You are on page 1of 14

Proceedings of the 12th International Conference on Environmental Degradation of Materials in Nuclear Power System Water Reactors Edited by T.R.

R. Allen, P.J. King, and L. Nelson TMS (The Minerals, Metals & Materials Society), 2005

FRACTURE SURFACE MORPHOLOGY OF STRESS CORROSION CRACKS IN NICKEL-BASE WELDS W. J. Mills Bechtel Bettis, Inc. P.O. Box 79, West Mifflin, PA 15125-0079 Keywords: Ni-based welds, Alloy 82H, Stress Corrosion Cracking, Intergranular Cracking, Fractography Abstract The fracture surface morphology of stress corrosion cracks (SCC) in Alloy 82H welds was characterized by metallographic and fractographic techniques. The compact tension specimens evaluated in this program were tested in 316C to 360C hydrogenated water and the macroscopic SCC behavior was detailed in Ref. [1]. Tests were conducted under constant-load conditions and with periodic unload-reload cycles every 10 or 100 min. Specimens tested under constant load and with 100-min. unload-reload cycles exhibit similar fracture surface morphologies; cracking was predominantly intergranular and very uneven. Uneven cracking is caused primarily by the heterogeneous nature of welds, as the most susceptible grain boundaries readily separate producing SCC fingers. The most resistant boundaries do not separate thereby causing unbroken ligaments in the wake of an advancing crack front. A 10-min. unload-reload cycle produces similar amounts of intergranular and transgranular cracking and a more uniform crack advance, although intergranular fingers jut out beyond the overall crack front. Even though 10-min. unloadreload cycles cause only a small to modest acceleration in crack growth rates, they should not be used to estimate SCC performance because the cracking mechanisms are different. The fatigue precrack/SCC interfaces were characterized to understand the crack incubation process whereby a transgranular precrack transitions into an intergranular SCC. Introduction Stress corrosion cracking (SCC) of Alloy 82 and 182 welds has been observed in pressurized water reactor (PWR) outlet nozzles and head penetrations and in some cases has caused through-wall Cracks propagate along cracks that produced leakage.[2-6] columnar grain boundaries and exhibited significant branching. To support disposition of stress corrosion cracks, extensive testing of Alloy 82, 82H and 182 welds[1,3,7-9] has been performed whereby crack growth rates (CGRs) were characterized as functions of temperature, stress intensity factor (K), weld orientation, and hydrogen concentration in the coolant. In this study, metallographic and scanning electron microscope (SEM) fractographic examinations of Alloy 82H weld specimens were performed to understand operative SCC mechanisms. Specimens were SCC tested in 316C to 360C hydrogenated water under constant-load conditions or with periodic unloadreload cycles.[1] Emphasis was placed on characterizing the fatigue precrack/SCC interface to understand the process by which transgranular precracks transition into intergranular cracks. Experimental Procedure SCC tests were conducted on Alloy 82H welds fabricated using a gas-tungsten-arc (GTA) process. Specific details regarding the welds and test methods are provided in Ref. [1]. The welds were fabricated using standard welding procedures, so the results are expected to apply in general to Alloy 82 and 82H welds. Two 669 crack orientations were studied: the T-L orientation where cracking occurs parallel to the welding direction and T-S orientation where cracking occurs in the weld root-to-crown direction. Both orientations share a common plane, but the crack growth directions differ relative to the dendrite and columnar grain structures. Compact tension specimens were SCC tested under constant-load conditions or with periodic unload-reload cycles every 10 or 100 min. Tests with periodic unload-reload cycles used a trapezoidal waveform with a 10 or 100 min. hold period, 0.5 min. rise time, 0.5 min. fall time and a stress ratio (R) of 0.65. Tests were performed in flowing autoclaves in water with a room temperature pH of 10.1 to 10.3, an oxygen concentration less than 14 ppb, and a conductivity of ~50 S/cm. For tests conducted in 360C water the hydrogen concentration was 150 cc H2/kg H2O while all 316C and 338C tests were conducted in 40 to 60 cc H2/kg H2O. The weld microstructure was characterized using standard metallographic procedures. Light optical microscopy was performed on metallographic specimens electrolytically etched in 5% nital or a nitric-acetic acid solution to reveal the general microstructure. SEM examination of samples etched in brominemethanol solution was performed to provide a more detailed characterization of grain boundaries. Broken specimen halves were examined on an SEM to characterize operating cracking mechanisms as a function of test conditions. Results and Discussion Crack Growth Rates The SCC behavior for the Alloy 82H welds tested in Ref. [1] are summarized in Figures 1 and 2, which plot CGRs corrected for the percent engagement (i.e., measured CGR are divided by the fraction of engagement) as a function of applied K level. Crack growth rates are also adjusted for temperature based on an activation energy of 31 kcal/mol. Inspection of the SCC data in Figures 1 and 2 reveals the following: The range in CGR scatter for a given weld (within weld scatter) at a given temperature and KI level is between 2X and 5X. The range in CGR scatter for all welds from multiple heats at a particular temperature and KI level is about 10X. CGRs in the T-S direction are slightly higher than those in the T-L direction, but there is considerable overlap in the data. Periodic unload-reload cycles every 100 min. have little effect on SCC at 338C, as CGRs are slightly above average constantload rates but well within the range of constant-load rates. Periodic unload-reload cycles every 10 min. have a small to modest effect on CGRs at 338C. Specifically, a 10-min. cycle increases CGRs by a factor of 1.5X to 4X. This effect is expected to become more pronounced at lower temperatures as the ratio of SCC rates to corrosion fatigue rates decreases. Periodic unload-reload cycles typically reduce the range in CGR scatter for individual welds to less than 2X.

Comparison of data ranges in Figures 1 and 2 shows that decreasing the dissolved hydrogen content from 150 to 50 cc H2/kg H2O causes a three-fold increase in CGRs. This behavior is consistent with dissolved hydrogen effects in Ref. [10] where CGRs were shown to decrease with increasing distance from the Ni/NiO transition, which occurs at ~15 cc H2/kg H2O at 338C. Based on data in Ref. [10], decreasing the hydrogen content from 150 to 50 cc H2/kg H2O would be expected to increase CGRs by a factor of 3, consistent with Figures 1 and 2. Nature of Uneven Cracking Macroscopic views of fracture surfaces for T-S and T-L oriented SCC specimens are shown in Figure 3. In the T-S orientation where the crack growth direction is parallel to the columnar grains, SCC is very usually uneven as cracks proceed along susceptible grain boundaries leaving large unbroken ligaments in regions that resist intergranular cracking. Particularly large ligaments are common and in some cases extend back to the precrack. In addition, many small and intermediate size ligaments are completely circumnavigated by stress corrosion cracks. The island ligaments, which appear as shiny specks, exist throughout the entire SCC region, even immediately adjacent to the precrack. Fracture surfaces of T-L oriented specimens exhibit a duplex appearance with some specimens exhibiting uniform crack extension (Figure 3e) while others have uneven cracking (Figure 3f). In general, SCC in T-L oriented specimens from weld C is more uniform than that from weld A, as SCC in the latter is consistently retarded on the side toward the weld crown. The uneven cracking does not appear to be caused by residual stresses because fatigue precracks are even. Hence, it appears to be associated with microstructural effects due to variations in timetemperature history during multiple pass welding, although the specific microstructural features have not been identified. Regardless of whether cracking is uniform or not, many unbroken shiny ligaments exist throughout the SCC region and along the precrack/SCC interface. Figure 4 plots the ratio of maximum to average crack extension ( aMAX/ aAVE) which provides a quantitative assessment of the size of SCC fingers. For aAVE values between 0.5 and 1.5 mm, aMAX values are 2 to 4 times greater than aAVE. At very low aAVE values, ratios typically range from 4 to 7, but ratios as high as 11 were observed. The large aMAX/ aAVE values for shallow cracks reflects the SCC fingers that jut out beyond the precrack during the early stages of cracking. With increasing crack extension, aMAX/ aAVE values decrease as SCC fingers coalesce. At aAVE values between 3 and 7 mm, the aMAX to aAVE ratio is 1.3 0.2. Even at large crack extension, the ratio is well above unity because crack fronts remain uneven. The data in Figure 4 were regressed to establish a relationship between the maximumto-average crack extension ratio and average crack extension: a MAX 1.745 1.026 (1) a AVE 0.380 a AVE where aMAX and aAVE are in mm. An alternate approach for displaying the uneven nature of SCC is to plot aMAX versus aAVE, as shown in Figure 5. The extent of unevenness is represented by the difference between maximum and average crack extensions, which corresponds to the vertical distance between each data point and the aMAX = aAVE line. When stress corrosion cracks start to grow (i.e., aAVE < 1 mm), the extent of unevenness increases rapidly as SCC fingers jut out well beyond the overall crack front. As cracking continues (i.e., 670

aAVE > 1.5 mm), the degree of unevenness remains nearly constant. This behavior indicates that SCC fingers continue to form in newly exposed regions with susceptible grain boundaries, while newly exposed regions with less susceptible grain boundaries resist cracking. Hence, uneven crack fronts continue to persist after stress corrosion cracks become fully engaged. Figures 4 and 5 show that constant load and unload-reload cycling produce similar aMAX to aAVE ratios. This observation indicates that periodic unloadings, especially with a 100-min. holdtime at 316C and 338C, have little effect on the propensity for SCC fingers to form in regions with susceptible grain boundaries. Even fracture surfaces generated in 288C to 338C water with 10-min. unload-reload cycles exhibit uneven cracking as SCC fingers extend beyond the advancing crack front. Microstructure The weld microstructure was characterized to correlate operative SCC mechanisms with microstructural features. Along the weld centerline, where specimen notches are located, the microstructure consists of coarse columnar grains that are aligned in the root-tocrown direction. Grain boundaries are high-angle boundaries that separate colonies of dendrites with different orientations. They have a convoluted shape, as shown in Figure 6, caused by the interpenetrating nature of dendrites. The degree of mismatch for these boundaries is controlled by the relative misorientation between adjacent dendrite bundles; grain boundaries with a large mismatch are high-energy, high-angle grain boundaries. The dark particles in the interdendritic regions are titanium carbonitride inclusions. Inclusions are also observed along grain boundaries where they hinder migration from the interdendritic regions. Submicron MgS inclusions are observed in interdendritic regions and along grain boundaries. Small Al,Mg,Si-rich oxide particles are often observed inside a Ti(C,N) shell, demonstrating that oxides serve as nucleation sites for primary inclusions during weld solidification. Figure 6b reveals that grain boundaries contain a high density of fine particles and a few micron-size Ti(C,N) inclusions. Analytical electron microscopy in Ref. [11] showed that fine niobium-rich MC-type carbonitrides decorate all grain boundaries and chromium-rich M23C6 carbides decorate some boundaries. Fine Nb,Ti(C,N) particles also precipitate in the vicinity of Ti(C,N) inclusions within the interdendritic regions. Metallographic profiles of stress corrosion cracks generated under constant-load conditions, which are shown in Figure 7, reveal that cracking is predominantly intergranular in both the T-L and T-S orientations. Cracks propagate along high-angle grain boundaries with preferential cracking occurring along boundaries that separate dendrite colonies with markedly different orientations. The different dendritic grain orientations along crack paths are apparent in the interference contrast image in Figure 7b. The other key observation is the discontinuous nature of intergranular cracks, whereby cracks arrest when they encounter grain boundaries or grain boundary segments that resist cracking. Variations in cracking resistance may be associated with differences in grain boundary structure (e.g., differences in mismatch, precipitation or segregation) or the relative orientation of grain boundaries to the applied stress. Intergranular cracks circumnavigate resistant areas leaving unbroken ligaments behind the advancing crack front. In some cases, transgranular cracks initiate within these ligaments (Figure 7d) due to high shear stresses that develop when ligaments are surrounded by intergranular cracks with different elevations. Separation along planar slip bands gives rise to the faceted appearance of islands

with transgranular cracks, as discussed in the next section. The planar slip bands that develop within ligaments are also responsible for the planar slip offsets superimposed on nearby intergranular faces. Fracture Surface Appearance of SCC Under Constant Load T-S Oriented Cracks. Fractographs in Figure 8 show that intergranular cracking is dominant in T-S oriented specimens tested under constant-load conditions. The convoluted nature of grain boundaries gives rise to the wavy or undulating morphology of intergranular faces. The long axis of dendritic grains lies within 45 of the root-to-crown direction (which is parallel to the cracking direction for T-S oriented specimens) with most grains being offset by only 10 to 20. Within the vast fields of intergranular cracking, there are islands of transgranular cracking or unbroken ligaments that failed during post-test fatigue apart. The unbroken ligaments, which form in regions with enhanced SCC resistance, range in width from 20 m (a single grain diameter) to 2 mm and can be found over the entire SCC region including the precrack/SCC interface. Because these ligaments are intact after SCC testing, their surfaces are not oxidized in test or during post-test heat tinting so they have a shiny appearance on a stereo light microscope. During SCC testing, some ligaments eventually fail due to a transgranular cracking mechanism along planar slip bands, as discussed in the previous section. Environmental cracking along slip bands induces a transgranular faceting mechanism that causes some ligaments to fail. The transgranular islands, which are often limited in size to a single grain or a few grains, exhibit ill-defined facets (Figure 8b). Secondary intergranular cracks, or off-angle cracks, often form in the vicinity of unbroken ligaments or transgranular islands as the intergranular cracks try to bypass regions with enhanced SCC resistance (Figure 7b). The cracking resistance of ligaments is then further enhanced because of the off-angle cracking (i.e., deflection of the crack tip region so it is not normal to the loading axis), loss of constraint and blunting of the crack tip. When shear stresses are sufficiently high to induce slip band cracking, a transgranular faceting mechanism leads to local separation. Other ligaments resist separation because of the loss of constraint, crack deflection and crack tip blunting. Precrack/SCC interfaces (Figure 9) were examined to understand the SCC incubation process by which transgranular fatigue precracks transition into intergranular cracks. Along much of the interface, it appears that intergranular cracks initiate directly from the tip of the precrack (Figure 9a), but stereo fractography shows that in most instances there is a steep, almost vertical, transgranular wall separating the two regions. Periodically, an intergranular crack crosses the plane of a precrack, such that a transgranular wall forms above the precrack on one side and below the precrack on the other side. Another key feature is off-angle intergranular cracks (i.e., nearly normal to the overall crack plane) that extend into both the SCC and precrack regions, as shown in Figure 9c and 9d. In other cases, off-angle intergranular cracks initiate at the precrack tip and extend only into the SCC region (center of Figures 9e and 9f). Off-angle cracks are important to the crack incubation process as discussed in the next paragraph. Finally, reverse intergranular fingers that extend into the precrack (Figures 9b, 9d and 9e) are also found along the precrack/SCC interface. The importance of each of these features in the SCC incubation process is discussed below. 671

Obviously, the first high-angle grain boundaries to separate are the susceptible boundaries that intersect a fatigue precrack. This includes grain boundaries or grain boundary segments that coincide with the tip of a precrack and off-angle boundaries that intersect a precrack plane at various angles. Cracks along offangle grain boundaries, which are driven by stresses in the z or through-thickness direction, can initiate at or just behind the tip of a fatigue precrack, as high plastic strains develop in these regions. The primary and off-angle intergranular cracks emanating from a precrack quickly propagate into the matrix, thereby creating a path for water to reach other high-angle grain boundaries. The offangle intergranular cracks transition into primary intergranular cracks1 oriented normal to the loading axis and then begin to grow axially and laterally. As these cracks spread out laterally, the intergranular crack and transgranular precrack are in close proximity but not in contact. The high shear stresses that develop between the two types of cracks initiate a transgranular faceting mechanism, similar to the process by which unbroken ligaments separate. Accordingly, the precrack and stress corrosion crack are linked up by a steep wall with a faceted appearance. In other regions, transgranular faceting nucleates at a precrack and then transitions into an intergranular crack when it encounters a susceptible grain boundary. This behavior is believed to result in transgranular walls that are less steep. Reverse SCC fingers that extend into a precrack (Figures 9b, 9d and 9e) are a manifestation of off-angle intergranular cracks and appears to be unique to welds. Reverse SCC fingers have a peninsula shape and form when SCC initiates along multiple offangle grain boundaries that intersect a precrack just behind its tip. The resulting intergranular cracks coalesce into a primary intergranular crack that extends into the precrack region. On one fracture half, a reverse finger is a depression below the precrack, while it protrudes above the precrack on the mating half. In many cases, the off-angle cracks that create reverse SCC fingers extend into the SCC region (Figures 9b and 9e). While reverse fingering is not a critical SCC incubation process, it points out the importance of the off-angle intergranular cracks in initiating SCC as well as the low stresses needed to incubate SCC along susceptible grain boundaries. The overall incubation process for Alloy 82H welds is almost identical to that for Alloy X-750. Figure 10 shows typical precrack/SCC interfaces for HTH Alloy X-750 specimens that were tested in 360C water with 50 cc H2/kg H2O under constantload conditions. The tests were interrupted and specimens fatigued apart just after SCC incubation so the incubation process could be interrogated. Examination of precrack/SCC interfaces reveals that SCC incubation occurs in local regions where susceptible grain boundary segments are aligned with a precrack or where off-angle boundaries intersect a precrack. Like the behavior in EN82H weld metal, off-angle grain boundaries in Alloy X-750 initiate cracking at and behind the tip of a precrack. The resulting off-angle cracks extend a few grain diameters ahead of a precrack where they transition into primary cracks that spread out laterally. Hence, off-angle cracks provide a conduit for water to reach susceptible grain boundaries beyond the precrack. In other regions, the precrack and stress corrosion crack are separated

Off-angle intergranular cracks near the precrack/SCC interface nucleate during the early stages of cracking and provide intergranular sites from which primary intergranular cracks nucleate. Although off-angle cracks are the first to form, they appear to be secondary cracks steeply inclined to the primary fracture surface when specimens are broken apart.

by a steep wall with transgranular facets. Transgranular cracking along planar slip bands causes environmental crack extension until a susceptible grain boundary is encountered at which point intergranular SCC commences. In other regions, it appears that the intergranular cracks extend laterally and then link up with the precrack by a steep wall of transgranular facets. In summary, the SCC incubation processes operative in HTH Alloy X-750 are very similar to those in EN82H welds. SCC incubation first occurs at local sites where susceptible grain boundaries intersect a precrack or where a transgranular faceting mechanism causes a transgranular crack to eventually reach a susceptible grain boundary. Rapid crack advance creates fingers of SCC ahead of a precrack, which then begin to grow laterally. In the base metal, the SCC fingers readily coalesce so crack extension is more uniform than in the weld metal. The base metal also develops unbroken ligaments as SCC resistant regions are encountered, but the ligaments are smaller and sparser than those in the weld metal. T-L Oriented Cracks. Fractographs in Figure 11 show the fracture surface appearance of T-L oriented specimens tested under constant-load conditions in 338C and 360C water. The overall fracture surface appearance for T-L oriented SCC is almost identical to its T-S counterpart. Vast portions of the fracture surface are covered with intergranular cracks with a wavy or convoluted morphology. The only notable difference in intergranular fracture surfaces for the two orientations is that the long axis of the dendritic grains is typically oriented 70 to 110 to the crack growth direction for the T-L orientation (Figure 11a). Islands of unbroken ligaments or transgranular facets are often found inside intergranular SCC regions. Unbroken ligaments formed when stress corrosion cracks bypass regions where grain boundaries resist cracking. Many of these ligaments did not fail during SCC testing so they have a shiny appearance as their surfaces are not oxidized. Other ligaments fail by a transgranular faceting mechanism (Figure 11b) due to local separation along intense slip bands, similar to the cracking mechanism described earlier for T-S oriented specimens. It is noteworthy that many T-L oriented specimens from weld C exhibited a high density of unbroken ligaments immediately beyond the precrack, while the overall crack extension region was rather uniform. The presence of many ligaments at the precrack/SCC interface demonstrates that SCC incubates in isolated regions where susceptible grain boundaries intersect the precrack region. In the T-L orientation, however, unbroken ligaments are readily bypassed, in contrast with T-S oriented specimens where massive ligaments tend to form. The different responses are believed to be associated with the orientation of susceptible columnar grains relative to the cracking direction. When an advancing stress corrosion crack encounters a susceptible columnar grain boundary in a T-S oriented specimen, local cracking rapidly occurs parallel to the cracking direction, which exacerbates ligament formation as SCC fingers extend beyond the advancing crack front. By contrast, when the same situation develops in T-L oriented specimens, local crack extension rapidly occurs normal to the macroscopic cracking direction. As a result, cracking along susceptible columnar grain boundaries penetrate into the ligaments and eventually choke them off. Examination of precrack/SCC interfaces in T-L oriented welds (Figures 11c and 11d) reveals three SCC incubation mechanisms. The first two mechanisms involve SCC initiation directly from fatigue precracks when grain boundary segments coincide with the tip of a precrack or when off-angle grain boundaries intersect a precrack region just behind its tip. The third mechanism involves 672

initiation and propagation of an environmentally induced transgranular crack that eventually encounters a susceptible grain boundary where it transitions into an intergranular crack (Figure 11c). These crack incubation mechanisms are similar to those operative in the T-S orientation, although there are important differences. In both orientations, local incubation occurs readily in regions where susceptible grain boundary segments lie along the crack tip. But because fewer grain boundaries coincide with the tip of a T-L oriented precrack (as the long axis of columnar grains is parallel to the crack tip), transgranular cracking is a more important mechanism than it is for T-S oriented specimens. Figure 11c shows evidence of faceted transgranular cracking that eventually encounters a grain boundary where it transitions into an intergranular crack. Crack incubation along off-angle grain boundaries that intersect a precrack just behind its tip also occurs in T-L oriented specimens (Figures 11d). In this orientation, however, off-angle cracks become an integral part of the primary fracture surface, because long grain boundary segments tend to be parallel to the crack tip, as opposed to the T-S orientation where they are perpendicular to the crack tip. As an off-angle crack starts to propagate, it readily transitions into an intergranular crack that is parallel to the precrack plane, so the off-angle crack is incorporated into the primary fracture surface. By contrast, most off-angle cracks that incubate SCC in T-S specimens are not incorporated into the primary fracture surface; rather they appear as secondary cracks. This behavior explains why few secondary intergranular cracks extend into T-L oriented precracks, in direct contrast with the T-S orientation where secondary cracks often intersect the precrack. SCC Generated with 100-Minute Unload-Reload Cycles Representative fractographs of T-S oriented fracture surfaces generated in 338C and 316C water with an unload-reload cycle every 100 min. are shown in Figure 12. The fracture surface morphology away from the precrack/SCC interface is almost identical to that generated under constant-load conditions. The dominant cracking mechanism is intergranular fracture where grain boundary faces have a wavy or convoluted nature (Figures 12a and 12b). Columnar grains are seen to be oriented within 45 of the root-to-crown direction (Figure 12b) with most grains being within 20 (Figure 12a). Islands of unbroken ligaments and transgranular facets (Figure 12c) are found inside vast fields of intergranular fracture. The faceted regions are inclined as cracking along planar slip bands links up intergranular crack segments at different elevations. Secondary intergranular cracks are often located in the vicinity of unbroken ligaments and transgranular islands as intergranular cracks attempt to propagate around SCC resistant regions (Figure 12c). Within the faceted regions, the local cracking direction can be determined because the ledges separating individual facets are parallel to the crack growth direction. As shown in Figure 12c, facet ledges are normal to the root-to-crown direction, thereby demonstrating the local cracking direction is perpendicular to the macroscopic crack growth direction. This behavior occurs because intergranular SCC fingers extend rapidly beyond the advancing crack front and bypass regions with SCC-resistant grain boundaries. A faceted growth mechanism then initiates along the intergranular crack/ligament interface and propagates into the uncracked ligament. Figures 12d through 12f, which show the precrack/SCC interface of specimens subjected to 100-min. unload-reload cycles, reveal that the SCC incubation mechanisms are similar to those for

constant-load specimens, with the exception of a corrosion fatigue mechanism that is operative just beyond the precrack. Consistent with the mechanisms in constant-load specimens, intergranular crack incubation in specimens with unload-reload cycles occurs where grain boundaries either intersect a precrack or coincide with the tip of a precrack. Accordingly, off-angle cracks extend into the precrack and SCC regions (Figure 12d), which create a path for water to reach other susceptible boundaries. Once intergranular SCC incubates along a precrack, it propagates rapidly into the matrix forming SCC fingers that jut out ahead of a precrack and then begin to extend laterally (Figures 12e and 12f). These aspects of intergranular crack incubation are identical to their constantload counterparts. The fundamental difference between specimens subjected to constant load versus 100-min. unload-reload cycles is the role of corrosion fatigue immediately ahead of a precrack. Unload-reload cycles cause corrosion fatigue crack extension in the regions between SCC fingers. The corrosion fatigue regions take on a transgranular faceted morphology. At 338C, facets are rather ill-defined (Figure 12d) whereas facets formed at 316C have a more crystallographic appearance (Figure 12e). When corrosion fatigue cracks intersect susceptible grain boundaries, they transition into intergranular stress corrosion cracks, as shown in Figure 12d. Thereafter, cracking remains predominantly intergranular with islands of unbroken ligaments and isolated patches of transgranular faceting, consistent with the SCC behavior under constant-load conditions. These findings demonstrate that periodic unload-reload cycles every 100 min. affect the crack incubation process, but do not affect the crack propagation mechanism. It is not surprising, therefore, that 100min. unload-reload cycles have little effect on overall crack growth rates at and above 316C. However, data scatter is diminished with periodic unload-reload cycles, because cycling promotes crack incubation in regions that resist intergranular cracking. SCC Generated with 10-Minute Unload-Reload Cycles SEM fractographs of welds C4 and A1 tested in 338C water with unload-reload cycles every 10 min. are shown in Figure 13. Both welds exhibit a mixed cracking mode. Weld C4 with a T-S orientation exhibits nearly equal amounts of intergranular and transgranular cracking, whereas weld A1 with a T-L orientation exhibits slightly more intergranular cracking. The intergranular cracks have a convoluted morphology that is identical to the stress corrosion cracks observed under constant-load conditions. However, large regions with transgranular facets are also observed. Transgranular facets have a variety of appearances, including fine crystallographic facets, ill-defined facets and coarse facets with river patterns. Metallographic profiles of cracks generated in 338C water with unload-reload cycles every 10 min. reveal a mixed cracking mode with evidence of both transgranular cracking and intergranular cracking along high-angle grain boundaries (Figure 14). Transgranular cracks associated with a corrosion fatigue mechanism follow persistent slip bands. Cracks that follow a single slip band (Figure 14a) exhibit broad facets, while those that propagate along a series of parallel slip bands (Figure 14b) produce fine crystallographic facets. Examination of the precrack/SCC interface in Figure 15 reveals extensive transgranular faceting due to corrosion fatigue. Transgranular cracks transition into intergranular cracks when susceptible grain boundaries are encountered, but it is common for the cracking mode to switch back and forth as the crack propagates. This mixed cracking mode is in direct contrast with 673

the dominant intergranular cracking mode operative away from the precrack/ SCC interface in welds subjected to 100 min. cycles. In a few regions, primary intergranular cracks and off-angle intergranular cracks incubate along a precrack during 10-min. cycle tests, as shown in Figure 15. Based on the facet orientation in adjacent regions, the intergranular cracks quickly jut out beyond the precrack long before corrosion fatigue cracks develop. Specifically, the faceted growth direction is normal to the intergranular/ligament interface (i.e., normal to the macroscopic cracking direction), thereby demonstrating that intergranular cracks form quickly and serve as nucleation sites from which the transgranular facets emanate. Substantial amounts of transgranular cracking indicate that cycling every 10 min. introduces a corrosion fatigue component that accelerates cracking, as CGRs are 1.5 to 4 times greater than their constant-load counterparts. However, CGRs for these two loading conditions are not radically different because significant intergranular SCC occurs in cyclic tests. For weld A1, CGRs obtained with a 10-min. hold time, where intergranular cracking covered over half of the fracture surface, are only ~30% greater than rates obtained during constant load tests. Nevertheless, a 10min. unload-reload cycle cannot be used to infer SCC rates because it introduces a significant transgranular fatigue component. Conclusions The fracture surface morphology for high temperature stress corrosion cracks in EN82H welds was characterized to understand operative SCC mechanisms under constant-load and with periodic unload-reload cycles. The uneven nature of SCC and the process by which transgranular precracks transition into intergranular SCC were also evaluated. Key findings are presented below: Constant-Load SCC Under constant-load conditions, the dominant SCC mechanism for both T-S and T-L oriented welds is intergranular cracking as cracks propagate along high-angle grain boundaries. The intergranular faces have a wavy appearance that reflects the convoluted nature of grain boundaries in as-cast structures. Within the intergranular regions, there are isolated islands that either failed by a transgranular faceting mechanism or resisted separation, resulting in unbroken ligaments. In the T-S orientation, SCC fingers jut out ahead of a precrack, as intergranular cracking occurs along the most susceptible columnar grain boundaries, which are oriented parallel to the cracking direction. The intergranular fingers spread out laterally and axially, but tend to bypass regions with enhanced SCC resistance, thereby producing small and massive unbroken ligaments behind the advancing crack front. Intergranular SCC incubation in T-S oriented welds involves three basic mechanisms: a direct transition to intergranular cracking when grain boundary segments are aligned with the tip of a precrack, separation of off-angle grain boundaries that produce secondary intergranular cracks (driven by throughthickness stresses) that eventually transition into a primary intergranular crack, and environmentally induced transgranular cracking that transitions into an intergranular crack when susceptible grain boundaries are encountered. The off-angle intergranular cracks that initiate at or behind the tip of a precrack provide a conduit for water to reach other high-angle boundaries that are normal to the applied stress direction.

The T-L oriented welds tested under constant load exhibit intergranular cracking with islands of unbroken ligaments or transgranular facets, consistent with T-S specimens. For the TL orientation, however, the long axis of columnar grains is nearly normal to the crack growth direction and for welds B and C unbroken ligaments are relatively small, unlike the massive ligaments in T-S specimens. The different ligament sizes are due to different columnar grain orientations relative to the cracking direction. When an advancing T-L crack encounters a region with particularly susceptible columnar grain boundaries, rapid crack extension occurs normal to the macroscopic cracking direction thereby causing the crack to penetrate into and eventually circumnavigate the ligament. By contrast, when an advancing T-S crack encounters a region with particularly susceptible columnar grain boundaries, SCC fingers rapidly advance in the macroscopic cracking direction, further exacerbating ligament formation. Intergranular SCC incubation in T-L oriented specimens involves the same three mechanisms operative in T-S oriented welds, although there are notable differences. Because fewer grain boundaries intersect the tip of a T-L oriented precrack, environmental transgranular cracking becomes a more important incubation mechanism. Once a transgranular crack reaches a susceptible grain boundary, it transitions into an intergranular crack. SCC incubation along off-angle grain boundaries is also important. Off-angle cracks in the T-L orientation are readily incorporated into the primary intergranular fracture surface, so there is an abrupt transition from a fatigue precrack to an intergranular stress corrosion crack. 100-Minute Unload-Reload Cycles Fracture surfaces generated at 316C and 338C with a 100min. unload-reload cycle, which consist of vast fields of wavy intergranular cracking with isolated islands of transgranular facets and unbroken ligaments, are almost identical to those generated under constant-load conditions. The only discernable difference occurs along the precrack/SCC interface where the 100-min. cycle produces a combination of intergranular SCC and transgranular corrosion fatigue. Intergranular fingers quickly jut out ahead of the precrack where susceptible grain boundaries coincide with or intersect a precrack, consistent with constant-load SCC. In between the SCC fingers, however, unload-reload cycles produce a transgranular corrosion fatigue mechanism consisting of ill-defined facets. Once corrosion fatigue cracks propagate a short distance, they encounter susceptible grain boundaries and transition into intergranular cracks. Thereafter, cracking is predominantly intergranular and indistinguishable from constant-load SCC. Because the intergranular crack propagation mechanism is unaffected by unload-reload cycles every 100 min., it is not surprising that CGRs for these cyclic tests are consistent with rates obtained in constant-load tests. Data scatter is diminished in tests with 100-min. cycles as corrosion fatigue promotes crack incubation in regions that resist intergranular SCC. 10-Minute Unload-Reload Cycles The T-L and T-S welds subjected to 10-min. unload-reload cycles exhibit a mixed cracking mode with nearly equal amounts of intergranular and transgranular cracks. While the convoluted nature of the intergranular cracks is comparable to that for constant-loaded specimens, extensive transgranular faceting indicates that a mechanical fatigue component

accelerates cracking. Indeed, CGRs obtained at 338C with a 10-min. unload-reload cycle are 1.5X to 4X greater than rates obtained under constant load or with a 100-min. cycle. Acknowledgement This work was performed under U.S. Department of Energy Contract with Bettis Atomic Power Laboratory. References
1. W.J. Mills and C.M. Brown, Stress Corrosion Crack Growth

Rates for Alloy 82H Welds in High Temperature Water, 11th International Conference on Environmental Degradation of Materials in Nuclear Power SystemsWater Reactors (ANS, 2003) 1240-1252. 2. R.S. Pathania, A.R. McIlree and J. Hickling, Overview of Primary Water Cracking of Alloys 182/82 in PWRs, Fontevraud V International Symposium: Contribution of Materials Investigation to the Resolution of Problems Encountered in Pressurized Water Reactors (SFEN, France, September 23-27, 2002) 13-27. 3. W.H. Bamford, J. Foster, K.R. Hsu, L. Tunon-Sanjur and A.R. McIlree, Alloy 182 Weld Crack Growth, and Its Impact on ServiceInduced Cracking in Operating PWR Plant Piping, 10th International Conference on Environmental Degradation of Materials in Nuclear Power Systems-Water Reactors (NACE, 2002). 4. G. Embring, J. Lagerstrom, A. Jenssen, K. Norrgard and D.R. Tice, Assessment of Cracking in Dissimilar Metal Welds, 10th International Conference on Environmental Degradation of Materials in Nuclear Power Systems-Water Reactors (NACE, 2002). 5. J. Economou, A. Assice, F. Cattant, J. Salin and M. Stindel, Conroles et Expertises Metallurgiques de Traversees de Coucercle de Cuve, Fontevraud III Contribution of Materials Investigation to the Resolution of Problems Encountered in Pressurized Water Reactors, French Nuclear Energy Society, 1(1994) 197. 6. W. Bamford and J. Hall, A Review of Alloy 600 Cracking in Operating Nuclear Plants: Historical Experience and Future Trends, 11th International Conference on Environmental Degradation of Materials in Nuclear Power Systems Water Reactors (ANS, 2003) 1071-1079. 7. S. Le Hong, J.M. Boursier, C. Amzallag and J. Daret, Measurements of Stress Corrosion Cracking Growth Rates in Weld Alloy 182 in Primary Water of PWR, 10th International Conference on Environmental Degradation of Materials in Nuclear Power Systems-Water Reactors (NACE, 2002). 8. D.S. Morton, S.A. Attanasio and G.A. Young, Primary Water SCC Understanding and Characterization Through Fundamental Testing in the Vicinity of the Nickel/Nickel Oxide Phase Transition, 10th International Conference on Environmental Degradation of Materials in Nuclear Power Systems-Water Reactors (NACE, 2002). 9. R. Lindstrm, P. Lidar and J. Lagerstrm, Crack Growth of Alloy 182 in a Simulated Primary Side PWR Environment, 8th International Symposium on Environmental Degradation of Materials in Nuclear Power SystemsWater Reactors, ANS, 1997, pp. 422-429. 10. S.A. Attanasio and D.S. Morton, Measurement of the Nickel/Nickel Oxide Transition in Ni-Cr-Fe Alloys and Updated Data Correlations to Quantify the Effect of Aqueous Hydrogen on Primary Water SCC, 11th International Conference on Environmental Degradation of Materials in Nuclear Power SystemsWater Reactors (ANS, 2003) 143-155. 11. C.M. Brown, W.J. Mills and M.G. Burke, Effect of Microstructure on Low Temperature Cracking Behavior of EN82H Welds, 10th International Conference on Environmental Degradation of Materials in Nuclear Power SystemsWater Reactors (NACE, 2002).

674

Figure 1. Equivalent CGRs at 338C as a function of KI for tests conducted in 50 cc H2/kg H2O. Each data set is regressed using a power-law relationship with a KI exponent of 1.6. Solid data points denote constant-load conditions. Dot or cross-hair inside symbol denotes unload-reload cycles every 100 or 10 minutes, respectively.

Figure 2. Equivalent CGRs at 338C as a function of KI for tests conducted in 150 cc H2/kg H2O. Each data set is regressed using a power-law relationship with a KI exponent of 1.6.

(a)

(b)

(c)

(d)

(e)

(f)

Figure 3. Macroscopic fracture surface morphology: fatigue precrack at bottom, SCC region near center and post-test fatigue apart near top. (a)-(d) T-S orientation where crack growth direction is nearly parallel to columnar grains. Uneven cracking with SCC fingers separated by unbroken ligaments. (e) T-L orientation where crack growth direction is nearly normal to columnar grains. Although SCC is rather uniform with 100% engagement, unbroken ligaments remain in wake of advancing crack. (f) T-L oriented specimen from weld A1. SCC was consistently retarded near the weld crown (left). 675

Figure 4. Ratio of maximum to average crack extension (DaMAX / DaAVE) as a function of DaAVE for stress corrosion cracks in Alloy 82H welds. As DaAVE values increase, the DaAVE to DaMAX ratio decreases and for DaAVE values between 3 and 7 mm, the ratio becomes 1.3 0.2.

Figure 5. Maximum versus average crack extension (DaMAX v. DaAVE) for stress corrosion cracks in Alloy 82H welds. The difference between DaMAX and DaAVE, given by the distance between each data point and the dashed line, represents the degree of unevenness for a crack.

(a)

50 mm

(b)

10 mm

Figure 6. Microstructure of Alloy 82H welds showing dendritic grain structure with grain boundaries separating colonies of dendrites with different orientations. (a) Columnar grains with convoluted grain boundaries due to interpenetrating nature of dendrites. (b) SEM micrograph showing wavy grain boundaries decorated with fine carbides. Ti(C,N) inclusions (white particles) in interdendritic regions were often surrounded by fine Nb-rich carbonitrides.

676

(a)

50 mm

(b)

50 mm

(c)

50 mm

(d)

20 mm

Figure 7. Metallographic profiles of stress corrosion cracks generated under constant load conditions. (a) Bright field micrograph showing primary and secondary IG SCC following convoluted grain boundaries. (b) Interference contrast micrograph of (a) showing that IG cracks propagate along grain boundaries with greater apparent misorientation. (c) Discontinuous IG SCC with unbroken ligaments. (d) Differential interference micrograph of TG cracks attempting to link up two IG cracks at different elevations.

TG IG IG TG IG

(a)

20 mm

(b)

20 mm

Figure 8. SEM fractographs of T-S oriented weld C4 that was tested in 338C water under constant load. (a) Classic IG SCC with columnar grains oriented nearly parallel to crack growth direction (arrow). (b) Isolated islands of transgranular faceting surrounded by intergranular cracking. Secondary intergranular cracks often form at IG/TG interfaces.
677

(a)

20 mm

(b)

50 mm

TG

(c)

20 mm

(d)

20 mm

(e)

50 mm

(f)

20 mm

Figure 9. SEM fractographs of fatigue precrack/SCC interface in T-S oriented welds tested in 338C water under constant load. (a) IG cracking appears to initiate directly ahead of fatigue precrack, but a steep, narrow band of TG faceting separates precrack and IG SCC. (b) On far left, IG SCC initiates directly from tip of precrack. Near center, a narrow band of facets separates precrack and IG SCC. On right side, a reverse IG finger extends into precrack. (c) Off-angle IG crack extends into precrack. (d) Off-angle IG crack (center) and reverse IG finger (far left) extend into precrack region. (e) Reverse IG finger (left of center) that initiated along an off-angle crack extends into precrack. A second off-angle crack initiated at tip of precrack (center). (f) High magnification of off-angle cracks in (e). Note steep narrow TG wall between precrack and IG SCC. 678

(a)

20 mm

(a)

100 mm

Figure 10. SEM fractographs of precrack/IG SCC/post-test fatigue crack interface (denoted by lines) in HTH Alloy X-750 tested in 680F water under constant load. Tests were interrupted just after SCC incubation. (a) IG SCC initiates at a single location where an off-angle IG crack extends into precrack. (b) IG SCC incubation in three regions: Off-angle IG crack on far left. Off-angle IG crack leads to primary IG cracking beyond precrack in center. On far right, secondary and primary IG cracking immediately ahead of precrack.

(a)

20 mm

(b)

20 mm

(c)

20 mm

(d)

100 mm

Figure 11. T-L oriented weld C1 tested under constant-load conditions. (a) Classic IG SCC with columnar grains oriented normal to crack growth direction (arrow). (b) Island of TG facets within a field of IG SCC. (c) Precrack/IG SCC interface in weld C2, where IG SCC initiates directly from precrack (left) or after narrow TG faceted region (right). (d) Precrack/IG SCC interface in weld A1. Jog in interface is a vertical shear crack. To left of jog, a TG faceted region separates precrack and IG SCC. To right of jog, there is a sharp transition from the precrack to IG SCC. Dashed line is a projection of the precrack. Apparently, off-angle IG cracking occurred behind precrack tip where a susceptible grain boundary intersected the precrack. The vertical jog formed when the off-angle IG crack linked up with the primary IG crack. On mating fracture surface, the precrack represented by the dashed line is a secondary crack lying below the primary IG crack. 679

(a)

10 mm

(b)

20 mm

IG
TG IG F(P-T)

CF IG

(c)

20 mm

(d)

Precrack

20 mm
F(P-T)

IG

F(P-T) F(P-T)

IG

F(P-T) IG CF CF
(e)

IG

Precrack

50 mm

(f)

Precrack

20 mm

Figure 12. T-S weld C4 tested in 316C and 338C water with unload-reload cycles every 100 minutes. Black arrows denote macroscopic crack growth direction and root-to-crown direction. (a) IG SCC with columnar grains aligned in cracking direction. (b) Typical IG SCC. (c) IG SCC, TG facets produced by environmental cracking (left of center), and unbroken ligament that failed during post-test fatigue apart [F(P-T)]. Local environmental cracking directions are indicated by white arrows. (d) Interface between precrack and environmental cracking. On left side of fractograph, IG SCC emanates from off-angle IG crack that extends into precrack. In center and right side, corrosion fatigue region separates precrack and IG SCC. Once corrosion fatigue region transitions into IG SCC, IG cracking remains dominant. (e)(f) IG SCC fingers jut out ahead of TG corrosion fatigue crack. Lines denote tip of precrack. 680

(a)

50 mm

(b)

10 mm

(c)

50 mm

(d)

5 mm

Figure 13. SEM fractographs of welds tested in 338C water with unload-reload cycles every 10 minutes. (a) T-S weld C4. Combination of IG cracking (left) and TG faceting (right). (b) T-S weld C4. Typical ill-defined facets in TG regions. (c) T-L weld A1. TG faceting surrounded by IG cracking. (d) High magnification of fine crystallographic facets in (c).

(a)

20 mm

(b)

20 mm

Figure 14. Metallographic profiles of T-L weld A1 that was tested in 338C water with unload-reload cycles every 10 minutes. (a) Differential interference contrast image showing TG crack along persistent slip band between two IG cracks, which gives rise to coarse crystallographic facets. (b) TG cracks emanating from secondary IG crack. TG cracking along a series of closely spaced slip bands gives rise to fine crystallographic facets. 681

20 mm
IG CF IG CF

CF

IG

CF

Precrack

(a)

Precrack

(b)

20 mm

Figure 15. Precrack/environmental crack interface (denoted by lines) for T-S oriented weld C4 that was tested in 640F water with unload-reload cycles every 10 minutes. Black and white arrows denote macroscopic and local crack growth directions. Corrosion fatigue cracks initiate along the fatigue precrack, except where susceptible grain boundaries are in close proximity to the precrack. Cracking along susceptible boundaries produces SCC fingers that jut out beyond the precrack during the early stages of SCC. The SCC fingers then serve as initiation sites for the TG facets (i.e., crack extension in faceted growth regions is away from IG fingers). (a) TG facets emanate from IG finger that juts out beyond precrack. Eventually, IG becomes the dominant mechanism in this region. (b) TG facets emanate from IG finger clearly demonstrating that the IG crack formed long before corrosion fatigue crack.

682

You might also like