logo

Fractography P2


Cleavage is a low-energy fracture that propagates along well-defined low-index crystallographic planes known as cleavage planes. Theoretically, a cleavage fracture should have perfectly matching faces and should be completely flat and featureless. However, engineering alloys are polycrystalline and contain grain and subgrain boundaries, inclusions, dislocations, and other imperfections that affect a propagating cleavage fracture so that true, featureless cleavage is seldom observed. These imperfections and changes in crystal lattice orientation, such as possible mismatch of the low- index planes across grain or subgrain boundaries, produce distinct cleavage fracture surface features, such as cleavage steps, river patterns, feather markings, chevron (herringbone) patterns, and tongues (Ref 7). As shown schematically in Fig. 11, cleavage fractures frequently initiate on many parallel cleavage planes. As the fracture advances, however, the number of active planes decreases by a joining process that forms progressively higher cleavage steps. This network of cleavage steps is known as a river pattern. Because the branches of the river pattern join in the direction of crack propagation, these markings can be used to establish the local fracture direction. Fig. 11 Schematic of cleavage fracture formation showing the effect of subgrain boundaries. (a) Tilt boundary. (b) Twist boundary A tilt boundary exists when principal cleavage planes form a small angle with respect to one another as a result of a slight rotation about a common axis parallel to the intersection (Fig. 11a). In the case of a tilt boundary, the cleavage fracture path is virtually uninterrupted, and the cleavage planes and steps propagates across the boundary. However, when the principal cleavage planes are rotated about an axis perpendicular to the boundary, a twist boundary results (Fig. 11b). Because of the significant misalignment of cleavage planes at the boundary, the propagating fracture reinitiates at the boundary as a series of parallel cleavage fracture connected by small (low) cleavage steps. As the fracture propagates away from the boundary, the numerous cleavage planes join, resulting in fewer individual cleavage planes and higher steps. Thus, when viewing a cleavage fracture that propagates across a twist boundary, the cleavage steps do not cross but initiate new steps at the boundary (Fig. 11b) Most boundaries, rather than being simple tilt or twist, are a combination of both types and are referred to as tilt-twist boundaries. Cleavage fracture exhibiting twist and tilt boundaries are shown in Fig. 12(a) and 13, respectively. Fig. 12 Examples of cleavage fractures. (a) Twist boundary, cleavage steps, and river patterns in an Fe-0.01C- 0.24Mn-0.02Si alloy that was fractured by impact. (b) Tongues (arrows) on the surface of a 30% Cr steel weld metal that fractured by cleavage Fig. 13 Cleavage fracture in Armco iron showing a tilt boundary, cleavage steps, and river patterns. TEM p-c replica Feather markings are a fan-shaped array of very fine cleavage steps on a large cleavage facet (Fig. 14a). The apex of the fan points back to the fracture origin. Large cleavage steps are shown in Fig. 14(b). Fig. 14 Examples of cleavage fractures. (a) Feather pattern on a single grain of a chromium steel weld metal that failed by cleavage. (b) Cleavage steps in a Cu-25 at.% Au alloy that failed by transgranular stress- corrosion cracking. (B.D. Lichter, Vanderbilt University) Tongues are occasionally observed on cleavage fracture (Fig. 12b). They are formed when a cleavage fracture deviates from the cleavage plane and propagates a short distance along a twin orientation (Ref 8). Wallner lines (Fig. 15) constitute a distinct cleavage pattern that is sometimes observed on fracture surfaces of brittle nonmetallic materials or on brittle inclusions or intermetallic compounds. This structure consist of two sets of parallel cleavage steps that often intersect to produce a crisscross pattern. Wallner lines result from the interaction of a simultaneously propagating crack front and an elastic shock wave in the material (Ref 9). Fig. 15 Wallner lines (arrow) on the surface of a fractured WC-Co specimen. TEM formvar replica. Etched with 5% HCl. (S.B. Luyckx, University of the Witwatersrand) Fatigue A fracture that is the result of repetitive or cyclic loading is known as a fatigue fracture. A fatigue fracture generally occurs in three stages: it initiates during Stage I, propagates for most of its length during Stage II, and proceeds to catastrophic fracture during Stage III. Fatigue crack initiation and growth during Stage I occurs principally by slip-plane cracking due to repetitive reversals of the active slip systems in the metal (Ref 10, 11, 12, 13, 14). Crack growth is strongly influenced by microstructure and mean stress (Ref 15), and as much as 90% of the fatigue life may be consumed in initiating a viable fatigue crack (Ref 16). The crack tends to follow crystallographic planes, but changes direction at discontinuities, such as grain boundaries. At large plastic-strain amplitudes, fatigue cracks may initiate at grain boundaries (Ref 14). A typical State I fatigue fracture is shown in Fig. 16. State I fatigue fracture surfaces are faceted, often resemble cleavage, and do not exhibit fatigue striations. Stage I fatigue is normally observed on high-cycle low-stress fractures and is frequently absent in low- cycle high-stress fatigue. Fig. 16 Stage I fatigue appearance. (a) Cleavagelike, crystallographically oriented State I fatigue fracture in a cast Ni-14Cr-4.5Mo-1Ti-6Al-1.5Fe-2.0(Nb + Ta) alloy. (b) Stair-step fracture surface indicative of Stage I fatigue fracture in a cast ASTM F75 cobalt-base alloy. SEM. (R. Abrams, Howmedica, Div. Pfizer Hospital Products Group Inc.) The largest portion of a fatigue fracture consists of Stage II crack growth, which generally occurs by transgranular fracture and is more influenced by the magnitude of the alternating stress than by the mean stress or microstructure (Ref 15, 17, 18). Fatigue fractures generated during Stage II fatigue usually exhibit crack-arrest marks known as fatigue striations (Fig. 17, 18, 19, 20, 21, 22), which are a visual record of the position of the fatigue crack front during crack propagation through the material. Fig. 17 Uniformly distributed fatigue striations in an aluminum 2024-T3 alloy. (a) Tear ridge and inclusion (outlined by rectangle). (b) Higher-magnification view of the region outlined by the rectangle in (a) showing the continuity of the fracture path through and around the inclusion. Compare with Fig. 18. Fig. 18 Local variations in striation spacing in a Ni-0.04C-21Cr-0.6Mn-2.5Ti-0.7Al alloy that was tested under rotating bending conditions. Compare with Fig. 17(b). Fig. 19 Fatigue striations in a 2024-T3 aluminum alloy joined by tear ridges Fig. 20 Fatigue striations on adjoining walls on the fracture surface of a commercially pure titanium specimen. (O.E.M. Pohler, Institut Straumann AG) Fig. 21 Fatigue striations on the fracture surface of a tantalum heat-exchanger tube. The rough surface appearance is due to secondary cracking caused by high-cycle low-amplitude fatigue. (M.E. Blum, FMC Corporation) Fig. 22 High-magnification views of fatigue striations. (a) Striations (arrow) on the fracture surface of an austenitic stainless steel. (C.R. Brooks and A. Choudhury, University of Tennessee). (b) Fatigue striations on the facets of tantalum grains in the heat-affected zone of a weldment. (M.E. Blum, FMC Corporation) There are basically two models that have been proposed to explain Stage II striation-forming fatigue propagation. One is based on plastic blunting at the crack tip (Ref 11). This model cannot account for the absence of striations when a metal is fatigue tested in vacuum and does not adequately predict the peak-to-peak and valley-to-valley matching of corresponding features on mating halves of the fracture (Ref 8, 19, 20, 21, 22, 23). The other model, which is based on slip at the crack tip, accounts for conditions where slip may not occur precisely at the crack tip due to the presence of lattice or microstructural imperfections (Ref 19, 20, 21). This model (Fig. 23) is more successful in explaining the mechanism by which Stage II fatigue cracks propagate. The concentration of stress at a fatigue crack results in plastic deformation (slip) being confined to a small region at the tip of the crack while the remainder of the material is subjected to elastic strain. As shown in Fig. 23(a), the crack opens on the rising-tension portion of the load cycle by slip on alternating slip planes. As slip proceeds, the crack tip blunts, but is resharpened by partial slip reversal during the declining-load portion of the fatigue cycle. This results in a compressive stress at the crack tip due to the relaxation of the residual elastic tensile stresses induced in the uncracked portion of the material during the rising load cycle (Fig. 23b). The closing crack does not reweld, because the new slip surfaces created during the crack- opening displacement are instantly oxidized (Ref 24), which makes complete slip reversal unlikely. Fig. 23 Mechanism of fatigue crack propagation by alternate slip at the crack tip. Sketches are simplified to clarify the basic concepts. (a) Crack opening and crack tip blunting by slip on alternate slip planes with increasing tensile stress. (b) Crack closure and crack tip resharpening by partial slip reversal on alternate slip planes with increasing compressive stress The essential absence of striations on fatigue fracture surfaces of metals tested in vacuum tends to support the assumption that oxidation reduces slip reversal during crack closure, which results in the formation of striations (Ref 19, 25, 26). The lack of oxidation in hard vacuum promotes a more complete slip reversal (Ref 27), which results in a smooth and relatively featureless fatigue fracture surface. Some fracture surfaces containing widely spaced fatigue striations exhibit slip traces on the leading edges of the striation and relatively smooth trailing edges, as predicted by the model (Fig. 23). Not all fatigue striations, however, exhibit distinct slip traces, as suggested by Fig. 23, which is a simplified representation of the fatigue process. As shown schematically in Fig. 24, the profile of the fatigue fracture can also vary, depending on the material and state of stress. Materials that exhibit fairly well-developed striations display a sawtooth-type profile (Fig. 24a) with valley-to- valley or groove-to-groove matching (Ref 23, 28). Low compressive stresses at the crack tip favor the sawtooth profile; however, high compressive stresses promote the groove-type fatigue profile, as shown in Fig. 24(c) (Ref 23, 28). Jagged, poorly formed, distorted, and unevenly spaced striations (Fig. 24b), sometimes termed quasi-striations (Ref 23), show no symmetrical matching profiles. Even distinct sawtooth and groove-type fatigue surfaces may not show symmetrical matching. The local microscopic plane of a fatigue crack often deviates from the normal to the principal stress. Consequently, one of the fracture surfaces will be deformed more by repetitive cyclic slip than its matching counterpart (Ref 29) (for an analogy, see Fig. 9). Thus, one fracture surface may show well-developed striations, while its counterpart exhibits shallow, poorly formed striations. Fig. 24 Sawtooth and groove-type fatigue fracture profiles. Arrows show crack propagation direction. (a) Distinct sawtooth profile (aluminum alloy). (b) Poorly formed sawtooth profile (steel). (c) Groove-type profile (aluminum alloy). Source: Ref 23 Under normal conditions, each striation is the result of one load cycle and marks the position of the fatigue crack front at the time the striation was formed. However, when there is a sudden decrease in the applied load, the crack can temporarily stop propagating, and no striations are formed. The crack resumes propagation only after a certain number of cycles are applied at the lower stress (Ref 4, 23, 30). This phenomenon of crack arrest is believed to be due to the presence of a residual compressive-stress field within the crack tip plastic zone produced after the last high-stress fatigue cycle (Ref 23, 30). Fatigue crack propagation and therefore striation spacing can be affected by a number of variables, such as loading conditions, strength of the material, microstructure, and the environment, for example, temperature and the presence of corrosive or embrittling gases and fluids. Considering only the loading conditions--which would include the mean stress, the alternating stress, and the cyclic frequency--the magnitude of the alternating stress (σmax - σmin) has the greatest effect on striation spacing. Increasing the magnitude of the alternating stress produces an increase in the striation spacing (Fig. 25a). While rising, the mean stress can also increase the striation spacing; this increase is not as great as one for a numerically equivalent increase in the alternating stress. Within reasonable limits, the cyclic frequency has the least effect on striation spacing. In some cases, fatigue striation spacing can change significantly over a very short distance (Fig. 25b). This is due in part to changes in local stress conditions as the crack propagates on an inclined surface. Fig. 25 Variations in fatigue striation spacing. (a) Spectrum-loaded fatigue fracture in a 7475-T7651 aluminum alloy test coupon showing an increase in striation spacing due to higher alternating stress. (b) Local variation in fatigue striation spacing in a spectrum-loaded 7050-T7651 aluminum alloy extrusion. (D. Brown, Douglas Aircraft Company) For a Stage II fatigue crack propagating under conditions of reasonably constant cyclic loading frequency and advancing within the nominal range of 10-5 to 10-3mm/cycle,* the crack growth rate, da/dN, can be expressed as a function of the stress intensity factor K (Ref 15, 31, 32): da = C (∆K ) m (Eq 1) dN where a is the distance of fatigue crack advance, N is the number of cycles applied to advance the distance a, m and C are constants, and ∆K = Kmax - Kmin is the difference between the maximum and minimum stress intensity factor for each fatigue load cycle. The stress intensity factor, K, describes the stress condition at a crack and is a function of the applied stress and a crack shape factor, generally expressed as a ratio of the crack depth to length. When a fatigue striation is produced on each loading cycle, da/dN represents the striation spacing. Equation 1 does not adequately describe Stage I or Stage III fatigue crack growth rates; it tends to overestimate Stage I and often underestimates Stage III growth rates (Ref 15). Stage III is the terminal propagation phase of a fatigue crack in which the striation-forming mode is progressively displaced by the static fracture modes, such as dimple rupture or cleavage. The rate of crack growth increases during Stage III until the fatigue crack becomes unstable and the part fails. Because the crack propagation is increasingly dominated by the static fracture modes, Stage III fatigue is sensitive to both microstructure and mean stress (Ref 17, 18). Characteristics of Fractures With Fatigue Striations. During Stage II fatigue, the crack often propagates on multiple plateaus that are at different elevations with respect to one another (Fig. 26). A plateau that has a concave surface curvature exhibits a convex contour on the mating fracture face (Ref 29). The plateaus are joined either by tear ridges or walls that contain fatigue striations (Fig. 19 and 20a). Fatigue striations often bow out in the direction of crack propagation and generally tend to align perpendicular to the principal (macroscopic) crack propagation direction. However, variations in local stresses and microstructure can change the orientation of the plane of fracture and alter the direction of striation alignment (Fig. 27). Fig. 26 Schematic illustrating fatigue striations on plateaus Fig. 27 Striations on two joining, independent fatigue crack fronts on a fracture surface of aluminum alloy 6061-T6. The two arrows indicate direction of local crack propagation. TEM p-c replica Large second-phase particles and inclusions in a metal can change the local crack growth rate and resulting fatigue striation spacing. When a fatigue crack approaches such a particle, it is briefly retarded if the particle remains intact or is accelerated if the particle cleaves (Fig. 18). In both cases, however, the crack growth rate is changed only in the immediate vicinity of the particle and therefore does not significantly affect the total crack growth rate. However, for low- cycle (high-stress) fatigue, the relatively large plastic zone at the crack tip can cause cleavage and matrix separation at the particles at a significant distance ahead of the advancing fatigue crack. The cleaved or matrix-separated particles, in effect, behave as cracks or voids that promote a tear or shear fracture between themselves and the fatigue crack, thus significantly advancing the crack front (Ref 33, 34). Relatively small, individual particles have no significant effect on striation spacing (Fig. 17b). The distinct, periodic markings sometimes observed on fatigue fracture surfaces are known as tire tracks, because they often resemble the tracks left by the tread pattern of a tire (Fig. 28). These rows of parallel markings are the result of a particle or a protrusion on one fatigue fracture surface being successively impressed into the surface of the mating half of the fracture during the closing portion of the fatigue cycle (Ref 23, 29, 34). Tire tracks are more common for the tension- compression than the tension-tension type of fatigue loading (Ref 23). The direction of the tire tracks and the change in spacing of the indentations within the track can indicate the type of displacement that occurred during the fracturing process, such as lateral movement from shear or torsional loading. The presence of tire tracks on a fracture surface that exhibits no fatigue striations may indicate that the fracture occurred by low-cycle (high-stress) fatigue (Ref 35). Fig. 28 Tire tracks on the fatigue fracture surface of a quenched-and-tempered AISI 4140 steel. TEM replica. (I. Le May, Metallurgical Consulting Services Ltd.) Decohesive Rupture
DMCA.com Protection Status Copyright by webtailieu.net