You are on page 1of 8

Corrosion Science 51 (2009) 23162323

Contents lists available at ScienceDirect

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

Intergranular corrosion susceptibility in supermartensitic stainless steel weldments


J.M. Aquino *, C.A. Della Rovere, S.E. Kuri
So Carlos Federal University (UFSCar), Materials Engineering Department, Rodovia Washington Lus, km 235, CEP 13565-905, So Carlos, SP, Brazil

a r t i c l e

i n f o

a b s t r a c t
The intergranular corrosion susceptibility in supermartensitic stainless steel (SMSS) weldments was investigated by the double loop electrochemical potentiokinetic reactivation (DL-EPR) technique through the degree of sensitization (DOS). The results showed that the DOS decreased from the base metal (BM) to the weld metal (WM). The heat affected zone (HAZ) presented lower levels of DOS, despite of its complex precipitation mechanism along the HAZ length. Chromium carbide precipitate redissolution is likely to occur due to the attained temperature at certain regions of the HAZ during the electron beam welding (EBW). Scanning electron microscopy (SEM) images showed preferential oxidation sites in the BM microstructure. 2009 Elsevier Ltd. All rights reserved.

Article history: Received 31 March 2009 Accepted 8 June 2009 Available online 14 June 2009 Keywords: C. Intergranular corrosion C. Pitting corrosion Stainless steel Welding

1. Introduction Intergranular corrosion is a selective process that occurs in sensitized regions of stainless steels as a result of inadequate heat treatments, welding, or high-temperature service [1]. In the classical sensitization process there is the formation of chromium rich precipitates at grain boundaries, with the subsequent impoverishment of that element in the adjacent matrix. Consequently, passive lms formed over those depleted regions are not stable, thus, it leads to a more susceptible region to a corrosion attack. In that aspect, intergranular corrosion could resemble a galvanic cell in which the grains are the cathodic area, and the corresponding grain boundaries are the anodic one; which results in a high cathodic area in relation to an anodic area [2]. Martensite induced sensitization is another type of sensitized region susceptible to an intragranular corrosion attack [1,3]. This process occurs due to the chromium carbide precipitation in the martensitic lath boundaries [3] which is more intensive in tempered than in quenched martensite [4]. Other metallurgical phases may contribute to inuence the corrosion resistance. According to the literature [4] austenite promotes carbon and nitrogen dissolution, having a consequent reduction in the chromium and molybdenum precipitates. Thus, pit potential demonstrated to be dependent on the austenite content; which gave noble potentials [5]. Otherwise, the d-ferrite phase presented in low carbon 13%Cr steels deteriorates the corrosion resistance [6] due to the higher carbide precipitation around that phase, that is occasioned by its low carbon solubility [7]. Other corrosion and mechanical problems are associated with that phase [8,9].
* Corresponding author. Tel.: +55 16 33518506; fax: +55 16 33615404. E-mail address: dsek@power.ufscar.br (J.M. Aquino). 0010-938X/$ - see front matter 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.corsci.2009.06.009

In welded joints, chromium and molybdenum carbonitride precipitations in the heat affected zone (HAZ), particularly in the supermartensitic stainless steel (SMSS), are responsible for its susceptibility to a corrosion attack. Those precipitations mainly occur at the prior austenite grain boundaries and also at the martensite/ d-ferrite phase boundaries during a multipass welding, with an additional intensication promoted by the post weld heat treatments (PWHT) [10]. Carbide precipitates occur due to the combination of carbon matrix saturation and tempering effect of the subsequent welding passes. That process is not expected to occur in the cap layer or in the single pass welded joints [10]. Intergranular stress corrosion cracking (IGSCC), which was not expected in the HAZ of the SMSS, can occur due to small amounts of Cr-carbide precipitates at the prior austenite grain boundaries [11]. According to the literature [12,13], the HAZ is the most susceptible region to a crack initiation, due to its sensitization in weldments made by conventional welding processes. Additionally, the welding technique exerts inuence on the corrosion behavior of the weldment. The high power density processes are interesting due to their high heat input; which is conned in a small region of the workpiece, and also to their high cooling rates. These characteristics promote the dissolution of Crcarbide precipitates as well as their suppression. That precipitate suppression is caused by very short time periods in the temperature precipitation zone [14]. Corrosion resistance and mechanical properties, when comparing the high power density to the conventional processes [14,15], can lead to better welded joints. Pitting corrosion is also likely to occur at chromium depleted regions [5] as well as near the MnS inclusions [1621]. However, a direct relationship was not established. According to the literature [22], the HAZ of a high power density process weldment is not the most susceptible region to the pitting corrosion due to

J.M. Aquino et al. / Corrosion Science 51 (2009) 23162323

2317

the redissolution process of the Cr-carbide precipitates. Furthermore, the HAZ microstructure does not seem to inuence the pitting corrosion [23]. The double loop electrochemical potentiokinetic reactivation (DL-EPR) is a corrosion testing method characterized by a rapid, quantitative, and non-destructive test to measure the degree of sensitization [24]. This technique was primarily developed for austenitic stainless steel using standard practices. However, intergranular corrosion concerning the martensitic stainless steel (MSS) is a potential problem due to its use in the tempered condition for hardness adjustment. As the conventional MSS contains high levels of carbon, it is not regularly welded. On the other hand, with the development of new weldable supermartensitic stainless steel (SMSS), which has lower carbon content, there is the need to investigate the behavior of those welded joints regarding the sensitization. In order to combine less susceptible welded joints, electron beam welding is inserted, considering intergranular and pitting corrosion. That process links high welding velocities to narrow heat affected zones of a few millimeters, with single passes, and no PWHT. These characteristics enable a more resistant weldment, considering precipitation as well as corrosion. The aim of this work was to measure, comparatively, the degree of sensitization (DOS) of the SMSS weldments, through the DL-EPR technique. 2. Experimental Electrochemical measurements were done in samples of the base metal (BM), the heat affected zone (HAZ), and the weld metal (WM); in two classes of supermartensitic stainless steel that were welded by an electron beam in a chamber of low vacuum. The full penetration butt weld was done between two 20 mm thick plates, which were previously hot-rolled and tempered at 600 C for 10 min. The welding was done by using a matching ller wire metal, in two passes, without PWHT. The chemical composition of the plates, the ller wire, and the welding operation conditions are showed in Tables 1 and 2, respectively. The 1 mm thick samples were extracted by electroerosion, from the inner part of the weldment top. Fig. 1(ac) shows a schematic representation of the samples extraction. Each extraction step was carried after etching the weldment sample with a villelas reagent, through the identication of the characteristic microstructures (BM, HAZ, and WM), as well as through an optical microscopic analysis. In order to attempt disregarding any galvanic inuence, the samples electrochemical investigation was carried separately in the BM and in the WM regions. A clear HAZ sample cutout was not possible to be obtained due to its small size. Fig. 1(d) shows that the sample also presented BM and WM region residues. X-ray diffraction measurements were done separately in the BM, HAZ, and WM samples for identifying and quantifying the metallurgical phases. The austenite phase quantication (volumetric fraction, %) was carried according to the standard practice ASTM-E975 [25]. The X-ray beam was focused in the middle of all samples. That was particularly important when considering the HAZ samples. The analysis was a 2hh with a continuous sweeping rate equal to 2 degrees a minute, ranging from 5 degrees to 120 degrees. During the X-ray measurements the samples were rotationed.
Table 1 Chemical composition (mass%) of steels and ller metal. Material A B Filler C 0.02 0.007 0.012 Si 0.3 0.07 0.45 Mn 0.9 1.6 0.65 P 0.03 0.007 0.005 S 0.004 0.003 0.005 Cr 12.6 11.4 12.3

Table 2 Electron beam welding parameters for the top pass. Parameters Vacuum Welding voltage Working distance Welding speed Root opening Wire feed rate Welding current Heat input Conditions 1 mbar 60 kV 50 mm 7 mm s1 0.5 mm 3.78 m min1 130 mA 1.1 kJ mm1

The electrochemical testing samples were mounted in polyester resin to avoid the presence of crevices after the electric contact. The exposed area was 0.5 cm2. Before the polarization measurements, the samples were wet-grinded on 600 silicon carbide (SiC) paper, washed in distilled water, and immersed in the electrochemical cell. The test solution was 0.5 mol L1 H2SO4 with 0.01 mol L1 KSCN, reagent-grade, in distilled water. The electrolyte was naturally aerated and the temperature was held at 25 C. A conventional electrochemical cell composed of a platinum counter electrode and a satured calomel reference electrode (SCE), connected to a Solartron SI 1287A potentiostat, was used. The DL-EPR polarization curves were obtained after two steps: First, the working electrode was subjected to open circuit conditions, until a steady state potential (Ecorr) was reached. This was accomplished in 25 min. Then, an anodic potentiodynamic sweeping rate of 1.67 mV s1, from 100 mV/Ecorr to +600 mVSCE, was imposed. At +600 mV, the potential scanning was reversed back to 100 mV/Ecorr. At least four curves for each sample region were done to obtain good reproducibility. The test results were expressed in the current densities ratio, iR/ia, which was used to evaluate the DOS. The iR term is the reactivation current density (maximum current density in the reverse scan), and the ia term is the activation current density (maximum current density in the anodic scan). The DOS was compared to the pit potential results, which were done in a separate work [22]. The used weldment samples were the same in both works. The pit potential determination was done in a 3.56 mass% (reagent-grade) sodium chloride solution. 3. Results and discussion Fig. 2(ac) shows some of the initial microstructures of the BM, the HAZ, and the WM. Due to their similar microstructural appearance in the weldment regions of both steels, there were chosen only three micrographs. Tempered martensite is the main metallurgical phase presented in the BM as a result of the tempering heat treatment. The HAZ microstructure also exhibited tempered martensite as a consequence of the thermal heat gradient. That process enabled distinct levels of tempering in the HAZ, as a consequence of the different temperatures reached. Additionally, the HAZ was characterized by a heterogeneous microstructure (relative to grain size). Quenched martensite is the WM characteristic phase that is formed due to the high cooling rates attained. Retained austenite was presented in the BM (A steel = 11.1% and B steel = 39.0%), in the HAZ (A steel = 5.5% and B steel = 4.3%), and

Ni 5.1 6.1 6.4

Mo 1.8 2.6 2.6

Ti 0.01 0.02

V 0.05 0.05

Cu 0.3 0.5

O 0.01 0.01 0.008

N 0.01 0.01 0.01

2318

J.M. Aquino et al. / Corrosion Science 51 (2009) 23162323

Fig. 1. Sample extraction schematic representation of an electron beam weldment top pass: (a) welded joint, (b) weldment slice extracted by electroerosion, (c) testing samples having dimensions of 10 mm 5 mm 1 mm, and (d) the HAZ sample was composed of the BM and the WM region. Weldment regions 1, 2, and 3 refer to the base metal (BM), heat affected zone (HAZ), and weld metal (WM). The asterisk indicates where the electrochemical testings were done.

in the WM (A steel = 26.4% and B steel = 18.5%). According to the literature [7], the austenitic phase in low carbon MSS can only be visualized through transmission electron microscopy (TEM). d-Ferrite was detected in the BM, and also next to the fusion line in the HAZ. Its morphological appearance resembles dark stringers as indicated by the arrows in Fig. 2(a and b). Fig. 3(a and b) shows representative curves of the DL-EPR test for the BM region of the A and B steels. The activation (anodic direction) and reactivation (cathodic direction) current densities are similar, and very high. That indicates a high sensitized material (iR/ia = 0.589 and 0.643 for the A and B steel, respectively). Scanning electron microscopy (SEM) images of Fig. 4(a and b), reveal intergranular and intragranular corrosion attacks, occasioned by the chromium carbide (Cr-carbide) precipitates, and possibly other precipitates like Fe2Mo [26]. That process occurred during the BM tempering heat treatment. Fig. 3 also shows the presence of two anodic peaks in the activation process of both steels. In order to separate and investigate the origin of these peaks, a potentiokinetic sweeping rate at 0.67 mV s1 was only conducted in the anodic direction. Fig. 5 shows that it was not possible to separate, completely, the two anodic peaks of the B steel. According to the literature [3,2730], there are some interpretations for the so-called second anodic current maximum (SACM), which are: (1) nickel enrichment on the surface; (2) adsorbed hydrogen oxidation; (3) Fe2+ ions effect; (4) chromium depleted zone effect, and (5) microstructural and compositional effects. Fig. 6(a and b) shows SEM images obtained after the potential sweeping rate interruption at E1 = 200 mVSCE, and at E2 = 50 mVSCE, respectively, which are indicated by the arrows in

Fig. 2. Optical micrograph of the weldment: (a) B steel BM, (b) A steel HAZ (next to the fusion line), and (c) A steel WM. The arrows indicate the d-ferrite phase.

Fig. 3. DL-EPR curves for the A and B steel BMs. The potential scanning direction is indicated herein (passive vertex potential at 600 mV). The ia and iR current densities are also indicated.

J.M. Aquino et al. / Corrosion Science 51 (2009) 23162323

2319

Fig. 4. SEM micrographs of the BM samples after a DL-EPR testing: (a) A steel, and (b) B steel. The samples were previously polished in alumina 1 lm.

Fig. 6. SEM micrographs of the B steel BM after a DL-EPR testing up to: (a) E1 = 200 mVSCE, and (b) E2 = 50 mVSCE. Microstructure revealed after a single polarization on the anodic direction, for the same testing sample. The sample was previously polished in alumina 1 lm.

Fig. 5. B steel BM anodic polarization at 0.67 mV/s from 100 mVOCP to +600 mVSCE.

Fig. 7. B steel BM cathodic polarization at 0.67 mV/s from +600 mVSCE to 100 mVOCP.

Fig. 5. The micrographs revealed a generalized corrosion attack over the microstructures, without a signicant difference between them. The expected chromium depleted zone effect for the SACM was not conrmed in the SMSS, according to the used experimental techniques. The same occurrence was observed on the A steel, but without any conclusion. During the reactivation process on Fig. 3, the B steel BM also presented two reactivation maximums. The separation and the

investigation of these processes were conducted through a voltage scan only in the cathodic direction, from +600 mVSCE to 250 mV/ Ecorr, after the open circuit conditions were reached. Fig. 7 shows the resulting polarization curve. The arrows at E3 = 100 mVSCE, and at E4 = 190 mVSCE, indicate the potential sweeping rate interruption in order to obtain SEM images. Fig. 8(a) shows that the rst reactivation peak (E3 = 100 mVSCE) corresponds to the martensite

2320

J.M. Aquino et al. / Corrosion Science 51 (2009) 23162323

Fig. 8. SEM micrographs of the B steel BM after a DL-EPR testing up to: (a) E3 = 100 mVSCE, and (b) E4 = 190 mVSCE. Microstructure revealed after a single polarization on the cathodic direction, for the same testing sample. The sample was previously polished in alumina 1 lm.

Fig. 10. SEM micrographs of the HAZ samples next to the BM (2 mm from the fusion line) after a DL-EPR testing: (a) A steel, and (b) B steel. The samples were previously polished in alumina 1 lm.

Fig. 9. DL-EPR curves for the A and B steel HAZs. The potential scanning direction is indicated herein (passive vertex potential at 600 mV).

phase contour laths corrosion attack. The second peak (E4 = 190 mVSCE) is due to the entire martensitic matrix corrosion attack (intragranular), as showed in Fig. 8(b). According to the literature [24], the second reactivation peak could correspond to a nickel rich phase, which is probably related to the retained austenite phase in the SMSS. In addition, the ascribed nickel phase corro-

Fig. 11. SEM micrographs in the middle of the HAZ samples (1 mm from the fusion line) after a DL-EPR testing: (a) A steel, and (b) B steel. The samples were previously polished in alumina 1 lm.

J.M. Aquino et al. / Corrosion Science 51 (2009) 23162323

2321

sion attack at E3 is associated with a chromium content decrease that caused the potential shift toward negative values. Such reactivation occurrence was not observed in the A steel BM sample due to its lower austenite content. The reactivation splitting occurrence of the current density peak occurs on the martensiteaustenite steels containing the following [31]: (i) at least 13% of chromium; (ii) alloyed with molybdenum; (iii) and an austenite content higher than 25% resulting from a suitable tempering heat treatment. That occurrence, on the DL-EPR reactivation curve, was detected in stainless steel type 304 [1]; however, having interpretations different from the ones herein. The DL-EPR curves of the HAZ samples are representatively shown in Fig. 9. The resulted curve is a measurement of the entire HAZ from its fusion line to its 2 mm-away point. The B steel degree of sensitization (iR/ia = 0.005) is much smaller than the A steel one (iR/ia = 0.08). That is because A steel has a higher carbon content than B steel, which results in a higher Cr-carbide precipitation, as it will be showed in the micrographs. However, when comparing the HAZ samples to the BM samples, a much lower DOS is obtained because of the Cr-carbide precipitates possible redissolution at the HAZ specic zones. That process occurred due to the gradients high-temperature in the HAZ, which is proper to the precipitates redissolution [14,32]. Furthermore, a reduced time for the Cr-carbide precipitation could also contribute to the DOS lower levels in that region due to the EBW high cooling rate [14]. Studies regarding the Cr-carbide precipitate redissolution are in progress. The second maximum current density in the activation/reactivation processes were not observed in the HAZ samples, even when 0.67 mV s1 sweeping rates were used.

Fig. 13. DL-EPR curves for the A and B steel WMs. The potential scanning direction is indicated herein (passive vertex potential at 600 mV).

SEM images were obtained in three distinct regions within the HAZ samples of both steels, which were: (1) next to the BM (2 mm from the fusion line), (2) in the middle of it (1 mm from the fusion line), and (3) next to the fusion line. Fig. 10(a and b) shows representative images of the HAZ next to the BM. An outstanding corrosion attack, with a high number of holes, can be observed on the grain boundaries, and also inside the A steel grains. The B steel samples are mainly characterized

Fig. 12. SEM micrographs of the HAZ samples next to the fusion line after a DL-EPR testing: (a) A steel, and (b) B steel. The samples were previously polished in alumina 1 lm.

Fig. 14. SEM micrographs of the WM samples after a DL-EPR testing: (a) A steel, and (b) B steel. The samples were previously polished in alumina 1 lm.

2322

J.M. Aquino et al. / Corrosion Science 51 (2009) 23162323

by an intragranular corrosion attack over the martensitic matrix. This difference is a consequence of the alloying element distinct levels, mainly carbon [33]. Fig. 11(a and b) shows representative images in the middle of the HAZ. Clearly, the intergranular and the intragranular corrosion attacks are decreased in both steels. This is a consequence of the generated thermal heat gradient, which enabled the Cr-carbide precipitate dissolution [14,33]. This process is more prominent in the martensitic matrix because of the higher carbon diffusion coefcient in that phase [7]. Fig. 12(a and b) shows HAZ images next to the fusion line. Within that region, there is a formation of the d-ferrite phase, which is the lighter color phase in the micrographs. An intensive corrosion attack can be observed adjacent to that phase, which was revealed by the DL-EPR tests. This process occurred due to the lower carbon solubility in the d-ferrite phase [7], which resulted in the matrix supersaturation and the consequent precipitation of the Cr-carbides that were adjacent to the d-ferrite phase. As the A steel has a higher carbon content than the B steel, it exhibited a more intensive precipitation, which was conrmed by the numerous holes of the corrosion attack parallel to the d-ferrite phase. The precipitation also occurred in the B steel samples, which were observed through the small cavities in Fig. 11(b), in spite of their very small carbon content. The HAZ micrographs elucidated the difference between the reactivation current densities of the A and B steels (Fig. 9). It was not possible to do the HAZ samples TEM due to their reduced number, and to the difculty of obtaining these samples.

However, new methodologies regarding sample extraction and analysis [34,35] are being considered. The WM region DL-EPR curves are shown in Fig. 13. Both steels presented very similar activation current densities. However, the A steel samples were more sensitized than the B steel ones due to their higher reactivation current density (iR/ia = 0.013 and 0.004 for the A and B steel, respectively). Nevertheless, the WM region was the least sensitized region within the weldment. No evidence of the second maximum current density was observed in the anodic nor in the cathodic scan directions, even after 0.67 mV s1 sweeping rates were applied. Fig. 14(a and b) shows scanning electron WM images after the DL-EPR test. The A steel presented very small corrosion attack cavities, which were almost absent in the B steel samples. There is no evidence of preferential precipitation sites, intragranular or intergranular corrosion attacks in the quenched martensite. The carbide precipitates lower levels in the WM region were observed due to the high cooling rates that occurred during the welding procedure [36]. This process enabled a very short remaining time in the Cr-carbide precipitation region, and a lower alloying element microsegregation which led to a reduced intergranular corrosion susceptibility. The high cooling rate effect also supported the sensitization lower levels in the HAZ due to the reduced time period within which the material stayed in the Cr-carbide precipitation region [14]. The degree of sensitization was compared within different regions of the weldment, and is shown in Fig. 15(a and b). The HAZ samples presented levels of sensitization extremely lower which were comparable to the WM ones. The HAZ performance concern-

Fig. 15. DOS and pitting potential values within the supermartensitic stainless steel weldments: (a) A steel, and (b) B steel.

J.M. Aquino et al. / Corrosion Science 51 (2009) 23162323

2323

ing the intergranular corrosion was completely different from what is reported in the literature [12,13,35]. However, few researches on SMSS, as well as on its weldments that are done through high power density processes, are available. The most important point to consider is the welding process, which has different characteristics when comparing it to the conventional process [14,15]. Among those characteristics, the sharp temperature variation of the HAZ is the most important because it mainly causes the Cr-carbide precipitation/dissolution in it. Fig. 15 also includes the weldment pitting potential values. These results are found in paper [22]. According to the literature [1620], the presence of chromium depleted regions are related to a pit nucleation and growth. As the sensitization increases, the pitting potential decreases, which results in a lower corrosion resistant material. 4. Conclusion The BM was the most sensitized region within the SMSS weldment due to its tempering process, which resulted in its Cr-carbide precipitation. The WM and the HAZ were the lower sensitized regions due to the WM high cooling rates, as well as to the Cr-carbide precipitate redissolution at certain regions of the HAZ. That was conrmed by the SEM images within the HAZ samples. The HAZ of an electron beam top pass weld is susceptible to sensitization. Moreover, the precipitation within that region showed to vary according to the attained temperature during welding. The d-ferrite phase (microstructural effect) also inuenced the HAZ sensitization, because of its low carbon solubility. The austenite content did not seem to inuence the sensitization or the pit potential. The splitted reactivation peak in the A steel BM samples corresponded to two oxidation processes, which were possibly related to: (1) the martensitic phase contour laths, and (2) the martensitic matrix, which may be associated to the austenitic phase content in the SMSS. The second maximum current density in the activation process was not splitted into two peaks. Furthermore, that occurrence did not seem to be related to the chromium depleted effect. A correlation between sensitization and pitting potential was established in the SMSS weldments, indicating that the probability of pitting corrosion enhanced as the sensitization increased. However, it was not conrmed that chromium depleted regions were the preferential sites for pit nucleation and growth. Acknowledgements The authors thank the nancial support provided by CAPES and Dr. Celso Roberto Ribeiro for supplying the steel samples. References
[1] V. Kain, K. Chandra, K.N. Adhe, P.K. De, Detecting classical and martensiteinduced sensitization using the electrochemical potentiokinetic reactivation test, Corrosion 62 (2005) 587593. [2] G.H. Aydogdu, M.K. Aydinol, Determination of susceptibility to intergranular corrosion and electrochemical reactivation behaviour of AISI 316L type stainless steel, Corros. Sci. 48 (2006) 35653583. [3] N. Alonso-Falleiros, M. Magri, I.G.S. Falleiros, Intergranular corrosion in a martensitic stainless steel detected by electrochemical tests, Corrosion 55 (1999) 769778. [4] M. Kimura, Y. Miyata, T. Toyooka, Y. Kitahaba, Effect of retained austenite on corrosion performance for modied 13% Cr steel pipe, Corrosion 57 (2001) 433439. [5] P.D. Bilmes, C.L. Llorente, L. Saire Huamn, L.M. Gassa, C.A. Gervasi, Microstructure and pitting corrosion of 13CrNiMo weld metals, Corros. Sci. 48 (2006) 32613270. [6] T. Hara, H. Asahi, Effect of d-ferrite on sulde stress cracking in a low carbon 13 mass% chromium steel, ISIJ Intern. 40 (2000) 11341141.

[7] E. Folkhard, Welding Metallurgy of Stainless Steel, rst ed., Springer-Verlag, New York, 1988. [8] K. Kondo, K. Ogawa, H. Amaya, H. Hirata, M. Ueda, H. Takabe, Y. Miyazaki, Alloy design of super 13Cr martensitic stainless steel (Development of super 13Cr martensitic stainless steel for line pipe 1), in: Paper of the Supermartensitic Stainless Steels, Brussels, Belgium, 1999. [9] Y.C. Lin, S.C. Chen, Effect of residual stress on thermal fatigue in a type 420 martensitic stainless steel weldment, J. Mater. Process. Technol. 138 (2003) 2227. [10] E. Ladanova, J.K. Solberg, T. Rogne, Carbide precipitation in haz of multipass welds in titanium containing and titanium free supermartensitic stainless steels part 1 proposed precipitation mechanisms, Corros. Eng. Sci. Technol. 41 (2006) 143151. [11] H. Nakamichi, K. Sato, Y. Miyata, M. Kimura, K. Masamura, Quantitative analysis of Cr-depleted zone morphology in low carbon martensitic stainless steel using FE-(S)TEM, Corros. Sci. 50 (2008) 309315. [12] L. Coudreuse, V. Ligier, C. Lojewski, P. Toussaint, Environmental induced cracking (SSC and SCC) in supermartensitic stainless steels (SMSS), in: Paper of the Supermartensitic Stainless Steels, Brussels, Belgium, 2002. [13] T. Rogne, M. Svenning, Intergranular corrosion of supermartensitic stainless steel a high temperature mechanism? in: Paper of the Supermartensitic Stainless Steels, Brussels, Belgium, 2002. [14] H.T. Lee, J.L. Wu, The effects of peak temperature and cooling rate on the susceptibility to intergranular corrosion of alloy 690 by laser beam and gas tungsten arc welding, Corros. Sci. 51 (2009) 439445. [15] H.T. Lee, J.L. Wu, Correlation between corrosion resistance properties and thermal cycles experienced by gas tungsten arc welding and laser beam welding alloy 690 butt weldments, Corros. Sci. 51 (2009) 733743. [16] G.T. Burstein, S.P. Vines, Repetitive nucleation of corrosion pits on stainless steel and the effects of surface roughness, J. Electrochem. Soc. 148 (2001) B504B516. [17] G.S. Frankel, Pitting corrosion of metals, J. Electrochem. Soc. 145 (1998) 2186 2198. [18] C.O.A. Olsson, D. Landolt, Passive lms on stainless steels: chemistry, structure and growth, Electrochim. Acta 48 (2003) 10931104. [19] I. Reynaud-Laporte, M. Vayer, J.P. Kauffmann, R. Erre, An electrochemical-AFM study of the initiation of the pitting corrosion of a martensitic stainless steel, Microsc. Microanal. Microstruct. 8 (1997) 175185. [20] M.P. Ryan, D.E. Williams, R.J. Chater, B.M. Hutton, D.S. McPhail, Why stainless steel corrodes, Nature 415 (2002) 770774. [21] 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) 17551764. [22] J.M. Aquino, C.A. Della Rovere, S.E. Kuri, Localized corrosion susceptibility of supermartensitic stainless steel in welded joints, Corrosion 64 (2008) 3539. [23] J. Enerhaug, O. Grong, U.M. Steinsmo, Factors affecting initiation of pitting corrosion in super martensitic steels weldments, Sci. Technol. Weld. Join. 6 (2001) 330338. [24] V. Chal, R. Stefec, On the development of the electrochemical potentiokinetic method, Electrochim. Acta 46 (2001) 38673877. [25] ASTM standard E 975-95, Standard practice for X-ray determination of retained austenite in steel with near random crystallographic orientation, ASTM, PA, 1995, pp. 16. [26] V. Vodarek, M. Tvrdy, A. Korgak, Heat treatment of supermartensitic steels, Inz. Mater. 5 (2001) 939941. [27] L. Felloni, S.S. Traverso, G.L. Zucchini, G.P. Cammarota, Investigation on the second anodic current maximum on the polarization curves of commercial stainless steels in sulphuric acid, Corros. Sci. 13 (1973) 773779. [28] A.A. Hermas, M.S. Morad, K. Ogura, A correlation between phosphorous impurity in stainless steel and a second anodic current maximum in H2SO4, Corros. Sci. 41 (1999) 22512266. [29] O.L. Riggs Jr., The second anodic current maximum for type 430 stainless steel in 0.1N H2SO4, Corrosion 31 (1975) 413415. [30] M.B. Rockel, Interpretation of the second anodic current maximum on polarization curves of sensitized chromium steels in 1N H2SO4, Corrosion 27 (1971) 95103. [31] V. Chal, M. Blahetov, J.Hubckov, Z. Krhutov, S. Lasek, K. Mazanec, Corrosion and structural testing of martensitic steels by electrochemical polarization method, in: Paper of the Supermartensitic Stainless Steels, Brussels, Belgium, 2002. [32] P.H.S. Cardoso, C. Kwietniewski, J.P. Porto, A. Reguly, T.R. Strohaecker, The inuence of delta ferrite in the AISI 416 stainless steel hot workability, Mater. Sci. Eng. A 351 (2003) 18. [33] O.M. Akselsen, G. Rorvik, P.E. Kvaale, C. van der Eijk, Microstructure-property relationships in HAZ of new 13% Cr martensitic stainless steels, Weld. J. 83 (2004) 160167. [34] H. Yanliang, B. Kinsella, T. Becker, Sensitisation identication of stainless steel to intergranular stress corrosion cracking by atomic force microscopy, Mater. Lett. 62 (2008) 18631866. [35] C. Garcia, M.P. de Tiedra, Y. Blanco, O. Martin, F. Martin, Intergranular corrosion of welded joints of austenitic stainless steels studied by using an electrochemical minicell, Corros. Sci. 50 (2008) 23902397. [36] S.A. David, S.S. Babu, J.M. Vitek, Welding: solidication and microstructure, J. Miner. Met. Mater. Soc. 55 (2003) 1420.

You might also like