later modified by Fehlner and Mort [73] who proposed that the rate-limiting step in passive film growth was the emission of anions from the environment into the passive film at the passive film/environment interface. Macdonald and co-workers proposed a comprehensive model of the passive film growth [74,75] which has been evolving since the late 1970s. This model, known as point defect model (PDM), accounts for the cation and anion mobility, vacancy mobility, interaction among cations, anions and vacancies, and reactions occurring at the passive film/electrolyte and passive film/metal interfaces. The PDM has been extended to account for the break-

down of the passive film and influence of alloying elements on passive film [76–79]. Various theories of passivity and fundamentals of SS corrosion have been covered in key review articles and monographs [10–13,74–77,80–86]. It is now well documented that the addition of Cr to Fe leads to a dramatic improvement in the corrosion resistance, owing to the replacement of the surface (passive) film from the native Fe-oxide to a stable Cr-oxide [56,87,88]. Whilst it is appreciated that a critical Cr content is essential to impart this phenomenon, understanding the enhanced corrosion resistance caused by the Cr derived passivation, and the precise role of Cr, took several decades. Further, it is generally reported that with the an increase in Cr content, the pH region of stable passivity enlarges, passive current density decreases, re-passivation potential shifts to lower (more negative) potentials, and pitting resistance increases significantly [87–90]. Corrosion behaviour of SS is attributed to depend upon the characteristics of the passive film which is largely influenced by Cr content of the alloy. It is generally believed that selective dissolution of Fe and oxidation of Cr leads to the formation of Cr rich passive layer [91–98]. Characteristics of the passive film and its relationship with the alloy composition have been investigated widely [99–104]. It was suggested that a Cr content >50% in passive film was required for stable passivity [105]. The Cr content of the passive film has been demonstrated to increases with increase in Cr content of the alloy [105–107]. An X-ray photo electron spectroscopy study of a series of Fe–Cr alloys indicated that the Cr content in the passive film formed in 0.5 M H2SO4 increased abruptly when Cr content in the alloy was above 13% [106,107]. These reports suggested requirement of the critical Cr content in SS to cause passivity, which was later explained on the basis of percolation model [108–112]. Passive film was proposed to be composed of an outer and inner layer [97,99–104,113]. Cr3+ was found to accumulate in the inner layer of passive film [99,100,106] which was suggested to be pure oxide whereas outer layer was reported to be composed of hydroxides [97,99–104,113]. Chemical composition of the passive film was found to depend upon the environment as well as applied potential [99–103]. It was shown that Fe2+, Fe3+ and Cr3+ were present at lower polarisation potentials, Fe3+ content was reported to increase with increase in potential [97,99,100,102]. The electronic properties of passive films developed over variety of materials are investigated in the past a few decades [114,115]. It is shown that the passive film behaves like a highly doped semiconductor and characteristics of the passive film depend upon alloying additions and Cr content of the SS [115–120]. Increasing Cr content of the SS alters the electronic properties of the passive film in a manner that improves stability of passive film [115,116]. Details of electronics structure of passive film and its dependency on composition can be found in [114–119,121–125]. The passive film on SS is susceptible to breakdown as a result of aggressive ions (i.e. halides), pH changes, and temperature [126,127]. Local breakdown of the passive film can lead to the localized corrosion known as pitting. Pitting can be characterised in three stages [128–130]: (1) Pit initiation or nucleation caused by the breakdown of the passive film. Nucleation events are followed by repassivation. (2) Metastable pitting where pit growth stops on the verge of stability. Metastable pits grow for a finite time and size before repassivation. Metastable pitting is proposed to be a measure of pitting susceptibility [130–136]. (3) Pit growth where pits grow as long as the pit interior can maintain the sufficiently aggressive electrolyte such that repassivation is prevented.

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

All the three stages of pitting are important and have attracted significant research attention. Passive film breakdown is complex and various theories for the passive film breakdown have been proposed. A competitive ion adsorption theory [137] suggests that both O2 and C1 anions can be adsorbed onto metal surfaces. In conditions where C1 adsorption is favoured over O2 adsorption, breakdown of passivity occurs. Hoar and Jacob [138] proposed a complex ion formation theory which suggested that Cl ions are adsorbed around a cation in the passive film surface which forms a complex. This complex is highly soluble and leads to thinning of the passive film locally and further formation of complex with Cl. The process of formation of a high energy complex and its dissolution leads to film breakdown. So-called ion penetration theory suggested that Cl incorporated into and migrates through the passive film and upon reaching the metal/passive film interface results in the film breakdown [137]. Hoar [139] and Sato [140] suggested that the passive film always contain a ‘‘film pressure’’. Passive film breakdown occurs when film pressure exceeds a critical value. Macdonald and co-workers extended the point defect model for passive film growth [74] to explain passive film breakdown [76]. It was proposed that metal vacancies are created at metal/passive film interface as a consequences of diffusion of metal cations from passive film/metal interface to passive film/electrolyte interface where metal vacancies submerge into the metal. Metal vacancies are suggested to pile up at metal/passive film interface when the rate of cation diffusion exceeds the rate of submergence of metal vacancies. The passive film collapses locally when metal vacancy concentration attains a critical value. These mechanisms proposes a plausible explanation of the passive film breakdown. Metastable pitting behaviour of various SS in various electrolytes has been widely investigated [130–136]. Williams et al. [131,132] proposed that pitting rate is proportional to the total number of metastable pitting events. A similar view has been supported by various authors for SS [129,141] as well as for Al and its alloys [142–145]. Since metastable pitting events are large in number and therefore provides a statistically reliable information about pitting corrosion. Punckt et al. [146] performed real-time microscopic in situ visualisation of pitting events on stainless steel in 0.05 M NaCl and suggested that onset of the pitting corrosion is a cooperative critical phenomenon resulting from interactions among metastable pits. Once a stable pit is formed, its propagation is caused because of the development of a highly concentrated solution condition within the growing pit [147]. The composition and microstructure of SS has a strong influence on the pitting corrosion. The Cr content of SS has been demonstrated to play a dominant role in pitting corrosion [90,148]. The pitting potential of Fe–Cr alloys as measured in 0.1 M NaCl increased with Cr content [90]. A dramatic increase in pitting potential was reported when Cr content increased the critical 13% value needed to create a stainless character [90]. Such an increase in pitting potential was attributed to a sudden increase in the Cr content of the passive film. This view is supported by Asami et al. [106] who investigated the Cr content of the passive film developed upon series of Fe–Cr alloys in 0.5 M H2SO4 at 500 mVSCE. Cr content of the passive film, as represented by Cr/ (Fe + Cr), increased by 6 times when Cr content in the alloy was increased from 13% to 15% [106]. These reports suggests that pitting corrosion of stainless steel has a dependence on the chemical composition of passive film which was found to depend upon Cr content of the alloy. It was proposed that in binary Fe–Cr alloys, pit initiation and propagation depends upon formation of Fe clusters [112]. Presence of Fe clusters leads to the local dissolution and depending upon size of cluster, repassivation may or may not occur. When Fe clusters are large enough, pits form and propagate. The size of Fe

3

clusters depends upon Cr content of the alloy. Ryan et al. [149] investigated pitting corrosion of various Fe–Cr binary alloys in HCl and proposed that pitting does not occur when Cr content of binary alloys was >16%. However, Fe cluster formation occurred in alloys having a Cr content below a so-called critical value. The presence of MnS particles and sulphur (S) content of the SS play critical role in pitting corrosion of SS. Most of the pitting events occur around MnS particles, particularly in environments with a low Cl concentration [150–152]. Various probable mechanisms for pitting due to MnS are discussed in the literature. The formation of sulphates, elemental sulphur, and thiosulphate were proposed to be probably reasons for pitting due to MnS particles [153–155]. These models were criticized by Ryan et al. [156] who proposed the formation of a Cr depleted zone around the MnS particles and pit initiation in Cr depleted zone. Focused ion beam/secondary ion mass spectroscopy (FIB/SIMS) analysis provided experimental evidence of such Cr depleted zone around the MnS [156]. A Cr depleted zone was questioned by some researchers [157] and was supported by others [158]. Alloying elements play crucial role in pitting corrosion of stainless steel [121,129,159–161]. Ni is reported to impart a moderate improvement in the pitting resistance [90,162]. Addition of small amounts of Mo [163–167], N [168–170], W [171], V [171], and Cu [119,172] are reported to impart significant improvement in pitting resistance of SS. Combined influence of these alloying elements is provided by pitting resistance equivalent number (PREN) [12,161,173–176]. The most common formulae to calculate PREN is [12]:

PREN ¼ %Cr þ 3:3ð%MoÞ þ 1:65ð%WÞ16ð%NÞ ðall compositions are in weight%Þ A major drawback of using PREN is that it ignores the influence of microstructural features and assumes that all the alloying elements are present in the matrix as solid solution. Carbon present in the SS forms carbides and therefore the amount of alloying elements present in the solid solution decreases with an increase in C content [12,175]. Various mechanisms for influence of these alloying elements on the pitting resistance are proposed in the literature and reviewed [129,148]. However, most of the proposed explanations are probable theories and a precise mechanistic understanding is not yet developed. A detailed discussion on the pitting behaviour of SS could be found in [76,112,126,141,147,177–182]. For the purposes of this paper, it is important to understand that Cr content, alloying elements and impurities, chemical homogeneity, morphology and distribution of secondary phases (i.e., MnS, inclusions), and compactness and homogeneity of the passive film are the main parameters influencing the pitting corrosion of SS in given environmental conditions [126,156,157,183–185]. These parameters are influenced by nanocrystalline structure and processing route. Therefore pitting corrosion of SS is expected to be altered due to the nanocrystalline structure as well as processing route used to fabricate nanocrystalline SS. In summary, classical theories of corrosion/passivation/pitting for conventional coarse grained SS must be considered in the light of novel properties of nanocrystalline materials in order to understand the influence of nanocrystalline structure and processing route on corrosion performance of SS. 3. Influence of nanocrystalline structure and processing route on corrosion of stainless steel Investigating corrosion behaviour of nanocrystalline SS and underlying mechanisms has attracted considerable interest. Nanocrystalline SS used to study corrosion behaviour were prepared

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

4

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

from several processing routes and were tested in various environments, which are expected to have a significant impact on corrosion performance. It should be noted that processing techniques used to produce a nanocrystalline structure not only refine grain size but also impart significant changes in other metallurgical parameters which influence corrosion behaviour. For instance, sputtering and high-energy ball milling are known to cause extended solid solubilities and a homogenous microstructure. Segregation and inclusion formation may not occur in alloys produced by these techniques. Some of the techniques, e.g., equal channel angular pressing (ECAP) and rolling may refine inclusions present in the parent material. The reported corrosion behaviour of SS also reflects the influence of processing methods and parameters employed. This may be one reasons for the reported inconsistency in corrosion behaviour of nanocrystalline SS. For example, most nanocrystalline SS produced by surface mechanical attrition treatment (SMAT) has inferior corrosion resistance that could be due to the dominating influence of defects generated during processing, or cross contamination by metals of differing nobility [186]. In order to develop a full understanding of the nanocrystalline structure, a detailed investigation of the nanocrystalline SS produced by various methods is required. Therefore a review of corrosion of nanocrystalline SS in this section is divided according to the processing routes being employed to fabricate nanocrystalline structure. 3.1. Nanocrystalline stainless steel produced by sputtering Sputtering is a vacuum evaporation process that physically removes portions of a coating material (called the target), and deposits a thin and firmly bonded film onto an adjacent surface, the substrate [187–191]. The process occurs by bombarding the surface of the sputtering target with gaseous ions under high voltage acceleration. As these ions collide with the target, atoms or occasionally entire molecules of the target material are ejected and propelled against the substrate, where they form a very tight bond. The resulting coating is held firmly to the surface by mechanical forces [187–189]. This technique is known to produce amorphous materials, super saturated solid solutions, and nanocrystalline materials. Sputtered materials are reported to be more homogenous and exhibit extended solid solubilities [187–189] that can influence corrosion behaviour. Inturi and Szklarska-Smialowska [192] were first to systematically investigate the effect of nanocrystalline structure on corrosion of SS. Corrosion behaviour of a conventional coarse grained 304 type SS (grain size of 30 lm) was compared with that of its nanocrystalline counterpart (grain size of 25 nm as shown in Fig. 1) in 0.3 wt.% NaCl. Corrosion resistance of the nanocrystalline SS was superior to that of coarse grained material, in spite of the same chemical composition and lower ratio of the Cr to that of Fe in the air born oxide on nanocrystalline SS. Breakdown potential of the nanocrystalline SS was reported to be approximately 850 mV nobler than that of coarse grained SS. The improved pitting resistance of the nanocrystalline SS was attributed to the high microstructural homogeneity and extremely refined grain size [192]. However, the role of MnS and other inclusions was not considered. Schneider et al. [193] reported that influence of nanocrystalline structure on the corrosion performance of a Fe–10Cr alloy in 0.1 M Na2SO4 was largely dependent upon pH. Nanocrystalline Fe– 10Cr alloy showed improved pitting corrosion resistance for pH > 4, which was attributed to the increase in selective dissolution of Fe and therefore Cr enrichment of the passive film [193]. Meng et al. [188] compared the corrosion behaviour of nanocrystalline Fe–10Cr alloy (grain size of 20–30 nm) with that of conventional coarse grained Fe–10Cr alloy (grain size 1000 lm) in 0.05 M H2SO4 + 0.25 M Na2SO4 and 0.05 M H2SO4 + 0.5 M NaCl. Corrosion

Fig. 1. Bright field transmission electron microscope (TEM) image of sputter deposited nanocrystalline stainless steel. A solid ring shaped selected area deflation pattern due to nanocrystalline structure is presented in the inset [192].

current density of nanocrystalline Fe–10Cr alloy was reported to be greater than that of its coarse grained counterpart which was attributed to the higher stored energy (caused by greater defects) in nanocrystalline alloy. Passivation behaviour of nanocrystalline alloys was improved and was attributed to the faster diffusion of Cr to form passive film. Mott–Schottky analysis showed lower donor density in the passive film developed on nanocrystalline alloy. Similarly, Ye et al. [194] reported improvement in pitting resistance of nanocrystalline 309 SS in acidic NaCl due to the formation of a more compact and stable passive film with lower donor density [194]. Inhibition of MnS formation in sputter deposited nanocrystalline coatings has been posited, but not experimentally studied. The influence of nanocrystalline structure on pitting resistance of a new type of austenitic SS was investigated in 3.5 wt.% NaCl where nanocrystalline structure exhibited significantly improved pitting resistance [195,196]. Electrochemical noise analysis in combination with in-situ atomic force microscopy under anodic potential control conditions indicated that mechanism of pit initiation and growth in nanocrystalline SS was significantly different than that of its coarse grained counterpart. Nanocrystalline structure exhibited a fast metastable pit initiation as well as re-passivation [195,196]. In a later study, this behaviour was attributed to the change in chemical structure (presence of elemental Mn and S) of SS surface due to nanocrystalline structure [197]. Presence of elemental Mn and S in nanocrystalline SS lead to higher metastable pitting rate but lowered the probability of transition from metastable to stable pit. On the other hand, in conventional coarse grained materials, these elements combined to form deleterious MnS which although decreased metastable pitting rate (due to decrease in number of electrochemical heterogeneities), probability of metastable pits transforming to stable pits was increased. One criticism of the work related to metastable pitting of nanocrystalline SS is that some critical experimental parameters (data acquisition rate, applied potential with respect to pitting potential, etc.) were not reported clearly in these studies. It becomes important in light of recent work [143,144] related to metastable pitting which shows the importance of applied potential. Greater Cr enrichment of the passive film developed upon nanocrystalline structure was speculated to be reason behind the improved corrosion resistance of sputtered SS. It was indeed confirmed later in year 2012 [68], where X-ray photo electron spectroscopy (XPS) studies revealed significantly greater Cr content of the passive film development upon nanocrystalline SS in 0.05 M H2SO4 + 0.2 M NaCl. Formation of more compact passive film,

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

5

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

3.2. Nanocrystalline stainless steel produced by cold rolling Wang et al. [198] compared the corrosion properties of bulk nanocrystalline 304 SS (produced by severe rolling technique) with that of the conventional coarse grained 304 SS in HCl. Nanocrystalline structure showed significantly improved pitting resistance which was attributed to the formation of a compact passive film due to stronger oxygen adsorption, larger work function, and weaker Cl adsorption [198]. Pitting resistance of nanocrystalline 316L SS produced by cold rolling followed by annealing was found to be superior in 3.5 wt.% NaCl [199]. Improvement in pitting resistance was speculated to be the Cr enrichment of surface layer due to faster diffusion of Cr through grain boundaries. To investigate the mechanism of improved pitting resistance caused by nanocrystalline structure, Pan el al. [200] compared the passive film growth mechanisms in: (1) deep rolled bulk nanocrystalline SS (average grain size 100 nm), (2) magnetron sputtered nanocrystalline SS (average grain size 50 nm), and (3) conventional rolled coarse grained SS (average grain size 100 lm) by electrochemical measurements and in situ AFM observation in 0.05 M H2SO4 + 0.2 M NaCl solution. The growth of the passive film on all three materials was three-dimensional and the growth rates were in decreasing order; nanocrystalline thin film (sputtered) > bulk nanocrystalline (deep rolled) > coarse grained 304 SS. The growth mechanism of passive film formation in coarse grained SS was termed ‘‘progressive’’ whereas it was termed ‘‘instantaneous’’ for nanocrystalline SS. The instantaneous mechanism was attributed to the faster nucleation of the passive film posited to be caused by a nanocrystalline structure. The grain size of the nanocrystalline SS produced by sputtering is much finer, whilst sputtering process is expected to result in a more uniform microstructure and lack of MnS formation. Deep rolling can also alter the morphology of inclusions and texture in addition to grain size. The influence of microstructural features on corrosion behaviour of stainless steel is expected to be important, and the difference in corrosion behaviour is a combined result of grain size and these microstructural changes. 3.3. Nanocrystalline stainless steel produced by high-energy ball milling followed by consolidation High-energy ball milling, also known as mechanical alloying (MA) is a solid-state powder processing technique involving repeated welding, fracture and rewelding of powder-particles in a high-energy ball mill [201,202]. This technique was originally developed to produce oxide-dispersion strengthened (ODS) nickel and iron-base superalloys for applications in aerospace industry [203]. Later, MA has shown to be capable of synthesizing a variety of equilibrium and non-equilibrium phases including nanocrystalline and amorphous materials, and recently it became the most versatile and economical process for synthesis of nanocrystalline materials, due to its simplicity, low cost, and ability to be scaled

up for large production. Synthesis of nanocrystalline metals and alloys, using high-energy ball milling and related phenomena, have recently been reviewed [201,202,204,205]. Gupta et al. prepared nanocrystalline Fe–Cr alloys using high energy ball milling and successfully consolidated them using a novel approach [45,206]. Nanocrystalline and coarse grained alloys thus prepared were used to investigate influence of nanocrystalline structure on corrosion performance in H2SO4 with and without Cl [48,62,207]. Corrosion behaviour of nanocrystalline Fe–Cr alloys (52 nm) was found to be superior to that of their coarse grained (1.5 lm) counterparts. Nanocrystalline structure was reported to decrease passive current density, increase breakdown potential (in presence of chloride ions), and decrease passivation potential and critical current density (Fig. 2). This marked improvement in corrosion behaviour was attributed to the greater Cr enrichment of the passive film which was confirmed by X-ray photo electron spectroscopy and secondary ion mass spectrometry (Fig. 3). Gupta et al. suggested that Cr enrichment of passive film by diffusion of Cr from the bulk to the passive film/metal interface at room temperature (as reported in literature) is less likely. An alternative mechanism for Cr enrichment: accelerated selective dissolution of Fe and oxidation of Cr due to nanocrystalline structure was proposed. Other possible mechanisms (formation of homogenous and compact passive film with improved stability) for enhanced corrosion resistance caused by nanocrystalline structure were also discussed. Most of the investigations comparing corrosion behaviour of nanocrystalline and coarse grained SS were carried out on commercial alloys where presence of inclusions, impurities, segregation, etc. plays crucial role and observed corrosion behaviour was a combined effect of all these parameters. Therefore effect of grain size could not be investigated exclusively. Investigation carried out by Gupta et al. [48,62,207] was a discrete effort to understand the role of nanocrystalline structure where they chose a simple binary system to keep the level of impurities same in coarse grained and nanocrystalline material. Both coarse grained (1.5 lm grain size) and nanocrystalline specimen were prepared by the same processing route which resulted in two identical samples with only difference in the grain sizes. However, the temperature used to sinter coarse grained Fe–Cr alloys (840 °C) was higher than that for nanocrystalline alloys (600 °C), and therefore possibilities of formation additional phases cannot be disregarded. Detailed investigation of microstructure would help in developing further insight into mechanism of corrosion of nanocrystalline SS. 1200 1000 800

Potential (mVSCE)

higher ratio of Cr oxide to Fe oxide in passive film, less incorporation of Cl in passive film and outstanding formation ability of the passive film were additional factors attributed to the increased corrosion resistance of nanocrystalline SS [68]. All the studies of nanocrystalline SS as produced by sputtering reported an improvement in pitting resistance [192,195–197], which was attributed to the refined grain size by most the authors, whilst none of the studies cited herein reported the influence of sputtering on microstructural features, i.e., texture, segregation, inclusions, twinning, dislocations, etc. – which may have an attendant role on corrosion of SS. Future studies dedicated to the chemical and microstructural analysis in conjunction with corrosion behaviour of sputter deposited nanocrystalline SS will help in developing further mechanistic understanding.

600 400 200 0 -200

nc Fe20Cr cg Fe20Cr

-400 -600 10-3

10-2

10-1

i

100

(mA/cm2)

Fig. 2. Anodic polarisation curves of nanocrystalline (nc) and coarse grained (cg) Fe20Cr in 0.05 M H2SO4 solution showing a significant decrease in passive current density, critical current density and passivation potential due to the nanocrystalline structure [48].

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

6

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

Fig. 3. Typical XPS survey scan spectrum of the nanocrystalline (nc) and coarse grained (cg) Fe–20Cr alloys after potentiostatic polarisation at 300 mV/30 min in 0.05 M H2SO4 solution [62].

3.4. Nanocrystalline stainless steel produced by equal channel angular pressing (ECAP) ECAP imparts severe plastic deformation (SPD) which results in significant grain refinement [208,209]. One advantage of this technique is that the cross section of the ECAP extruded billet remains unchanged, and therefore a high level of strain can be imparted without reducing cross section [208,209]. ECAP has been widely used for the preparation of ultrafine grained light metals [210– 212], whilst the use of ECAP in producing nanocrystalline SS is relatively new [213,214]. Nanocrystalline SS (80–120 nm grain size) produced by ECAP showed and improved corrosion behaviour (as represented by lower corrosion current density and nobler corrosion potential in 0.5 M H2SO4) in comparison to its conventional coarse grained counterpart [215]. XPS characterisation performed on air born film showed no obvious difference in Cr content or thickness due to nanocrystalline structure. Improvement in corrosion resistance of nanocrystalline stainless steel was attributed to the improved compactness of the passive film, not to the chemical composition. However, it should be noted that the mechanism of formation of airborne film and therefore its composition is expected to be different than passive film developed in aqueous electrolyte. In addition to grain refinement, ECAP is expected to influence twinning, texture, and perhaps refinement of intermetallics (i.e., MnS). The influence of these parameters on corrosion of SS remains to be investigated. 3.5. Nanocrystalline stainless steel produced by cavitation Nanocrystalline surface layers of 316L with various grain sizes (varying from 91 nm to 167 lm) were prepared by cavitation followed by low-temperature annealing and their corrosion performance was compared in 0.9 wt.% NaCl via potentiodynamic polarisation tests [216]. The nanocrystalline structure was reported lower susceptibility to pitting corrosion and increase repassivation power as indicated by shift in the pitting and protection potentials in more noble directions, and lower corrosion current density. Increased electron work function (EWF) and Cr enrichment of passive film, and more uniform passive film were speculated to be responsible for increased corrosion resistance. 3.6. Nanocrystalline stainless steel produced by surface mechanical attrition treatment (SMAT) Surface mechanical attrition treatment (SMAT) is a relatively new technique where dynamic plastic deformation of surface is

induced by repeated impacts with high-velocity balls. Due to the severe plastic deformation, the coarse grained structure at the surface is refined into the nanostructure without changing the overall chemical compositions [217–219]. The SMAT has been successfully applied in many metallic material systems including SS and various surface properties after SMAT have been investigated [219– 222]. In addition to grain refinement, SMAT has been reported to impart significant surface defects, inclusions from processing media, refinement of inclusions present in parent material, and residual stresses, which all influences corrosion behaviour [12,186,222–225]. Corrosion performance (general corrosion as well as pitting corrosion) of nanocrystalline layer of 304 SS produced by sandblasting was found to be inferior to that of its coarse grained counter parts in 3.5 wt.% NaCl [223,224]. Annealing (350 °C/1 h) of the sandblasted surface, which retained the nanocrystalline structure (grain size of 20 nm), was reported to remove the surface imperfections and therefore an improvement in corrosion behaviour was reported. Scratch tests performed at various applied potentials, showed that passivation abilities of nanocrystalline (annealed) surface were superior to that of coarse grained 304 SS. Electron work function (EWF) of passive layer developed on nanocrystalline surface (annealed) was found to be greater than that on coarse grained SS. Improved corrosion behaviour of nanocrystalline (annealed) SS was attributed to the improved adhesion of passive film and the faster diffusion of Cr. Hao et al. [225] compared corrosion resistance of nanocrystalline 316 stainless steel (produced by SMAT) in 0.1 M NaCl and reported significant decrease in pitting resistance (as demonstrated by pitting potential and critical temperature for pitting), which was attributed to the formation of cracks during SMAT. Annealing was reported to improve corrosion resistance of nanocrystalline surface. Most recently, stress corrosion cracking susceptibility of nanocrystalline SS produced by SMAT [226,227] was found to higher than that of its coarse grained counterpart and this behaviour was attributed to the defects induced by SMAT process. Corrosion behaviour of coarse grained 1Cr18Ni9Ti SS was compared with that of nanocrystalline SS (produced via shot pinning, grain size of 145 nm) in 3.5 wt.% NaCl [228]. Nanocrystalline structure showed improved pitting resistance as shown by significantly higher pitting potential and decrease in number of pits. Toppo et al. [229] used a novel thermo-mechanical surface treatment approach, involving conventional shot blasting followed by laser surface heating to engineer microstructural modification in type 304 austenitic SS. Thermo-mechanical surface treatment resulted in the formation of fine recrystallized grains with some straininduced martensite on the modified surface, and a significant improvement in its resistance against uniform as well as pitting corrosion in deaerated acidified 0.5 M NaCl. Grain refinement and dispersion of alumina inclusions on the modified surface were considered to be the key factors responsible for improvement in corrosion resistance. It was shown that the processing parameters used for SMAT have large influence on the corrosion properties of resulting surface [230,231]. Corrosion resistance was not only influenced by the extent of grain refinement but also by the micro strain and defects induced during SMAT [230]. With increasing SMAT time, simultaneous decrease in grain size and increase in surface defects (e.g., cracks) were reported. Resulting surface properties (including corrosion) can be controlled by choosing right processing parameters, i.e., diameter of balls and SMAT time. Inferior corrosion resistance of nanocrystalline surfaces prepared by SMAT as reported in a few reports could be because of the deleterious influence of surface defects surpassed beneficial effect of grain refinement on corrosion performance. However, a verity of materials in a wide range of SMAT conditions must be investigated to reach any general con-

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

7

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

clusion. Moreover, the influence of SMAT on phase transformation, inclusions from processing media, dislocation density, and micro cracks need to be characterised and their influence on corrosion behaviour remains to be investigated.

low disk shaped. Such pit geometry offers easier repassivation. Lager and semi-elliptic pits covered with finely perforated covers (Lacy covers) were observed in coarse grained SS. Perforated covers act as diffusion barrier and help in maintaining aggressive electrolyte inside the pits, leading to pit stabilisation [9,197,232]. Most of the investigations attributed the nanocrystalline structure alone as the primary reason for distinct corrosion behaviour, which seems a reasonable first order interpretations, and in many cases is supported by the studies reviewed. For example, highenergy ball milled nanocrystalline and coarse grained materials had very similar chemical composition and microstructure except grain size – in essence allowing the investigation of grain size exclusively [48,62]. However, in case of commercial alloys where significant amount of C, S and impurities are present; influence of processing parameters on metallurgical factors other than grain size becomes very important, and need future research attention.

3.7. General discussion Corrosion resistance of nanocrystalline SS produced by various processing routes is summarised in Table 1. Above discussion and Table 1 show that nanocrystalline structure as prepared by various processing routes could lead to a significant improvement in corrosion performance of SS. Tables 2 and 3 summarises reported effects of nanocrystalline structure on the passivation behaviour and pitting resistance respectively. The observed difference as reported in the literature is largely dependent on the processing routes. Some of the studies showed that nanocrystalline SS prepared by SMAT had inferior pitting corrosion resistance, which was attributed to the micro cracks and impurities introduced by processing. It was shown that by choosing appropriate SMAT parameters and post-SMAT treatments (i.e., low temperature annealing), corrosion resistance can be increased. Corrosion current density of nanocrystalline SS in most of the cases was either close to that of coarse grained SS or higher which was attributed to increased surface energy. Critical current and passive current densities were decreased significantly and passivation potential became less noble, indicating improved passivation abilities due to nanocrystalline structure. An example is given in Fig. 2 which shows a clear influence of nanocrystalline structure on passivation of a Fe20Cr alloy in acidic media. Critical current density, passivation potential, and passive current densities decreases significantly. Detailed mechanisms of the influence of nanocrystalline structure on corrosion performance of SS are discussed in the following section. In chloride containing media, pitting potential of nanocrystalline stainless steel was found significantly nobler than that of microcrystalline stainless steel. Pit morphology in nanocrystalline and conventional coarse grained SS was reported to be significantly different (Fig. 4). Pits in nanocrystalline SS were smaller and shal-

4. Mechanisms for the influence of nanocrystalline structure on corrosion of stainless steel Published literature related to the corrosion of nanocrystalline SS as produced by various processing routes has been reviewed in the previous section. Most of the investigations showed that the corrosion performance, in particular passivation abilities and pitting resistance improved significantly due to nanocrystalline structure. Recent studies, based on X-ray photoelectron spectroscopy (XPS), secondary ion mass spectrometry (SIMS), and electrochemical impedance spectroscopy (EIS) have helped in developing mechanistic understanding. In this section, mechanistic aspects of influence of nanocrystalline structure on the corrosion behaviour of SS are discussed. 4.1. Cr enrichment of the passive film Most of the early investigations on corrosion behaviour of nanocrystalline SS speculated that the greater Cr enrichment of the passive film developed upon nanocrystalline SS was responsible for enhanced corrosion performance. However, experimental

Table 1 Summary of reported corrosion behaviour (derived from potentiodynamic polarisation curves) of nanocrystalline stainless steel (produced by various processing routes). In this table corrosion parameters (Ecorr: corrosion potential, Ep: passivation potential, Eb: breakdown potential, icrit: critical current density and ip: passive current density, icorr: corrosion current density) obtained from various papers are presented. Superscript ‘‘nc’’ denotes nanocrystalline and ‘‘cg’’ denotes coarse grained materials. Material

Electrolyte

Preparation technique

Grain size (nm)

cg inc corr/icorr

cg inc crit/icrit

cg inc p /ip

cg Enc p  Ep (mV)

cg Enc corr  Ecorr (mV)

cg Enc b  Eb (mV)

Reference

SS304 Fe10Cr

0.3 wt.% NaCl 0.05 M H2SO4 + 0.25 M Na2SO4 0.1 N H2SO4 (pH < 1) 0.1 N Na2SO4 (pH > 4) 0.05 M H2SO4 + 0.25 M Na2SO4 0.05 M H2SO4 + 0.5 M NaCl 3.5 wt.% NaCl 3.5 wt.% NaCl 0.05 M H2SO4 + 0.5 M NaCl 3.5 wt.% NaCl 3.5 wt.% NaCl 3.5 wt.% NaCl 1 M H2SO4

DC sputtering Magnetron sputtering

25 20–30

– 2.4

– 0.16

– 0.8

– 89

– 61

850 0

[192] [188]

Magnetron sputtering Magnetron sputtering Magnetron sputtering

40 40 <50

Similar Similar –

cg – inc p > ip Similar cg nc cg 2.5 inc p < ip Ep < Ep cg cg inc Enc crit < icrit – p < Ep

Similar – –

– – –

[193] [193] [194]

Magnetron sputtering

<50

Similar







700

[194]

Magnetron sputtering Magnetron sputtering Magnetron sputtering

20 50 50

– – –

– –

Similar – Similar – 100 –

100 50 –

900 700 650

[195,196] [197] [68]

Sandblasting Sandblasting + annealing High energy shot pitting Shot blasting + laser treatment Multi-pass cold rolling Deep rolling

20 20 18

– – –

– – –

cg inc p > ip – cg inc p < ip – cg inc < i p p –

150 150 –

150 0 600

[223,224] [223,224] [228] [229]

40 100

0.075



cg inc p < ip – cg inc p < ip –

36 mV cg Enc corr > Ecorr

500 400

[199] [200]

– –

0.25 0.17

0.55 0.37

0 0

0 119

[48] [48]

Fe10Cr Fe10Cr 309SS 309SS Austenitic SS 304SS 304SS 304SS 304SS 1Cr18Ni9Ti 304SS 316L SS 304SS Fe20Cr Fe20Cr

3.5% NaCl 0.05 M H2SO4 + 0.2 M NaCl 0.05 M H2SO4 0.05 M H2SO4 + 0.5 M NaCl

High-energy ball milling 52 High-energy ball milling 52



180 190

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

8

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

Table 2 Summary of reported effect of nanocrystalline structure on the passivation behaviour of stainless steel. Material

Electrolyte

Passivation ability

Comments

Reference

Fe10Cr/magnetron sputtering

0.05 M H2SO4 + 0.25 M Na2SO4 0.1 N H2SO4 (pH < 1) 0.1 N Na2SO4 (pH > 4) 0.05 M H2SO4 + 0.5 M NaCl 3.5 wt.% NaCl

Improved

Lower critical current density and passivation potential. Higher repassivation potential. Lower done density and increased Cr content caused by diffusion

[188]

Decreased

Due to higher microcracks, inclusions from processing media, and residual stresses

[193]

Improved

Due to faster selective dissolution of Fe and therefore Cr enrichment

[193]

Improved

Homogenous, compact and stable passive film

[194]

Improved

[223,224]

Fe10Cr/magnetron sputtering Fe10Cr/magnetron sputtering 309SS/magnetron sputtering 304SS/sand blasting and annealing 1Cr18Ni9Ti/high energy shot pitting 316L SS/multi-pass cold rolling 304SS/deep rolling

Fe20Cr/high-energy ball milling Fe20Cr/high-energy ball milling

3.5 wt.% NaCl

Improved

Improved diffusion of Cr, higher EWF of passive film, Purportedly improved bonding between metal and interface No discussion

3.5% NaCl

Improved

Faster diffusion of Cr

[199]

0.05 M H2SO4 + 0.2 M NaCl 0.05 M H2SO4

Improved

Instantaneous nucleation and faster three dimensional growth of passive film was suggested

[200]

Improved

[48]

0.05 M H2SO4 + 0.5 M NaCl

Improved

Cr enrichment of passive film through a purportedly more rapid selective dissolution of Fe and oxidation of Cr, formation of homogenous and compact passive film, improved oxygen adsorption Cr enrichment of passive film through a purportedly more rapid selective dissolution of Fe and oxidation of Cr, formation of homogenous and compact passive film, improved oxygen adsorption

[228]

[48]

Table 3 Summary of the reported effects of nanocrystalline structure on the pitting corrosion of stainless steel. Material

Electrolyte

Pitting Comments resistance

Reference

SS304/DC sputtering Fe10Cr/magnetron sputtering Austenitic SS/ magnetron sputtering 304SS/magnetron sputtering 304SS/magnetron sputtering 304SS/sand blasting and annealing 316L SS/multi-pass cold rolling 304SS/deep rolling Fe20Cr/highenergy ball milling

0.3 wt.% NaCl

Improved Attributed to nanocrystalline grain size and homogeneity of sputtered surface

[192]

0.05 M H2SO4 + 0.25 M Na2SO4 3.5 wt.% NaCl

Improved Attributed to lower done density and increased Cr content caused by faster diffusion

[188]

Improved Increased passivation rate of metastable pits. Improved formation and continuous growth of passive film

[195,196]

3.5 wt.% NaCl

Improved Suppression of combination of Mn and S and therefore decrease in deleterious MnS

[197]

0.05 M H2SO4 + 0.5 M NaCl

[68]

3.5 wt.% NaCl

Improved Decreased diffusivity of defects in passive films, decreased adsorption of Cl on the surface, higher Cr content in the passive film caused by diffusion of Cr Improved Improved diffusion, higher EWF of passive film, better bonding between metal and interface

3.5% NaCl

Improved Faster diffusion of Cr

[199]

0.05 M H2SO4 + 0.2 M NaCl 0.05 M H2SO4 + 0.5 M NaCl

Improved Instantaneous nucleation and faster three dimensional growth of passive film Improved Cr enrichment of passive film through faster selective dissolution of Fe and oxidation of Cr, formation of homogenous and compact passive film, improved oxygen adsorption

[200] [48]

validation of greater Cr enrichment the passive film was presented most recently [48,62,68,207]. Gupta et al. compared the Cr content of the passive film developed upon nanocrystalline and coarse grained FeCr alloys in acidic medium (with and without chloride ions) using SIMS and XPS and reported that the passive film in nanocrystalline alloys had significantly higher Cr content [48,62,207]. Sputtering time required to reach to the base composition was similar in the two alloys, indicating similar thickness of passive film developed upon nanocrystalline and coarse grained FeCr alloys. XPS scans as obtained after passivation at potential in the middle of passive region (Fig. 3), shows higher Cr content of the passive film developed upon nanocrystalline alloy. Similarly, greater Cr content in the passive film developed upon nanocrystalline 304 SS was reported by Pan et al. [68]. It was reported that in spite of higher Cr content, the oxidation states of Cr and Fe present

[223,224]

in the passive films were similar [62]. Passive film was composed of both oxides and hydroxides of iron and chromium. Additional peaks relating to metallic Fe and Cr were observed, which were attributed to the presence of thin passive film and therefore signals from the base metal contributed in the analysis of passive film [62]. Greater Cr enrichment caused by a nanocrystalline structure has been speculated due to higher fraction of grain boundaries and other structural defects, which are known to provide faster diffusion paths for Cr [68]. However, it was suggested that diffusion coefficient of Cr at room temperature, even in nanocrystalline alloys was too low to cause any significant influence on Cr content of passive layer through diffusion of Cr from the bulk to the passive film [48,62]. This view was also supported by two previous studies where Cr content of the air born oxide films on nanocrystalline and coarse grained SS was found similar. Gupta et al. [48,62] suggested

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

9

4.3. Microstructural homogeneity

Fig. 4. SEM micrographs showing typical pit morphology in (a) coarse grained (b) nanocrystalline stainless steel [197].

Electrochemical heterogeneities, i.e., segregation, inclusions, precipitates, and intermetallics are present in the conventional SS, which are deleterious to overall corrosion performance and are major cause of pitting corrosion. Processing techniques used to produce nanocrystalline SS are known to develop a more homogenous microstructure and many of the investigations, particularly sputtered deposited and high-energy ball milled alloys, suggested that removal of segregation and formation of more uniform microstructure to be the main reason of improved corrosion performance of nanocrystalline SS [192,194]. According to the percolation theory, passivation is achieved only when a certain number of Cr atoms surround Fe atoms [108–112]. In conventional stainless steel, segregation of Fe atoms decreases the percolation tendency which increases pitting susceptibility. However, in nanocrystalline alloys, atoms are more uniformly distributed and tendency of segregation is less which leads to more percolation and more uniform passive layer causing improved passivation abilities and higher pitting resistance [108–112]. The size and chemical composition of the inclusions/precipitates in SS has a significant effect on the pitting resistance. Inclusions of size below a critical limit are known not to initiate pit irrespective to the chemical composition. The critical size was reported to largely depend upon the chemical composition of the inclusion. It was reported that MnS particles <300 nm as present in type 304 SS do not cause pitting corrosion when exposed to 0.6 M NaCl [233]. Nanocrystallization is reported to remove and/ or refine the MnS particles present in the commercial stainless steel which is expected to improve over all corrosion performance [194]. The morphology of MnS seems highly dependent on the processing route, whilst the sputtering process was posited to inhibit formation of MnS. High-energy ball milling of commercial SS was not performed and hence whether it is a technique which removes MnS is unknown. Pan et al. [197] reported the existence of elemental Mn and S in nanocrystalline SS as prepared by DC magnetron sputtering, whereas Mn and S combined to form MnS in conventional SS. Elemental Mn and S were reported to be less deleterious in comparison to MnS particles and therefore corrosion performance of nanocrystalline SS was found to be improved. 4.4. Increased surface activity

that greater Cr enrichment of passive film caused by nanocrystalline structure seems possible to explain based on selective dissolution of Fe and oxidation of Cr during corrosion process. Nanocrystalline structure increases reactivity, which accelerates the selective dissolution of Fe and oxidation of Cr to Cr-oxide or hydroxide [48,62]. This process leads to the development of a passive film with higher Cr content. 4.2. Electronic structure of passive film The point defect model (PDM) as proposed by Macdonald and coworkers [74–76] suggests that density and diffusivity of point defects (e.g., oxygen vacancies and metal cation vacancies) play important role in passivation and pitting. The growth and breakdown of the passive film involves generation and annihilation of these point defects which is influenced by the substrate material. Thus, the key parameters in determining the passivity and pitting are density and the diffusivity of the defects in the passive film. These parameters can be determined by Mott–Schottky analysis. Most of the early investigations reported that donor density of the passive film formed on nanocrystalline SS was significantly less than that of coarse grained SS which indicated improved passivation behaviour and pitting resistance of nanocrystalline alloys [188].

Increased surface activity of nanocrystalline SS leads to the acceleration of reactions causing the passivation, i.e., dissolution of Fe and oxidation of Cr. Therefore, passivation potential of nanocrystalline SS was reported to be less noble than that of conventional coarse grained SS of same chemical composition [48,62,207,234]. Some of the studies showed that electron work function (EWF) is a measure of activity and reported that passive film developed upon nanocrystalline alloys had higher EFW, indicating them to be less reactive and more protective [223]. Some of the studies showed than nanocrystalline structure lowers EFW of the alloy, leading to the higher reaction rate and also easier formation of passive film [69,198]. Attempts were made to link EFW of alloy with the adhesive strength of the passive film and reported that higher reactivity of nanocrystalline surfaces leads to enhanced adhesion strength between passive film and the alloy surface due to the increase in the electron activity at grain boundaries and possible pegging of the passive film into the grain boundaries [69,235]. 4.5. Formation of compact passive film Nucleation and growth is the main mechanism of passive film formation. The nucleation of oxide is favoured at the high energy sites, i.e., surface defects in the form of dislocations, grain bound-

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

10

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

aries, triple points, impurities, etc. Since nanocrystalline materials are composed of the large fraction of surface defects therefore they offer a large fraction of closely spaced nucleation sites. The presence of closely spaced nucleation sites reduces the lateral distance necessary for the lateral growth of a uniform oxide layer to cover the entire surface. Change in the nucleation and growth mechanism due to nanocrystalline structure and therefore enhanced passivation abilities of 304 SS is reported in literature [200]. Compactness of the passive film was investigated by monitoring the passive current as a function of time as it was shown that current versus time plot usually follows the following relationship [68,194,236]

I ¼ 10ðAþk log tÞ

ð1Þ

where k represents the slope of double-log plot for the potentiostatic polarisation. k = 1 indicates the formation of compact, highly protective passive film, whilst k = 0.5 indicates the presence of porous film growing as a result of a dissolution and precipitation processes. Based on above analysis, many of the investigations proposed formation of compact passive film on nanocrystalline alloys where k was close to 1 for nanocrystalline SS [68,194]. More recently AFM studies suggested formation of more uniform passive film over nanocrystalline surface as large fraction of grain boundaries leads to instantaneous formation of passive film [200]. 4.6. Decreased probability of metastable to stable pit formation Pitting corrosion of nanocrystalline stainless steel has been investigated widely and reviewed recently [53]. Morphology of pits in nanocrystalline and microcrystalline materials was very different (Fig. 4). Current transients, a representative of metastable pitting events, obtained under potentiostatic control showed a marked influence of nanocrystalline structure [53]. Metastable pitting rate and its repassivation probability in nanocrystalline material was significantly higher than that in microcrystalline alloy. It was suggested that increased metastable pitting rate was due to: (1) elemental distribution of Mn and S, and (2) higher activity of nanocrystalline surface [53]. Nanocrystalline structure lead to faster repassivation rate and therefore probability of metastable to stable pit transition in nanocrystalline SS were lower. 4.7. Incorporation of elements from processing media Processes used to produce nanocrystalline alloys are known to form extended solid solubilities and formation of metastable phases. Possibilities of incorporation of elements from processing media (e.g., high-energy ball milling media, atmosphere used in high-energy ball milling or sputtering) cannot be overruled. Corrosion of SS is reported to be sensitive to small additions of elements, e.g., N, S, C and therefore incorporation of elements from processing media, even in very small amounts, is expected to have significant impact on corrosion properties. For instance, sputter deposited pure Al thin films exhibited high pitting potential which was initially attributed to the grain refinement and improved electrochemical homogeneity [237–240]. However such large influence on pitting corrosion due to these parameters in pure metals like Al seems less likely. More recently, it has been shown that a significant amount of O was incorporated in sputter deposited Al [241]. The incorporation of O to the sputtered Al film is attributed to the enhanced pitting resistance [242]. These studies could be very important in explaining improved corrosion resistance of nanocrystalline SS. N, O, or other elements from atmosphere could be introduced into SS during synthesis of nanocrystalline surface and could influence corrosion properties. Possibilities of incorporation of these elements depend upon processing method. Dedicated

studies investigating the possibilities of incorporation of O, N, Ar, etc. from the processing atmosphere could be very useful in explaining change in corrosion/mechanical properties of these alloys. 4.8. Oxygen adsorption According to adsorption theory [85,243,244], the passivity is regarded as adsorption of atoms which retarded the metal dissolution. Increased adsorption of oxygen on nanocrystalline alloys is reported [34,35,198,245,246] and therefore it may be one of the reasons of increased passivation ability of nanocrystalline SS. Wang et al. [198] showed an increase in oxygen adsorption on nanocrystalline SS surface and attributed it to be one of the reasons leading enhanced corrosion resistance of nanocrystalline SS. 5. Prospects and gaps Corrosion behaviour of SS as prepared by a variety of processing routes has been reported. Most of the investigations regarding the corrosion behaviour of nanocrystalline SS have associated grain refinement with improved pitting corrosion resistance. The role of processing route and the associated processing parameters that impart significant microstructural changes are not well investigated. Processing routes used for preparation of a nanocrystalline structure are expected to result in homogenous microstructures without any clustering, which is expected to result in improved resistance against pitting corrosion. Synthesis techniques, e.g. sputtering, high-energy ball millings are known to cause extended solid solubilities and therefore incorporation of elements from processing media/atmosphere is plausible. Microstructure and electrochemical characteristics of nanocrystalline SS must be investigated at atomic level to understand the corrosion mechanisms. Alloying SS with low levels of Mo, Mn, S, W, N, Ta, C, Cu, etc. plays an important role in microstructure and corrosion. The review of the role of such elements is beyond the scope of this study, whilst such alloying effects are discussed briefly in section two of this paper and covered in monographs for conventional SS [11,12]. The influence of these alloying elements on the corrosion behaviour of nanocrystalline stainless steel has not been investigated and needs future research attention. The proposed mechanisms for the Cr enrichment of the passive films as presented in the associated works remain somewhat rudimentary. The faster diffusion of Cr leading to higher Cr enrichment of passive films is yet to be validated by specific experiments. The investigations of the diffusivity of Cr from bulk to metal/electrolyte interfaces would help in developing a better understanding. Compositional characterisation of passive film developed over nanocrystalline SS in various environmental conditions has attracted only limited attention. Detailed characterisation of the passive film (i.e., oxidation states and quantification of elements as a function of thickness of the passive film) should help in understanding the mechanistic aspects of passivity in nanocrystalline SS. The influence of nanocrystalline structure on the various stages of pitting corrosion (i.e., pit initiation, metastable pitting, and pit growth) has not been investigated in detail, and therefore the influence of a nanocrystalline structure on these stages is not known. Perhaps most critically, if nanocrystalline SS are ever considered for in-service use, the repassivation potential as obtained via cyclic potentiodynamic polarisation, is a key parameter that is not reported in literature in the context of nanocrystalline SS. Detailed future studies investigation various stages of pitting corrosion and repassivation mechanism will further help in understanding role of nanocrystalline structure in passivation and pitting of SS.

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

One of the biggest challenges limiting the application of nanocrystalline SS is their poor ductility [247–249] and thermal stability [250–252], and therefore lack of processing techniques capable of producing large samples [206]. Some recent studies [45,253– 255], however, have shown that inferior ductility as reported in nanocrystalline materials it is not the inherent property of nanocrystalline structure and could be a manifestation of defects induced during the processing. Consequently, dedicated work is required to explore the possibilities of producing nanocrystalline SS with reasonable ductility whilst retaining a high yield strength. Compaction of nanocrystalline SS powders as produced by highenergy ball milling (powder metallurgy route) seems a viable method to produce large samples; however, high strength and poor thermal stability of nanocrystalline materials are the biggest hurdles. Issues of high temperature stability were investigated and various methods to hinder grain growth were proposed recently [256–259]. Use of small amount of grain growth stabilizer (i.e., Zr addition) can pin down the grain growth and these materials can be processed at higher temperature leading to improved properties [259–261]. Recently various methods, e.g. spark plasma sintering, annealing prior to compaction, in situ consolidation, etc. have been proposed for consolidation of ball milled nanocrystalline powder. Dedicated work, combining modern processing techniques and understanding of properties of nanocrystalline SS should help in developing techniques for production of large samples. Stress corrosion cracking (SCC) susceptibility of fine microstructure generated by various machining [262–264] or intentionally [227] developed nanocrystalline structures was analysed and it was found that a nanocrystalline structure was more susceptible to SCC. However, SCC susceptibility seemed to be induced by the micro cracks generated during processing. Nanocrystalline SS possess a large fraction of defects and therefore diffusivity of hydrogen can be increased, which is expected to influence hydrogen embrittlement susceptibility of SS. Similarly, intergranular failure (be it corrosion or embrittlement due to potentially enhanced H-levels) of SS may be influenced by a nanocrystalline structure. However, the influence of nanocrystalline structure on the crevice corrosion, intergranular corrosion (IGC), and hydrogen embrittlement of SS has not yet been investigated. Any investigation regarding the effect of a nanocrystalline structure (produced by various processing routes) on the SCC, IGC, HE, and crevice corrosion of SS will be of great interest.

6. Potential applications of nanocrystalline stainless steel The nanocrystalline SS can be used to prepare corrosion/wear resistant coatings – the latter reliant on superior mechanical properties from nanocrystalline structure. Many of the reported techniques, i.e., sputtering, pulse electrochemical deposition, and SMAT can be readily used for this purpose. Rolling can be used to produce bulk nanocrystalline or ultrafine grained SS with improved mechanical properties, and in the appropriate environment, improved corrosion resistance. Methods for grain refinement such as ECAP are more difficult to upscale, however for particular applications may be suited; whilst extreme deformation imparted via high pressure torsion (HPT) only remains relevant to scientific studies. Due to the improved diffusivity of elements in nanocrystalline materials, nitriding, carburizing, etc. may be performed at lower temperatures or different operating conditions, and could be used to prepare functionally graded materials. The strength of nanocrystalline SS has been reported to be significantly higher than that of conventional coarse grain SS [44–47]. For instance, the yield strength of nanocrystalline 316L SS as produced via SMAT was reported to be 1450 MPa, which is signifi-

11

cantly higher than that of conventional 316L SS [44]. Similarly, the yield strength of in situ consolidated nanocrystalline Fe– 20Cr–10Ni (grain size <10 nm) was reported to be 2.6 GPa [45]. Such increased strength leads to very high specific strength, which is in addition to corrosion resistance and high stiffness, may provide nanocrystalline SS unique property profiles. Nanocrystalline SS produced by high-energy ball milling has the potential to replace sintered stainless steel components (prepared by compaction and sintering of SS powder) which are used in the automotive industry [265]. These components usually require excellent high temperature corrosion resistance (as engine parts are exposed to high temperatures), wear resistance, and high strength. Published data related to high temperature oxidation of Fe–Cr alloys shows that nanocrystalline Fe–Cr alloys offer significantly improved oxidation resistance at lower Cr content in elevated temperature applications [38–43]. Use of nanocrystalline SS powders are expected to improve the quality of conventional sintered components. A criticism of high-energy ball milling for production of nanocrystalline SS is the associated high cost, which involves production of metal powder prior to high-energy ball milling. However, recent studies have shown that starting material for high-energy ball milling need not to be in powder form. The starting material could be industrial by-products such as swarf/ chips formed during machining [266]. High-energy ball milling of such industrial by-products is attractive; moreover, nanocrystalline powders as produced by high-energy ball milling may also be a starting material for net shape manufacturing of components. Cost associated with the production of nanocrystalline SS could be compensated with the benefits of net shape manufacturing (i.e., reduced material and machining cost) and improved properties of nanocrystalline SS. Nanocrystalline SS exhibited significantly improved corrosion resistance at lower Cr content and therefore this may contribute to alloys for use in clean energy generating systems where harsh conditions (i.e., high temperatures in solid oxide fuel cells, corrosive biomass, etc.) exist. Dedicated work is however required to explore the possibilities of economical production of nanocrystalline materials and applications of nanocrystalline SS, in the face of conventional corrosion and heat resistant alloys that can presently be commercially acquired (i.e. Ni alloys).

7. Conclusions This review has covered the body of literature that pertains to the corrosion of nanocrystalline stainless steel. Based on the works reviewed, the general consensus is that corrosion resistance of stainless steel improves due to the nanocrystalline structure for the majority of test electrolytes. It was also determined from a wider analysis, that corrosion resistance was largely dependent upon the processing route used to generate or impart the nanocrystalline structure. Of the studies surveyed, the passive film developed over nanocrystalline stainless steel in various electrolytes was purported to be more stable, more compact and contain a lower defect density, whilst also possessing a higher Cr content in the passive film (for an equivalent bulk Cr content). Nanocrystalline stainless steels were noted for having a more uniform microstructure with implications in improvement of resistance against pitting corrosion. The associated mechanisms related to the corrosion and instances of improvement in corrosion resistance have been discussed herein. In addition to nanocrystalline structure, processing route had strong influence on corrosion behaviour of stainless steel. Dedicated future investigation on nanocrystalline stainless steel prepared by various processing routes are required in order to develop further understanding of the influence of nanocrystalline

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

12

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

structure and processing routes on corrosion behaviour of stainless steel. References [1] N. Baddoo, 100 years of stainless steel: a review of structural applications and the development of design rules, Struct. Eng. 91 (2013) 10–18. [2] N.R. Baddoo, Stainless steel in construction: a review of research, applications, challenges and opportunities, J. Constr. Steel Res. 64 (2008) 1199–1206. [3] L. Gardner, The use of stainless steel in structures, Prog. Struct. Eng. Mater. 7 (2005) 45–55. [4] R.M. Davison, T.R. Laurin, J.D. Redmond, H. Watanabe, M. Semchyshen, A review of worldwide developments in stainless steels, Mater. Des. 7 (1986) 111–119. [5] J. Beddoes, J.G. Parr, Introduction to Stainless Steels ASM International, ASM International, Materials Park, OH, 1999. [6] K.H. Lo, C.H. Shek, J.K.L. Lai, Recent developments in stainless steels, Mater. Sci. Eng. R: Rep. 65 (2009) 39–104. [7] A. Iversen, B. Leffler, Aqueous corrosion of stainless steels, Shreir’s Corros. (2010) 1802–1878. [8] R.W. Revie, Uhlig’s Corrosion Handbook, third ed., Wiley, Hoboken, 2011. http://dx.doi.org/10.1002/9780470872864. [9] R.C. Newman, 2001 W.R. Whitney award lecture: understanding the corrosion of stainless steel, Corrosion 57 (2001) 1030–1041. [10] D.D. Macdonald, Passive films: nature’s exquisitely nano-engineered protection system, Curr. Appl. Phys. 4 (2004) 129–132. [11] H. Kaesche, Corrosion of Metals: Physicochemical Principles and Current Problems Erlangen, Springer, 2003. [12] A.J. Sedriks, Corrosion of Stainless Steels, second ed., John Wiley & Sons, New York, 1996. [13] P. Schmuki, From Bacon to barriers: a review on the passivity of metals and alloys, J. Solid State Electrochem. 6 (2002) 145–164. [14] S. Narahari Prasad, M. Narayana Rao, Stainless steel – a versatile engineering material for critical applications, Adv. Mater. Res. 794 (2013) 44–49. [15] A. Saha Podder, A. Bhanja, Applications of stainless steel in automobile industry, Adv. Mater. Res. 794 (2013) 731–740. [16] K. Lu, The future of metals, Science 328 (2010) 319–320. [17] H. Dong, S-phase surface engineering of Fe–Cr, Co–Cr and Ni–Cr alloys, Int. Mater. Rev. 55 (2010) 65–98. [18] M. Laleh, F. Kargar, M. Velashjerdi, Low-temperature nitriding of nanocrystalline stainless steel and its effect on improving wear and corrosion resistance, J. Mater. Eng. Perform. (2012) 1–7. [19] X. Ma, L. Wang, S.V. Subramanian, C. Liu, Studies on Nb microalloying of 13Cr super martensitic stainless steel, Metall. Mater. Trans. A: Phys. Metall. Mater. Sci. 43 (2012) 4475–4486. [20] D. Ye, J. Li, W. Jiang, J. Su, K. Zhao, Effect of Cu addition on microstructure and mechanical properties of 15%Cr super martensitic stainless steel, Mater. Des. 41 (2012) 16–22. [21] J.A. Scott, J.E. Spruiell, Grain refinement of wrought austenitic stainless steels by rapid heating, Metall. Trans. 5 (1974) 255–259. [22] B. Ravi Kumar, S. Sharma, B. Mahato, Formation of ultrafine grained microstructure in the austenitic stainless steel and its impact on tensile properties, Mater. Sci. Eng. A 528 (2011) 2209–2216. [23] H.b. Li, Z.h. Jiang, Z.r. Zhang, Y. Yang, Effect of grain size on mechanical properties of nickel-free high nitrogen austenitic stainless steel, J. Iron Steel Res. Int. 16 (2009) 58–61. [24] A. Di Schino, M. Barteri, J.M. Kenny, Effects of grain size on the properties of a low nickel austenitic stainless steel, J. Mater. Sci. 38 (2003) 4725–4733. [25] P. Rodriguez, Influence of metallurgical variables on corrosion, Key Eng. Mater. 36 (1989) 31–42. [26] T. Tsuru, R.M. Latanision, Corrosion resistance of microcrystalline stainless steels, J. Electrochem. Soc. 129 (1982) 1402–1408. [27] W.B. Nowak, Microcrystalline/amorphous iron alloy films for corrosionresistant coatings, Mater. Sci. Eng. 23 (1976) 301–305. [28] D. Liu, F. Wang, C. Cao, H. Lou, L. Zhang, H. Lin, Pitting corrosion resistance of microcrystalline coatings of sputtered 321 stainless steel, Corrosion 46 (1990) 975–977. [29] R. Sandstrom, H. Bergqvist, Strengthening of nitrogen alloyed austenitic stainless steel, vol. 1, 1976, pp. 281–285. [30] M. Hasegawa, M. Osawa, Corrosion behaviour of ultrafine grained austenitic stainless steel, Corrosion 40 (1984) 371–374. [31] N.M. Pakhomova, I.A. Levin, Effect of grain growth on the intercrystalline corrosion of 18–8 stainless steel, Prot. Met. 9 (1973) 598–600. [32] V.K. Gouda, R.W. Staehle, Effect of Metallurgical variables on the corrosion fatigue behaviour of 304 stainless steel in sulphuric/chloride media, Br. Corros. J. 15 (1980) 111–117. [33] H. Gleiter, Nanostructured materials: basic concepts and microstructure, Acta Mater. 48 (2000) 1–29. [34] H. Gleiter, Nanocrystalline materials, Prog. Mater. Sci. 33 (1989) 223–315. [35] H. Gleiter, Our thoughts are ours, their ends none of our own: are there ways to synthesize materials beyond the limitations of today?, Acta Mater 56 (2008) 5875–5893. [36] G. Liu, G.J. Zhang, F. Jiang, X.D. Ding, Y.J. Sun, J. Sun, E. Ma, Nanostructured high-strength molybdenum alloys with unprecedented tensile ductility, Nat. Mater. 12 (2013) 344–350.

[37] T.G. Langdon, Twenty-five years of ultrafine-grained materials: achieving exceptional properties through grain refinement, Acta Mater. 61 (2013) 7035–7059. [38] R.K. Gupta, N. Birbilis, J. Zhang, Oxidation resistance of nanocrystalline alloys, in: H. Shih (Ed.), Corrosion Resistance, Intech Publisher, 2012, pp. 213–238. [39] R.K. Gupta, R.K. Singh Raman, C.C. Koch, Fabrication and oxidation resistance of nanocrystalline Fe10Cr alloy, J. Mater. Sci. 45 (2010) 4884–4888. [40] R.K.S. Raman, R.K. Gupta, Oxidation resistance of nanocrystalline vis-à-vis microcrystalline Fe–Cr alloys, Corros. Sci. 51 (2009) 316–321. [41] R.K.S. Raman, R.K. Gupta, C.C. Koch, Resistance of nanocrystalline vis-a-vis microcrystalline Fe–Cr alloys to environmental degradation and challenges to their synthesis, Philos. Mag. 90 (2010) 3233–3260. [42] S.G. Wang, M. Sun, H.B. Han, K. Long, Z.D. Zhang, The high-temperature oxidation of bulk nanocrystalline 304 stainless steel in air, Corros. Sci. 72 (2013) 64–72. [43] M. Govindaraju, N.S. Cheruvu, K. Natesan, Development and evaluation of nanostructured coatings for long-term corrosion and oxidation performance, Am. Soc. Mech. Eng., Pressure Vessels Piping Div. (Publ.) PVP 9 (2010) 231– 237. [44] X.H. Chen, J. Lu, L. Lu, K. Lu, Tensile properties of a nanocrystalline 316L austenitic stainless steel, Scr. Mater. 52 (2005) 1039–1044. [45] R.K. Gupta, K.S. Darling, R.K. Singh Raman, K.R. Ravi, C.C. Koch, B.S. Murty, R.O. Scattergood, Synthesis, characterization and mechanical behaviour of an in situ consolidated nanocrystalline FeCrNi alloy, J. Mater. Sci. 47 (2012) 1562– 1566. [46] T. Roland, D. Retraint, K. Lu, J. Lu, Enhanced mechanical behavior of a nanocrystallised stainless steel and its thermal stability, Mater. Sci. Eng. A 445–446 (2007) 281–288. [47] H.W. Ni, H. He, G.Q. Li, J. Liu, Preparation of nanocrystalline 430L stainless steel by HEBM and SPS, J. Iron Steel Res. Int. 15 (2008) 73–76. [48] R.K. Gupta, R.K.S. Raman, C.C. Koch, B.S. Murty, Effect of nanocrystalline structure on the corrosion of a Fe20Cr alloy, Int. J. Electrochem. Sci. 8 (2013) 6791–6806. [49] M. Tikhonova, Y. Kuzminova, A. Belyakov, R. Kaibyshev, Nanocrystalline S304H austenitic stainless steel processed by multiple forging, Rev. Adv. Mater. Sci. 31 (2012) 68–73. [50] G. Palumbo, S.J. Thorpe, K.T. Aust, On the contribution of triple junctions to the structure and properties of nanocrystalline materials, Scr. Metall. Mater. 24 (1990) 1347–1350. [51] R. Birringer, H. Gleiter, H.P. Klein, P. Marquardt, Nanocrystalline materials an approach to a novel solid structure with gas-like disorder?, Phys Lett. A 102 (1984) 365–369. [52] K.D. Ralston, N. Birbilis, Effect of grain size on corrosion: a review, Corrosion 66 (2010) 0750051–07500513. [53] L. Liu, Y. Li, F. Wang, Electrochemical corrosion behavior of nanocrystalline materials – a review, J. Mater. Sci. Technol. 26 (2010) 1–14. [54] H.J. Fecht, Intrinsic instability and entropy stabilization of grain boundaries, Phys. Rev. Lett. 65 (1990) 610–613. [55] H.J. Fecht, E. Hellstern, Z. Fu, W.L. Johnson, Nanocrystalline metals prepared by high-energy ball milling, Metall. Trans. A 21 (1990) 2333–2337. [56] P. Zimmer, R. Birringer, Measuring the interface stress of nanocrystalline iron, Appl. Phys. Lett. 92 (2008). [57] R. Kirchheim, X.Y. Huang, P. Cui, R. Birringer, H. Gleiter, Free energy of active atoms in grain boundaries of nanocrystalline copper, nickel and palladium, Nanostruct. Mater. 1 (1992) 167–172. [58] Z.B. Wang, N.R. Tao, W.P. Tong, J. Lu, K. Lu, Diffusion of chromium in nanocrystalline iron produced by means of surface mechanical attrition treatment, Acta Mater. 51 (2003) 4319–4329. [59] H. Mehrer, Diffusion in nanocrystalline materials, Diffusion in Solids, vol. 155, Springer, Berlin Heidelberg, 2007. pp. 593–620. [60] I.V. Belova, G.E. Murch, Diffusion in nanocrystalline materials, J. Phys. Chem. Solids 64 (2003) 873–878. [61] R. Würschum, S. Herth, U. Brossmann, Diffusion in nanocrystalline metals and alloys – a status report, Adv. Eng. Mater. 5 (2003) 365–372. [62] R.K. Gupta, Synthesis and corrosion behaviour of nanocrystalline Fe–Cr alloys, PhD Thesis, Monash University, 2010. [63] H. Jung, A. Alfantazi, An electrochemical impedance spectroscopy and polarization study of nanocrystalline Co and Co–P alloy in 0.1 M H2SO4 solution, Electrochim. Acta 51 (2006) 1806–1814. [64] H. Jung, A. Alfantazi, Corrosion behavior of nanocrystalline Co and Co–P alloys in a NaOH solution, Corrosion 66 (2010) 0350021–03500212. [65] M. Crobu, A. Scorciapino, B. Elsener, A. Rossi, The corrosion resistance of electroless deposited nano-crystalline Ni–P alloys, Electrochim. Acta 53 (2008) 3364–3370. [66] Y. Gao, Z.J. Zheng, M. Zhu, C.P. Luo, Corrosion resistance of electrolessly deposited Ni–P and Ni–W–P alloys with various structures, Mater. Sci. Eng. A 381 (2004) 98–103. [67] R. Rofagha, U. Erb, D. Ostrander, G. Palumbo, K.T. Aust, The effects of grain size and phosphorus on the corrosion of nanocrystalline Ni–P alloys, Nanostruct. Mater. 2 (1993) 1–10. [68] C. Pan, L. Liu, Y. Li, B. Zhang, F. Wang, The electrochemical corrosion behavior of nanocrystalline 304 stainless steel prepared by magnetron sputtering, J. Electrochem. Soc. 159 (2012) C453–C460. [69] D.Y. Li, Electron work function at grain boundary and the corrosion behavior of nanocrystalline metallic materials, in: Materials Research Society Symposium Proceedings, vol. 887, 2006, pp. 227–235.

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx [70] E.J.W. Verwey, Electrolytic conduction of a solid insulator at high fields. The formation of the anodic oxide film on aluminium, Physica 2 (1935) 1059– 1063. [71] N. Cabrera, N.F. Mott, Theory of the oxidation of metals, Rep. Progr. Phys. 12 (1949) 163–184. [72] N.F. Mott, The theory of the formation of protective oxide films on metals – III, Trans. Faraday Soc. 43 (1947) 429–434. [73] F.P. Fehlner, N.F. Mott, Low-temperature oxidation, Oxid. Met. 2 (1970) 59– 99. [74] C.Y. Chao, L.F. Lin, D.D. Macdonald, Point defect model for anodic passive films, J. Electrochem. Soc. 128 (1981) 1187–1194. [75] D.D. Macdonald, Point defect model for the passive state, J. Electrochem. Soc. 139 (1992) 3434–3449. [76] L.F. Lin, C.Y. Chao, D.D. Macdonald, Point defect model for anodic passive films – 2. Chemical breakdown and pit initiation, J. Electrochem. Soc. 128 (1981) 1194–1198. [77] D.D. MacDonald, The history of the point defect model for the passive state: a brief review of film growth aspects, Electrochim. Acta 56 (2011) 1761–1772. [78] M. Urquidi, D.D. Macdonald, Solute/vacancy interaction model for the effect of minor alloying elements on the breakdown of the passive films, Electrochem. Soc. Extended Abstr. 85–1 (1985) 29–30. [79] M. Urquidi-Macdonald, D.D. Macdonald, Theoretical analysis of the effects of alloying elements on distribution functions of passivity breakdown, J. Electrochem. Soc. 136 (1989) 961–967. [80] J.W. Schultze, M.M. Lohrengel, Stability, reactivity and breakdown of passive films. Problems of recent and future research, Electrochim. Acta 45 (2000) 2499–2513. [81] D.A. Jones, Principles and Prevention of Corrosion, Prentice Hall, Upper Saddle River, NJ, 1996. [82] I. Olefjord, The passive state of stainless steels, Mater. Sci. Eng. 42 (1980) 161–171. [83] D.D. Macdonald, Passivity – the key to our metals-based civilization, Pure Appl. Chem. 71 (1999) 951–978. [84] D.D. Macdonald, M. Urquidi-Macdonald, Theory of steady-state passive films, J. Electrochem. Soc. 137 (1990) 2395–2402. [85] H.H. Uhlig, Passivity in metals and alloys, Corros. Sci. 19 (1979) 777–791. [86] N. Sato, An overview on the passivity of metals, Corros. Sci. 31 (1990) 1–19. [87] P. Monnartz, Beitrag zum Studium der Eisenchromlegierungen unter besonderer Berücksichtigung der Säurebeständigkeit, Metallurgie 8 (1911) 161–170. [88] N.D. Tomashov, Passivity and corrosion resistance of metal systems, Corros. Sci. 4 (1964) 315–334. [89] K. Osozawa, H.J. Engell, The anodic polarization curves of iron–nickel– chromium alloys, Corros. Sci. 6 (1966) 389–393. [90] J. Horvath, H.H. Uhlig, Critical potentials for pitting corrosion of Ni, Cr–Ni, Cr– Fe, and related stainless steels, J. Electrochem. Soc. 115 (1968) 791–795. [91] I. Odnevall Wallinder, J. Lu, S. Bertling, C. Leygraf, Release rates of chromium and nickel from 304 and 316 stainless steel during urban atmospheric exposure – a combined field and laboratory study, Corros. Sci. 44 (2002) 2303–2319. [92] D. Hamm, K. Ogle, C.O. Olsson, S. Weber, D. Landolt, Passivation of Fe–Cr alloys studied with ICP-AES and EQCM, Corros. Sci. 44 (2002) 1443–1456. [93] V.S. Rao, L.K. Singhal, Corrosion behavior and passive film chemistry of 216L stainless steel in sulphuric acid, J. Mater. Sci. 44 (2009) 2327–2333. [94] M. Prazak, D. Kuchynka, T.X. Sen, Selectetive dissolution of Fe, Cr and Ni from stainless steel, in: EUROCOR ‘77, Eur Congr on Met Corros, 92nd Event of the Eur Fed of Corros, 1977, pp. 287–291. [95] N. Ramasubramanian, N. Preocanin, R.D. Davidson, Analysis of passive films on stainless steel by cyclic voltammetry and auger spectroscopy, J. Electrochem. Soc. 132 (1985) 793–798. [96] M. Seo, N. Sato, Theoretical approach to corrosion resistivity of alloys, Langmuir 3 (1987) 917–921. [97] A.R. Brooks, C.R. Clayton, K. Doss, Y.C. Lu, On the role of Cr in the passivity of stainless steel, J. Electrochem. Soc. 133 (1986) 2459–2464. [98] B.O. Elfström, The effect of chloride ions on passive layers on stainless steels, Mater. Sci. Eng. 42 (1980) 173–180. [99] I. Olefjord, B. Brox, U. Jelvestam, Surface composition of stainless steels during anodic dissolution and passivation studied by ESCA, J. Electrochem. Soc. 132 (1985) 2854–2861. [100] I. Olefjord, B. Brox, Quantitative ESCA analysis of the passive state of an Fe–Cr alloy and an Fe–Cr–Mo alloy, Thin Films Sci. Technol. (1983) 561–570. [101] I. Olefjord, B. Brox, U. Jelvestam, Surface composition of stainless steels during anodic dissolution and passivation studied by ESCA, Proc. – Electrochem. Soc. 84–9 (1984) 388–401. [102] K. Hashimoto, K. Asami, An X-ray photo-electron spectroscopic study of the passivity of ferritic 19Cr stainless steels in 1 NHCl, Corros. Sci. 19 (1979) 251– 260. [103] K. Teramoto, K. Asami, K. Hashimoto, Composition of passive films on ferritic 30Cr stainless steel in H2SO4, Corros. Eng. (Tokyo) 27 (1978) 57–61. [104] G. Okamoto, Passive film of 18–8 stainless steel structure and its function, Corros. Sci. 13 (1973) 471–489. [105] H. Fischmeister, U. Roll, Passive layers on stainless steels: a survey of surface analysis results, Fresenius’ Zeitschrift für Analytische Chemie 319 (1984) 639–645.

13

[106] K. Asami, K. Hashimoto, S. Shimodaira, An XPS study of the passivity of a series of iron–chromium alloys in sulphuric acid, Corros. Sci. 18 (1978) 151– 160. [107] H.J. Rocha, G. Lennartz, Arcg Eisenhuttenwes 16 (1955) 117. [108] K. Sieradzki, R.C. Newman, Percolation model for passivation in stainless steels, J. Electrochem. Soc. 133 (1986) 1979–1980. [109] M.P. Ryan, S. Fujimoto, G.E. Thompson, R.C. Newman, Disorder and structural relaxation in passive films on Fe–Cr alloys, Mater. Sci. Forum 185–188 (1995) 233–240. [110] S. Qian, R.C. Newman, R.A. Cottis, K. Sieradzki, Computer simulation of alloy passivation and activation, Corros. Sci. 31 (1990) 621–626. [111] Q. Song, R.C. Newman, R.A. Cottis, K. Sieradzki, Validation of a percolation model for passivation of Fe–Cr alloys. Two-dimensional computer simulations, J. Electrochem. Soc. 137 (1990) 435–439. [112] D.E. Williams, R.C. Newman, Q. Song, R.G. Kelly, Passivity breakdown and pitting corrosion of binary alloys, Nature 350 (1991) 216–219. [113] C.R. Clayton, Y.C. Lu, Bipolar model of the passivity of stainless steel: the role of Mo addition, J. Electrochem. Soc. 133 (1986) 2465–2473. [114] P. Schmuki, H. Boehni, Metastable pitting and semiconductive properties of passive films, J. Electrochem. Soc. 139 (1992) 1908–1913. [115] N.B. Hakiki, S. Boudin, B. Rondot, M. Da Cunha Belo, The electronic structure of passive films formed on stainless steels, Corros. Sci. 37 (1995) 1809–1822. [116] H. Tsuchiya, S. Fujimoto, O. Chihara, T. Shibata, Semiconductive behavior of passive films formed on pure Cr and Fe–Cr alloys in sulfuric acid solution, Electrochim. Acta 47 (2002) 4357–4366. [117] E. Cho, H. Kwon, D.D. Macdonald, Photoelectrochemical analysis on the passive film formed on Fe–20Cr in pH 8.5 buffer solution, Electrochim. Acta 47 (2002) 1661–1668. [118] S.B. Arya, V.S. Raja, A.N. Tiwari, A comparative study of semiconducting behavior of passive film of high nitrogen and Ni and Mn free stainless steels in 3.5 wt.% NaCl, Adv. Mater. Res. 794 (2013) 626–631. [119] E.E. Oguzie, J. Li, Y. Liu, D. Chen, Y. Li, K. Yang, F. Wang, The effect of Cu addition on the electrochemical corrosion and passivation behavior of stainless steels, Electrochim. Acta 55 (2010) 5028–5035. [120] A. Irhzo, Y. Segui, N. Bui, F. Dabosi, The role of alloyed tungsten on the conductivity of stainless steel passive layers, Corros. Sci. 26 (1986) 769–780. [121] K. Osozawa, Y. Fukase, K. Yokota, Effects of minor elements on corrosion resistance of austenitic stainless steel in dilute sulphuric acid, Corros. Eng. (Tokyo) 20 (1971) 11–16. [122] K. Osozawa, The effects of nitrogen on the passivity and the pitting corrosion of stainless steels, Zairyo to Kankyo/Corros. Eng. 47 (1998) 561–569. [123] A. Fattah-alhosseini, F. Soltani, F. Shirsalimi, B. Ezadi, N. Attarzadeh, The semiconducting properties of passive films formed on AISI 316 L and AISI 321 stainless steels: a test of the point defect model (PDM), Corros. Sci. 53 (2011) 3186–3192. [124] H.H. Ge, X.M. Xu, L. Zhao, F. Song, J. Shen, G.D. Zhou, Semiconducting behavior of passive film formed on stainless steel in borate buffer solution containing sulfide, J. Appl. Electrochem. 41 (2011) 519–525. [125] R. Babic´, M. Metikoš-Hukovic´, Semiconducting properties of passive films on AISI 304 and 316 stainless steels, J. Electroanal. Chem. 358 (1993) 143–160. [126] G.T. Burstein, Passivity and localized corrosion, Shreir’s Corros. (2010) 731– 752. [127] G.T. Burstein, J.J. Moloney, Cyclic thermammetry, Electrochem. Commun. 6 (2004) 1037–1041. [128] G.T. Burstein, C. Liu, R.M. Souto, S.P. Vines, Origins of pitting corrosion, Corros. Eng. Sci. Technol. 39 (2004) 25–30. [129] G.S. Frankel, Pitting corrosion of metals: a review of the critical factors, J. Electrochem. Soc. 145 (1998) 2186–2198. [130] G.T. Burstein, P.C. Pistorius, S.P. Mattin, The nucleation and growth of corrosion pits on stainless steel, Corros. Sci. 35 (1993) 57–62. [131] D.E. Williams, J. Stewart, P.H. Balkwill, The nucleation, growth and stability of micropits in stainless steel, Corros. Sci. 36 (1994) 1213–1235. [132] D.E. Williams, C. Westcott, M. Fleischmann, Stochastic models of pitting corrosion of stainless steel. I. Modeling of the initiation and growth of pits at constant potential, J. Electrochem. Soc. 132 (1985) 1796–1804. [133] P.C. Pistorius, G.T. Burstein, Aspects of the effects of electrolyte composition on the occurrence of metastable pitting on stainless steel, Corros. Sci. 36 (1994) 525–538. [134] P.C. Pistorius, G.T. Burstein, Growth of corrosion pits on stainless steel in chloride solution containing dilute sulphate, Corros. Sci. 33 (1992) 1885– 1897. [135] G.S. Frankel, L. Stockert, F. Hunkeler, H. Boehni, Metastable pitting of stainless steel, Corrosion 43 (1987) 429–436. [136] N.J. Laycock, M.H. Moayed, R.C. Newman, Metastable pitting and the critical pitting temperature, J. Electrochem. Soc. 145 (1998) 2622–2628. [137] N.D. Greene, M.G. Fontana, A critical analysis of pitting corrosion, Corrosion 15 (1959) 41–47. [138] T.P. Hoar, W.R. Jacob, Breakdown of passivity of stainless steel by halide ions, Nature 216 (1967) 1299–1301. [139] T.P. Hoar, The production and breakdown of the passivity of metals, Corros. Sci. 7 (1967) 341–355. [140] N. Sato, A theory for breakdown of anodic oxide films on metals, Electrochim. Acta 16 (1971) 1683–1692. [141] G.S. Frankel, N. Sridhar, Understanding localized corrosion, Mater. Today 11 (2008) 38–44.

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

14

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx

[142] S.T. Pride, J.R. Scully, J.L. Hudson, Metastable pitting of aluminum and criteria for the transition to stable pit growth, J. Electrochem. Soc. 141 (1994) 3028– 3040. [143] R.K. Gupta, N.L. Sukiman, M.K. Cavanaugh, B.R.W. Hinton, C.R. Hutchinson, N. Birbilis, Metastable pitting characteristics of aluminium alloys measured using current transients during potentiostatic polarisation, Electrochim. Acta 66 (2012) 245–254. [144] R.K. Gupta, B.R.W. Hinton, N. Birbilis, The effect of chromate on the pitting susceptibility of AA7075-T651 studied using potentiostatic transients, Corros. Sci. 82 (2014) 197–207. [145] R.K. Gupta, A. Deschamps, M.K. Cavanaugh, S.P. Lynch, N. Birbilis, Relating the early evolution of microstructure with the electrochemical response and mechanical performance of a Cu-rich and Cu-lean 7xxx aluminum alloy, J. Electrochem. Soc. 159 (2012) C492–C502. [146] C. Punckt, M. Bölscher, H.H. Rotermund, A.S. Mikhailov, L. Organ, N. Budiansky, J.R. Scully, J.L. Hudson, Sudden onset of pitting corrosion on stainless steel as a critical phenomenon, Science 305 (2004) 1133–1136. [147] R.C. Newman, E.M. Franz, Growth and repassivation of single corrosion pits in stainless steel, Corrosion 40 (1984) 325–330. [148] C.O.A. Olsson, D. Landolt, Passive films on stainless steels—chemistry, structure and growth, Electrochim. Acta 48 (2003) 1093–1104. [149] M.P. Ryan, N.J. Laycock, R.C. Newman, H.S. Isaacs, The pitting behavior of iron–chromium thin film alloys in hydrochloric acid, J. Electrochem. Soc. 145 (1998) 1566–1571. [150] J. Stewart, D.E. Williams, The initiation of pitting corrosion on austenitic stainless steel: on the role and importance of sulphide inclusions, Corros. Sci. 33 (457–463) (1992) 465–474. [151] T.L.S.L. Wijesinghe, D.J. Blackwood, Real time pit initiation studies on stainless steels: the effect of sulphide inclusions, Corros. Sci. 49 (2007) 1755–1764. [152] M.A. Baker, J.E. Castle, The initiation of pitting corrosion at MnS inclusions, Corros. Sci. 34 (1993) 667–682. [153] G. Wranglen, Pitting and sulphide inclusions in steel, Corros. Sci. 14 (1974) 331–349. [154] S.E. Lott, R.C. Alkire, Role of inclusions on initiation of crevice corrosion of stainless steel: I. Experimental studies, J. Electrochem. Soc. 136 (1989) 973– 979. [155] G.S. Eklund, Initiation of pitting at sulfide inclusions in stainless steel, J. Electrochem. Soc. 121 (1974) 467–473. [156] M.P. Ryan, D.E. Williams, R.J. Chater, B.M. Hutton, D.S. McPhail, Why stainless steel corrodes, Nature 415 (2002) 770–774. [157] Q. Meng, G.S. Frankel, H.O. Colijn, S.H. Goss, Stainless-steel corrosion and MnS inclusions, Nature 424 (2003) 389–390. [158] D.E. Williams, M.R. Kilburn, J. Cliff, G.I.N. Waterhouse, Composition changes around sulphide inclusions in stainless steels, and implications for the initiation of pitting corrosion, Corros. Sci. 52 (2010) 3702–3716. [159] R.M. Fernández-Domene, E. Blasco-Tamarit, D.M. García-García, J. GarcíaAntón, Effect of alloying elements on the electronic properties of thin passive films formed on carbon steel, ferritic and austenitic stainless steels in a highly concentrated LiBr solution, Thin Solid Films 558 (2014) 252–258. [160] Y.C. Lu, M.B. Ives, C.R. Clayton, Synergism of alloying elements and pitting corrosion resistance of stainless steels, Corros. Sci. 35 (1993) 89–96. [161] A.U. Malik, N.A. Siddiqi, S. Ahmad, I.N. Andijani, The effect of dominant alloy additions on the corrosion behavior of some conventional and high alloy stainless steels in seawater, Corros. Sci. 37 (1995) 1521–1535. [162] K. Oh, S. Ahn, K. Eom, K. Jung, H. Kwon, Observation of passive films on Fe– 20Cr–xNi (x = 0, 10, 20 wt.%) alloys using TEM and Cs-corrected STEM–EELS, Corros. Sci. 79 (2014) 34–40. [163] J.R. Galvele, J.B. Lumsden, R.W. Staehle, Effect of molybdenum on the pitting potential of high purity 18% Cr ferritic stainless steel, J. Electrochem. Soc. 125 (1978) 1204–1208. [164] G.O. Ilevbare, G.T. Burstein, Role of alloyed molybdenum in the inhibition of pitting corrosion in stainless steels, Corros. Sci. 43 (2001) 485–513. [165] M. Kaneko, H.S. Isaacs, Effects of molybdenum on the pitting of ferritic- and austenitic-stainless steels in bromide and chloride solutions, Corros. Sci. 44 (2002) 1825–1834. [166] P.I. Marshall, G.T. Burstein, Effects of alloyed molybdenum on the kinetics of repassivation on austenitic stainless steels, Corros. Sci. 24 (1984) 463–478. [167] K. Sugimoto, Y. Sawada, Role of alloyed molybdenum in austenitic stainless steel in the inhibition of pitting in neutral halide solutions, Corrosion 32 (1976) 347–352. [168] H.Y. Ha, T.H. Lee, S.J. Kim, Role of nitrogen in the active–passive transition behavior of binary Fe–Cr alloy system, Electrochim. Acta 80 (2012) 432–439. [169] H.Y. Ha, T.H. Lee, S.J. Kim, Synergistic effect of Ni and N on improvement of pitting corrosion resistance of high nitrogen stainless steels, Corros. Eng. Sci. Technol. 49 (2014) 82–86. [170] R.F.A. Jargelius-Pettersson, Electrochemical investigation of the influence of nitrogen alloying on pitting corrosion of austenitic stainless steels, Corros. Sci. 41 (1999) 1639–1664. [171] R. Goetz, J. Laurent, D. Landolt, The influence of minor alloying elements on the passivation behaviour of iron–chromium alloys in HCl, Corros. Sci. 25 (1985) 1115–1126. [172] A. Tomio, M. Sagara, T. Doi, H. Amaya, N. Otsuka, T. Kudo, Role of alloyed copper on corrosion resistance of austenitic stainless steel in H2S–Clenvironment, Corros. Sci. 81 (2014) 144–151.

[173] G. Herbsleb, Influence of SO2, H2S and CO on pitting corrosion of austenitic chromium–nickel stainless steels with up to 4 wt.% molybdenum in 1 M NaCl, Werkstoffe und Korrosion 33 (1982) 334–340. [174] R. Merello, F.J. Botana, J. Botella, M.V. Matres, M. Marcos, Influence of chemical composition on the pitting corrosion resistance of non-standard low-Ni high-Mn–N duplex stainless steels, Corros. Sci. 45 (2003) 909–921. [175] B.A. Betts, S.A. Hebdon, Determining PREN for high-chromium cast irons, Mater. Perform. 39 (2000) 78–81. [176] R.A. Perren, T.A. Suter, P.J. Uggowitzer, L. Weber, R. Magdowski, H. Böhni, M.O. Speidel, Corrosion resistance of super duplex stainless steels in chloride ion containing environments: investigations by means of a new microelectrochemical method. I. Precipitation-free states, Corros. Sci. 43 (2001) 707–726. [177] R.C. Newman, H.S. Isaacs, Dissolution and passivation kinetics of Fe–Cr–Ni alloys during localized corrosion, Thin Films Sci. Technol. (1983) 269–274. [178] R.C. Newman, M.A.A. Ajjawi, H. Ezuber, S. Turgoose, An experimental confirmation of the pitting potential model of Galvele, Corros. Sci. 28 (1988) 471–477. [179] R.C. Newman, Protection potentials for pitting of stainless steel in neutral chloride solutions, Corros. Sci. 23 (1983) 1045–1046. [180] R. Ke, R. Alkire, Initiation of corrosion pits at inclusions on 304 stainless steel, J. Electrochem. Soc. 142 (1995) 4056–4062. [181] W.R. Cieslak, D.J. Duquette, Electrochemical study of the pit initiation resistance of ferritic stainless steel, J. Electrochem. Soc. 132 (1985) 533–537. [182] B. Malki, T. Souier, B. Baroux, Influence of the alloying elements on pitting corrosion of stainless steels: a modeling approach, J. Electrochem. Soc. 155 (2008) C583–C587. [183] G. Salvago, G. Fumagalli, The distribution of stainless steel breakdown potentials: experimental method and the effect of metallurgical conditions, Corros. Sci. 33 (1992) 985–995. [184] W.E. White, Observations of the influence of microstructure on corrosion of welded conventional and stainless steels, Mater. Charact. 28 (1992) 349–358. [185] M.P. Ryan, D.E. Williams, R.J. Chater, B.M. Hutton, D.S. McPhail, Stainless-steel corrosion and MnS inclusions: commentary, Nature 424 (2003) 390. [186] D. Fabijanic, A. Taylor, K.D. Ralston, M.X. Zhang, N. Birbilis, Influence of surface mechanical attrition treatment attrition media on the surface contamination and corrosion of magnesium, Corrosion 69 (2013) 527–535. [187] J. Duchene, M. Terraillon, M. Pailly, R.F. and D.C. reactive sputtering for crystalline and amorphous VO2 thin film deposition, Thin Solid Films 12 (1972) 231–234. [188] G. Meng, Y. Li, F. Wang, The corrosion behavior of Fe–10Cr nanocrystalline coating, Electrochim. Acta 51 (2006) 4277–4284. [189] K. Wasa, S. Hayakawa, Special features of thin compound films prepared by magnetron sputtering, Surf. Sci. 86 (1979) 300–307. [190] H. Uchida, Y. Matsumura, H. Uchida, H. Kaneko, Progress in thin films of giant magnetostrictive alloys, J. Magn. Magn. Mater. 239 (2002) 540–545. [191] T.W. Barbee Jr., B.E. Jacobson, D.L. Keith, Microstructure of amorphous 304 stainless steel–carbon alloys synthesized by magnetron sputter deposition, Thin Solid Films 63 (1979) 143–150. [192] R.B. Inturi, Z. Szklarska-Smialowska, Localized corrosion of nanocrystalline 304 type stainless steel films, Corrosion 48 (1992) 398–403. [193] M. Schneider, W. Zeiger, D. Scharnweber, H. Worch, Electrochemical behavior of nanocrystalline FeAl8 and FeCr10 alloys, Mater. Sci. Forum 225–227 (1996) 819–824. [194] W. Ye, Y. Li, F. Wang, Effects of nanocrystallization on the corrosion behavior of 309 stainless steel, Electrochim. Acta 51 (2006) 4426–4432. [195] L. Liu, Y. Li, F. Wang, Pitting mechanism on an austenite stainless steel nanocrystalline coating investigated by electrochemical noise and in-situ AFM analysis, Electrochim. Acta 54 (2008) 768–780. [196] L. Liu, Y. Li, F. Wang, Influence of nanocrystallization on pitting corrosion behavior of an austenitic stainless steel by stochastic approach and in situ AFM analysis, Electrochim. Acta 55 (2010) 2430–2436. [197] C. Pan, L. Liu, Y. Li, F. Wang, Pitting corrosion of 304ss nanocrystalline thin film, Corros. Sci. 73 (2013) 32–43. [198] S.G. Wang, M. Sun, K. Long, The enhanced even and pitting corrosion resistances of bulk nanocrystalline steel in HCl solution, Steel Res. Int. 83 (2012) 800–807. [199] M. Eskandari, M. Yeganeh, M. Motamedi, Investigation in the corrosion behaviour of bulk nanocrystalline 316L austenitic stainless steel in NaCl solution, Micro Nano Lett. 7 (2012) 380–383. [200] C. Pan, L. Liu, Y. Li, S. Wang, F. Wang, Passive film growth mechanism of nanocrystalline 304 stainless steel prepared by magnetron sputtering and deep rolling techniques, Electrochim. Acta 56 (2011) 7740–7748. [201] C. Suryanarayana, Mechanical alloying and milling, Prog. Mater. Sci. 46 (2001) 1–184. [202] B.S. Murty, S. Ranganathan, Novel materials synthesis by mechanical alloying/milling, Int. Mater. Rev. 43 (1998) 101–141. [203] J. Benjamin, Dispersion strengthened superalloys by mechanical alloying, Metall. Trans. 1 (1970) 2943–2951. [204] C.C. Koch, Synthesis of nanostructured materials by mechanical milling: problems and opportunities, Nanostruct. Mater. 9 (1997) 13–22. [205] C.C. Koch, Top-down synthesis of nanostructured materials: mechanical and thermal processing methods, Rev. Adv. Mater. Sci. 5 (2003) 91–99. [206] R. Gupta, R.K. Singh Raman, C.C. Koch, Grain growth behaviour and consolidation of ball-milled nanocrystalline Fe–10Cr alloy, Mater. Sci. Eng. A 494 (2008) 253–256.

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041

R.K. Gupta, N. Birbilis / Corrosion Science xxx (2014) xxx–xxx [207] R.K. Gupta, R.K. Singh Raman, C.C. Koch, Electrochemical characteristics of nano and microcrystalline Fe–Cr alloys, J. Mater. Sci. 47 (2012) 6118–6124. [208] R.Z. Valiev, T.G. Langdon, Principles of equal-channel angular pressing as a processing tool for grain refinement, Prog. Mater. Sci. 51 (2006) 881–981. [209] M. Furukawa, Z. Horita, M. Nemoto, T.G. Langdon, Processing of metals by equal-channel angular pressing, J. Mater. Sci. 36 (2001) 2835–2843. [210] Z. Horita, T. Fujinami, M. Nemoto, T.G. Langdon, Improvement of mechanical properties for Al alloys using equal-channel angular pressing, J. Mater. Process. Technol. 117 (2001) 288–292. [211] M. Furukawa, Z. Horita, M. Nemoto, R.Z. Valiev, T.G. Langdon, Microhardness measurements and the Hall–Petch relationship in an Al–Mg alloy with submicrometer grain size, Acta Mater. 44 (1996) 4619–4629. [212] W.J. Kim, C.W. An, Y.S. Kim, S.I. Hong, Mechanical properties and microstructures of an AZ61 Mg Alloy produced by equal channel angular pressing, Scr. Mater. 47 (2002) 39–44. [213] C.X. Huang, Y.L. Gao, G. Yang, S.D. Wu, G.Y. Li, S.X. Li, Bulk nanocrystalline stainless steel fabricated by equal channel angular pressing, J. Mater. Res. 21 (2006) 1687–1692. [214] C.X. Huang, G. Yang, B. Deng, S.D. Wu, S.X. Li, Z.F. Zhang, Formation mechanism of nanostructures in austenitic stainless steel during equal channel angular pressing, Philos. Mag. 87 (2007) 4949–4971. [215] Z.J. Zheng, Y. Gao, Y. Gui, M. Zhu, Corrosion behaviour of nanocrystalline 304 stainless steel prepared by equal channel angular pressing, Corros. Sci. 54 (2012) 60–67. [216] C.T. Kwok, F.T. Cheng, H.C. Man, W.H. Ding, Corrosion characteristics of nanostructured layer on 316L stainless steel fabricated by cavitationannealing, Mater. Lett. 60 (2006) 2419–2422. [217] K. Lu, J. Lu, Surface nanocrystallization (SNC) of metallic materialspresentation of the concept behind a new approach, J. Mater. Sci. Technol. 15 (1999) 193–197. [218] N.R. Tao, J. Lu, K. Lu, Surface nanocrystallization by surface mechanical attrition treatment, Mater. Sci. Forum 579 (2008) 91–108. [219] K. Lu, J. Lu, Nanostructured surface layer on metallic materials induced by surface mechanical attrition treatment, Mater. Sci. Eng. A 375–377 (2004) 38–45. [220] W.P. Tong, N.R. Tao, Z.B. Wang, J. Lu, K. Lu, Nitriding iron at lower temperatures, Science 299 (2003) 686–688. [221] L. Waltz, D. Retraint, A. Roos, Semi-massive nanocrystallised composites: from the process to mechanical and microstructural investigations, vol. 762, 2013, pp. 487–492. [222] Y. Samih, B. Beausir, B. Bolle, T. Grosdidier, In-depth quantitative analysis of the microstructures produced by surface mechanical attrition treatment (SMAT), Mater. Charact. 83 (2013) 129–138. [223] X.Y. Wang, D.Y. Li, Mechanical and electrochemical behavior of nanocrystalline surface of 304 stainless steel, Electrochim. Acta 47 (2002) 3939–3947. [224] X.Y. Wang, D.Y. Li, Mechanical, electrochemical and tribological properties of nano-crystalline surface of 304 stainless steel, Wear 255 (2003) 836–845. [225] Y.w. Hao, B. Deng, C. Zhong, Y.m. Jiang, J. Li, Effect of surface mechanical attrition treatment on corrosion behavior of 316 stainless steel, J. Iron Steel Res. Int. 16 (2009) 68–72. [226] T. Bai, K. Guan, Evaluation of stress corrosion cracking susceptibility of nanocrystallized stainless steel 304L welded joint by small punch test, Mater. Des. 52 (2013) 849–860. [227] T. Bai, P. Chen, K. Guan, Evaluation of stress corrosion cracking susceptibility of stainless steel 304L with surface nanocrystallization by small punch test, Mater. Sci. Eng. A 561 (2013) 498–506. [228] T. Wang, J. Yu, B. Dong, Surface nanocrystallization induced by shot peening and its effect on corrosion resistance of 1Cr18Ni9Ti stainless steel, Surf. Coat. Technol. 200 (2006) 4777–4781. [229] A. Toppo, R. Kaul, M.G. Pujar, U. Kamachi Mudali, L.M. Kukreja, Enhancement of corrosion resistance of type 304 stainless steel through a novel thermomechanical surface treatment, J. Mater. Eng. Perform. 22 (2013) 632–639. [230] T. Balusamy, S. Kumar, T.S.N. Sankara Narayanan, Effect of surface nanocrystallization on the corrosion behaviour of AISI 409 stainless steel, Corros. Sci. 52 (2010) 3826–3834. [231] T. Balusamy, T.S.N. Sankara Narayanan, K. Ravichandran, I.S. Park, M.H. Lee, Influence of surface mechanical attrition treatment (SMAT) on the corrosion behaviour of AISI 304 stainless steel, Corros. Sci. 74 (2013) 332–344. [232] N.J. Laycock, S.P. While, J.S. Noh, P.T. Wilson, R.C. Newman, Perforated covers for propagating pits, J. Electrochem. Soc. 145 (1998) 1101–1108. [233] J. Jun, K. Holguin, G.S. Frankel, Pitting corrosion of very clean 304 stainless steel, Corrosion (2013). [234] R. Kirchheim, B. Heine, H. Fischmeister, S. Hofmann, H. Knote, U. Stolz, The passivity of iron–chromium alloys, Corros. Sci. 29 (1989). [235] S. Tao, D.Y. Li, Tribological, mechanical and electrochemical properties of nanocrystalline copper deposits produced by pulse electrodeposition, Nanotechnology 17 (2006) 65–78. [236] M. Lakatos-Varsányi, F. Falkenberg, I. Olefjord, The influence of phosphate on repassivation of 304 stainless steel in neutral chloride solution, Electrochim. Acta 43 (1998) 187–197.

15

[237] G.S. Frankel, M.A. Russak, C.V. Jahnes, M. Mirzamaani, V.A. Brusic, Pitting of sputtered aluminum alloy thin films, J. Electrochem. Soc. 136 (1989) 1243– 1244. [238] V. Brusic, G.S. Frankel, B.M. Rush, A.G. Schrott, C. Jahnes, M.A. Russak, T. Petersen, Corrosion and passivation of Fe and FeN films, J. Electrochem. Soc. 139 (1992) 1530–1535. [239] G.S. Frankel, R.C. Newman, C.V. Jahnes, M.A. Russak, On the pitting resistance of sputter-deposited aluminum alloys, J. Electrochem. Soc. 140 (1993) 2192– 2197. [240] G.S. Frankel, The growth of 2-D pits in thin film aluminum, Corros. Sci. 30 (1990) 1203–1218. [241] F. Tang, D.S. Gianola, M.P. Moody, K.J. Hemker, J.M. Cairney, Observations of grain boundary impurities in nanocrystalline Al and their influence on microstructural stability and mechanical behaviour, Acta Mater. 60 (2012) 1038–1047. [242] G.S. Frankel, X. Chen, R.K. Gupta, S. Kandasamy, N. Birbilis, Effect of vacuum system base pressure on corrosion resistance of sputtered Al thin films, J. Electrochem. Soc. 161 (2014) C195–C200. [243] H.H. Uhlig, Structure and growth of thin films on metals exposed to oxygen, Corros. Sci. 7 (1967) 325–339. [244] H.H. Uhlig, P.F. King, Passivity in the iron–chromium binary alloys, J. Phys. Chem. 63 (1959) 2026–2032. [245] H. Zhang, R.L. Penn, R.J. Hamers, J.F. Banfiel, Enhanced adsorption of molecules on surfaces of nanocrystalline particles, J. Phys. Chem. B 103 (1999) 4656. [246] C. Lemier, J. Weissmuller, Grain boundary segregation, stress and stretch: effects on hydrogen absorption in nanocrystalline palladium, Acta Mater. 55 (2007) 1241. [247] M.A. Meyers, A. Mishra, D.J. Benson, Mechanical properties of nanocrystalline materials, Prog. Mater. Sci. 51 (2006) 427–556. [248] T.R. Malow, C.C. Kqch, Mechanical properties, ductility, and grain size of nanocrystalline iron produced by mechanical attrition, Metall. Mater. Trans. A: Phys. Metall. Mater. Sci. 29 (1998) 2285–2295. [249] P.G. Sanders, J.A. Eastman, J.R. Weertman, Elastic and tensile behavior of nanocrystalline copper and palladium, Acta Mater. 45 (1997) 4019–4025. [250] V.Y. Gertsman, R. Birringer, On the room-temperature grain growth in nanocrystalline copper, Scr. Metall. Mater. 30 (1994) 577–581. [251] L. Lu, N.R. Tao, L.B. Wang, B.Z. Ding, K. Lu, Grain growth and strain release in nanocrystalline copper, J. Appl. Phys. 89 (2001) 6408–6414. [252] C. Suryanarayana, C.C. Koch, Nanocrystalline materials – current research and future directions, Hyperfine Interact. 130 (2000) 5–44. [253] C.C. Koch, D.G. Morris, K. Lu, A. Inoue, Ductility of nanostructured materials, MRS Bull. 24 (1999) 54–58. [254] I.A. Ocid’Ko, T.G. Langdon, Enhanced ductility of nanocrystalline and ultrafine-grained metals, Rev. Adv. Mater. Sci. 30 (2012) 103–111. [255] K.M. Youssef, R.O. Scattergood, K.L. Murty, J.A. Horton, C.C. Koch, Ultrahigh strength and high ductility of bulk nanocrystalline copper, Appl. Phys. Lett. 87 (2005). [256] T. Chookajorn, H.A. Murdoch, C.A. Schuh, Design of stable nanocrystalline alloys, Science 337 (2012) 951–954. [257] J.R. Weertman, Retaining the nano in nanocrystalline alloys, Science 337 (2012) 921–922. [258] K.A. Darling, B.K. VanLeeuwen, J.E. Semones, C.C. Koch, R.O. Scattergood, L.J. Kecskes, S.N. Mathaudhu, Stabilized nanocrystalline iron-based alloys: guiding efforts in alloy selection, Mater. Sci. Eng. A 528 (2011) 4365–4371. [259] C.C. Koch, R.O. Scattergood, B.K. VanLeeuwen, K.A. Darling, Thermodynamic stabilization of grain size in nanocrystalline metals, vol. 715–716, 2012, pp. 323–328. [260] R.K. Gupta, R.K.S. Raman, C.C. Koch, Grain growth behaviour and consolidation of ball milled nanocrystalline iron–chromium alloys, in: TMS Annual Conference, vol. 1, 2008, pp. 151–157. [261] K.A. Darling, R.N. Chan, P.Z. Wong, J.E. Semones, R.O. Scattergood, C.C. Koch, Grain-size stabilization in nanocrystalline FeZr alloys, Scr. Mater. 59 (2008) 530–533. [262] G. Hinds, L. Wickström, K. Mingard, A. Turnbull, Impact of surface condition on sulphide stress corrosion cracking of 316L stainless steel, Corros. Sci. 71 (2013) 43–52. [263] A. Turnbull, K. Mingard, J.D. Lord, B. Roebuck, D.R. Tice, K.J. Mottershead, N.D. Fairweather, A.K. Bradbury, Sensitivity of stress corrosion cracking of stainless steel to surface machining and grinding procedure, Corros. Sci. 53 (2011) 3398–3415. [264] S. Ghosh, V. Kain, Microstructural changes in AISI 304L stainless steel due to surface machining: effect on its susceptibility to chloride stress corrosion cracking, J. Nucl. Mater. 403 (2010) 62–67. [265] E. Klar, P.K. Samal, Powder Metallurgy Stainless Steels: Processing, Microstructures, and Properties, ASM International, Materials Park, OH, 2007. [266] M.H. Enayati, M.R. Bafandeh, S. Nosohian, Ball milling of stainless steel scrap chips to produce nanocrystalline powder, J. Mater. Sci. Technol. 42 (2007) 2844–2848.

Please cite this article in press as: R.K. Gupta, N. Birbilis, The influence of nanocrystalline structure and processing route on corrosion of stainless steel: A review, Corros. Sci. (2014), http://dx.doi.org/10.1016/j.corsci.2014.11.041