Вы находитесь на странице: 1из 13

Fatigue & Fracture of Engineering Materials & Structures

doi: 10.1111/j.1460-2695.2008.01234.x

Fatigue behaviour of friction stir welded AA2024-T3 alloy: longitudinal and transverse crack growth
M. T. MILAN, W. W. BOSE FILHO, C. O. F. T. RUCKERT and J. R. TARPANI
Department of Materials, Aeronautics and Automotive Engineering, Engineering School of S ao Carlos, University of S ao Paulo, Av. Trabalhador S ao-Carlense, 400, Centro, CEP. 13.566-590, S ao Carlos-SP, Brazil Received in final form 20 April 2007

A B S T R A C T The fatigue crack growth properties of friction stir welded joints of 2024-T3 aluminium

alloy have been studied under constant load amplitude (increasing- K ), with special emphasis on the residual stress (inverse weight function) effects on longitudinal and transverse crack growth rate predictions (Glinkas method). In general, welded joints were more resistant to longitudinally growing fatigue cracks than the parent material at threshold K values, when beneficial thermal residual stresses decelerated crack growth rate, while the opposite behaviour was observed next to K C instability, basically due to monotonic fracture modes intercepting fatigue crack growth in weld microstructures. As a result, fatigue crack growth rate (FCGR) predictions were conservative at lower propagation rates and non-conservative for faster cracks. Regarding transverse cracks, intense compressive residual stresses rendered welded plates more fatigue resistant than neat parent plate. However, once the crack tip entered the more brittle weld region substantial acceleration of FCGR occurred due to operative monotonic tensile modes of fracture, leading to non-conservative crack growth rate predictions next to K C instability. At threshold K values non-conservative predictions values resulted from residual stress relaxation. Improvements on predicted FCGR values were strongly dependent on how the progressive plastic relaxation of the residual stress field was considered. Keywords aluminium alloy; crack growth rate prediction; fatigue; friction stir welding; residual stress.
NOMENCLATURES

a = crack length AA2024-T3 = high-strength aluminium alloy grade d = slot aperture da/dN = crack growth rate E = plane-stress Youngs modulus E = plane-strain Youngs modulus EL = elongation at fracture FCGR = fatigue crack growth rate(s) FSW = friction stir welding h(x,a) = weight function HAZ = heat-affected zone K C = critical stress intensity factor K MAX = maximum applied stress intensity in fatigue K Ir = residual stress intensity factor in mode I of crack opening

Correspondence: J. R. Tarpani. E-mail: jrpan@sc.usp.br

526

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

FATIGUE BEHAVIOUR OF FRICTION STIR WELDED AA2024-T3 ALLOY

527

K rx , K ry = residual stress intensity factor distribution on x and y directions L = length of generic test piece L 0 = original gage length M = location of strain-gage R = stress ratio R = effective stress ratio RA = reduction in area at fracture S = engineering, nominal, remote or gross stress S, L(x), T (y) = three main orthogonal metallographic axes or directions TP = test piece TMAZ = thermo-mechanically affected zone UTS = ultimate tensile strength W = width of generic test piece WEDM = wire electro-discharge machine YS = yield strength Z(a) = influence function K (th) = range of stress intensity factor in fatigue (threshold value) = strain M = strain measured at M position = Poissons coefficient rx , ry = residual stress distribution on x and y directions On the other hand, compressive residual stresses reduce the fatigue crack growth rate (FCGR) by decreasing the effective stress ratio. Additionally, residual stresses were found to affect initiation fracture toughness values (K C ) of aluminium alloys.6,7 Transverse cracks present an even more complex situation, inasmuch as the defect propagates from the parent material towards the weld region. The crack tip intersects different microstructures and distinct intrinsic fatigue cracking behaviours can be expected. This paper presents data obtained from studying the fatigue crack resistance of FSW joints of aeronautical grade AA2024-T3 high-strength alloy containing either longitudinal or transverse cracks. The effective stress ratio method (Glinkas R )8 was employed to predict FCGR in the welded alloy taking into account the residual stress intensity factor calculated by the slitting or cut compliance method9 and parent material fatigue properties. Predicted crack growth rates are then compared to experimental values. The authors expect to contribute to the still limited body of knowledge regarding FSW materials by providing useful fatigue crack growth data for both safe-life and damage-tolerant designs.
INCREMENTAL SLITTING AND WEIGHT FUNCTION METHODS

INTRODUCTION

In recent years, friction stir welding (FSW), a solid-state joining technique, has been considered a potential technique to replace conventional riveting operations and fusion welding methods (e.g. laser and electron beam) in aircraft manufacture. However, the weld still results in a continuous medium for crack propagation and hence, knowledge of the fatigue and fracture properties of such classes of materials is vital if a damage-tolerant design is adopted. Advantages of FSW process include: design simplification (easy periodical inspection, less macro stressconcentrators), low distortion, poreless welding process (less micro-stress concentrators), static strength as high as 80100% of parent material, improved fatigue performance and better load distribution. FSW disadvantages comprise: lack of extensive data on mechanical properties, continuous crack propagation medium, mismatching between plastic properties of weld and parent metals and residual stress effects. Residual stresses are invariably present in welded structures after fabrication. They are likely to affect mechanical and corrosion properties of the materials and therefore influence the in-service performance of structural components. The effects of residual stresses on fatigue crack propagation have been reported by several authors such as Itoh et al.,1 Bussu and Irving2 and Milan and Bowen.3,4 Based on Parkers superposition principle,5 they concluded that tensile residual stresses increase the crack growth rate due to increasing effective stress ratio (R ).

The cut compliance or incremental slitting method is a helpful technique to determine both the near surface and through thickness residual stress profiles. It is based on

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

528

M . T . M I L A N et al.

the fact that when a cut, simulating a crack, is incrementally introduced into a part, the residual stresses are relieved, causing the part to deform. Such deformation can be sensed by strain gauges attached at specific positions of the part (Fig. 1) and the residual stress intensity factor profile can be derived.9,10 Assuming a sufficiently narrow slot (d a), linear elastic fracture mechanics (LEFM) can be employed to establish a relationship (Eq. 1) between the measured strains, , and the corresponding residual stress intensity factor in opening mode, K Ir :9 K Ir (a ) = E d M , Z(a ) d a (1)

stress distribution in the flawless test specimen, r :12


a

K Ir (a ) =
0

h ( x , a ) r ( x ) d x

(4)

where M is the measured strain at the back face position M during the cutting procedure, a is the slot length, E , the generalized form of the Youngs modulus (E = E for plane-stress, and E = E/1 2 for plane-strain conditions), and Z (a) the influence function that depends on the test-piece geometry, cut plane location and strain measurement position. Here Z (a) is considered as being independent of the residual stress profile. For a rectangular plate, where L > 2W , and taking strain measurements at the back face (position M ), Z (a) is given as follows:11
- for a/W < 0.2 (shallow crack):
2

Insofar as the weight function h(x,a) is universal for a given crack geometry, one can determine the inverse weight function, so that r can be determined from K Ir (i.e. inverse path). Therefore, by employing the slitting technique and the incremental stress method (inverse weight function) it is possible to obtain the original residual stress profile in the uncracked body, as well as the residual stress profile redistribution due to crack growth.
MATERIAL AND TEST SPECIMENS

Z(a ) =

2,532 (Wa )1.5

1 25
a W

a W 2

0.2

a W

5.926 0.2

0.288 0.2

+1 (2)

AA2024-T3 3.2 mm-thick plates were friction stir welded along the rolling direction (Fig. 2). Welding parameters are proprietary to the Brazilian aircraft manufacturer, Embraer S/A. Basic tensile properties of the parent alloy are provided in Table 1 for both longitudinal and transverse directions. Tri-dimensional views of, respectively, chemically etched and unetched microstructures of the material tested are shown in Fig. 3. Full-plate-thickness test pieces with in-plane dimensions of 60 120 mm2 were wire electro-discharge machined according Figs. 4 and 5 for, respectively, longitudinal and transverse crack growth specimens. No post-weld heat treatments were applied to the FSW plates, so that all experiments were carried out after all natural aging had ceased.
EXPERIMENTS AND ANALYSIS

- for 0.2<a/W < 1 (deep crack):

Residual stress intensity factor evaluation (3) A strain gauge was attached to the back face of each test piece, as shown in Figs. 4 and 5. The strain gauge and connection wires were covered with a plastic resin in order to avoid reaction with the environment. Shielded wires

Z(a ) =

2.532 (W a )1.5

Equation (4) dictates the relation between the residual stress concentration factor, K Ir (Eq. 1), and the residual

Fig. 1 Rectangular plate containing a slot along the central plane, with strain measurements taken at the location M.

Fig. 2 Freshly friction stir welded plates tested in this study.

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

FATIGUE BEHAVIOUR OF FRICTION STIR WELDED AA2024-T3 ALLOY

529

Table 1 Mechanical properties of the Al-alloy tested at ambient temperature Orientation/Property AA2024-T3 longitudinal Standard deviation AA2024-T3 transverse Standard deviation E (GPa) 79 4.6 81 2.0 UTS (MPa) 477 4.2 463 3.8 YS (MPa) 350 6.0 308 4.5 ELa (%) 21.1 0.36 21.4 0.31 RA (%) 20.8 2.46 23.0 1.45

Average of three test pieces for each specimen orientation. a L = 25 mm. 0

Fig. 3 Typical microstructures of the rolled A1-2024-T3 alloy: (a) Kellers etching; (b) Simply polished. Note different image magnifications (approximately 2:1).

Fig. 4 T-L testpiece configurations for measuring transverse residual stress intensity factor on y direction, K ry , and for fatigue crack propagation tests along the rolling or longitudinal (x) direction: (a) Slot at the centre of the weld line (0 mm), and (b) 5 mm displaced from zero position. Tool shoulder width is 16 mm. FSW tools travelling from the bottom to the top of the page and rotating in counter-clockwise direction.

were used to connect the strain gauges to the strain data acquisition apparatus in order to minimize electromagnetic interferences. Plate cutting was performed by wire electro-discharge machining (WEDM) either longitudi-

nally or transversally to the weldline. For the former case, as depicted in Fig. 4, the slot was introduced in two different positions: 0 mm and 5 mm distant from the weld nugget centre line (on the FSW tool advancing side), to

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

530

M . T . M I L A N et al.

Fig. 5 L-T testpiece configuration for measuring longitudinal residual stress intensity factor on x direction, K rx , and for fatigue crack propagation tests along transverse (y) direction. Strain measurments are taken at the point M. Tool shoulder width is 16 mm.

simulate a crack growing in the weld and in the heat affected zone, respectively. Recent work by Milan et al.13 has found WEDM more practical, precise and less likely to introduce additional stresses compared to abrasive saw method. In all experiments, cutting increments of 0.5 mm were chosen. Readings were taken 5 min after each single slotting procedure had finished. After the strain data were obtained as a function of the slot length, the secant method was used to derive d /da data, which were then employed in Eq. (1) to determine corresponding K ry and K rx values, for longitudinal and transverse cracks, respectively. Fatigue crack growth tests A 5-mm-long notch was introduced at the edge of the welded test pieces, in the positions indicated in Figs. 4 and 5. Once a 5-mm-long fatigue pre-crack emanated from the notch tip, mode I fatigue loading tests were conducted at ambient temperature under constant amplitude loading condition (increasing- K testing). Sinusoidal waveform under a frequency of 60 Hz was applied to all the test specimens. The parent metal was tested under stress ratios (R) of 0, 0.3 and 0.5, respectively, in order to provide data for the Glincas R method predictions, while the weld metals where tested at R = 0.5 only. A detailed description of the R method can be found in Milan and Bowen.4

and negative in the entire range of acquired data. Rigorously speaking, a negative K r value does not exist for a closed crack, but it means that when an external load is applied, the thermal residual stress profile shields the crack, that is, effectively reduces the locally applied K factor, and hence decelerates FCGR. Clearly, this effect is a welcome safety margin against failure in damage-tolerant approaches. It is observed that K ry profiles vary only slightly from one test piece to another, so that no systematic trend can be established regarding the slot position in relation to the weldline, which means that nugget and thermomechanically affected zone (TMAZ) microstructures are similar to some extent with regard to fatigue properties. Using the inverse weight function method, as detailed by Schindler,9 it was possible to obtain the initial residual stress profile present in the material before the slot was introduced, as depicted in Fig. 6b. Tensile residual stress in the centre of the test piece and balancing compressive residual stresses near the edges can be noticed. From Fig. 6, it can be observed that compressive residual stresses may delay crack nucleation at the edge of a pristine component. On the other hand, tensile residual stresses may favour nucleation at the centre of the piece. These findings have important implications in safe-life approaches. Transverse cracking Figure 7a shows longitudinal residual stress intensity factor profile, K rx , for test pieces with the slot axis perpendicular to the weldline. A negative peak value develops at approximately 16 mm from the edge of the test piece, so that K rx remains negative up to the weld centre line (30 mm position from the specimen edge). Therefore, the FCGR of a growing crack approaching perpendicularly to the weldline is expected to be slowed down. However, after the crack crosses the weld region (i.e. positive K rx values in Fig. 7a) the FCGR will accelerate as compared to the crack growth rate in the parent material

RESULTS AND DISCUSSION

Residual stress intensity factor profile Longitudinal cracking Figure 6a presents transverse residual stress intensity factor profiles, K ry , obtained for longitudinal slot test pieces. For both cases, that is, for the slot introduced at 0 mm and 5 mm from the weld centre line, K ry values remain low

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

FATIGUE BEHAVIOUR OF FRICTION STIR WELDED AA2024-T3 ALLOY

531

Fig. 6 (a) Transverse residual stress intensity factor profiles, and (b) Initial residual stress profiles derived for longitudinal slot testpieces.

under identically applied stress intensity factor range and R ratio. In terms of non-destructive inspection programs of damage-tolerant structures, the above provided results point out the need for a much more complex surveillance schedule for welded structures as compared to monolithic (non-welded) ones in order to guarantee structural integrity for both short- and long-fatigue growing cracks. Figure 7b depicts the original residual stress profile in the flawless material, that is, before the slot introduction, as derived through the inverse weight function method.9 It is possible to observe the tensile residual stress approaching 80% of the yield strength of the parent plate in the weld region (in the advancing side of the tool), while near the edges the residual stresses remain compressive. A likely

crack nucleation at the TMAZ/HAZ zone can then be predicted.

FCGR curves (experimental vs. predicted) Longitudinal cracking Fatigue crack propagation rate curves of longitudinally slotted test pieces are presented in Fig. 8. For the same applied K level, a crack positioned on the centre of the weld nugget (0 mm position) exhibits fatigue growth rates slightly lower than a crack located along the TMAZ at about 4 mm from the weld centre line. Compared to the parent material however, all the welded test pieces

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

532

M . T . M I L A N et al.

Fig. 7 (a) Residual stress intensity factor, and (b) Corresponding initial residual stress profiles for transverse slot testpieces.

presented higher fatigue crack growth resistance at low applied K values. This is consistent with the presence of a negative peak of K ry at a slot length of order of 5 mm (Fig. 6a). Nonetheless for higher applied K val ues (above 10 MPa. m), that is, longer slots (cracks) and faster crack growth rate, although K ry remains negative during the entire range of testing, the welded material presented higher values of FCGR. This leads to the assumption that whereas microstructure and residual stresses act simultaneously in the threshold region, at K values approaching K C , monotonic fracture modes (dimple and/or cleavage micromechanisms) intercept fatigue crack propagation, that is, they accelerate FCGR as compared to

the parent plate behaviour. It is worth mentioning that at these crack tip positions only microstructural effects can play a role, because K ry values approach zero (refer to Fig. 6a). Figure 9 shows the fatigue crack propagation rates estimated via the so-called R method, for a test piece containing a crack in the weld centre line. For K values below 8 MPa. m, predicted values are conservative compared to the experimental ones, whereas for K values above 10 MPa. m, predicted values are otherwise nonconservative. Because the prediction model takes into account only the parent metal fatigue properties and the K ry profile of the welded material, it is possible that

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

FATIGUE BEHAVIOUR OF FRICTION STIR WELDED AA2024-T3 ALLOY

533

Fig. 8 Fatigue crack propagation curves of parent and welded materials, with cracks propagating along the weld centre line (0 mm) and TMAZ (5 mm), respectively.

Fig. 9 Experimental and predicted FCGR of the welded material, with the crack propagating along the weld centre line (0 mm).

likely that the intrinsic fatigue crack growth resistance of the nugget microstructure differs significantly from that of the parent plate. Such a hypothesis is supported by the fact that the discrepancies between predicted and experimental values are higher at low K values and near the final fracture region, where FCGR is much more dependent on microstructural features. Conversely, in the Paris region, where microstructure plays a minor role in FCGR, there is good correlation between experimental and predicted values. Briefly, recalling that at the threshold region (i.e. next to the K th value) conservative FCGR predictions were derived for the weld microstructure on the basis of the intrinsic fatigue behaviour of the parent plate, the results strongly suggest that at low applied K values the former microstructure is more damage tolerant than the parent metal. On the other hand, near the final catastrophic fracture region (i.e. approaching K C value), the nonconservatism of FCGR predictions for the weld nugget indicates that monotonic fracture modes (i.e. K MAX controlled mechanisms) are not precisely accounted for by Glinkas model, and evidences the parent metal as the toughest microstructure at high applied K values. For cracks positioned on the TMAZ and subjected to low K values, predicted and experimental FCGR are in good agreement (Fig. 11), indicating that the TMAZ microstructure exhibits the same intrinsic fatigue crack growth resistance as the parent material. It should be mentioned that as the crack grew during the increasingK tests, it deflected towards the interface between the TMAZ and the heat affected zone (HAZ), as seen in Fig. 12. As a result, FCGR predictions at higher K values were underestimated (i.e. non-conservative approach). Assuming that residual stress profiles for both TMAZ and HAZ regions are similar, the underestimated prediction suggests that the microstructure of the TMAZ/HAZ interface offers lower resistance to fatigue crack propagation than the parent plate material probably due to a lower fracture toughness value, which would contribute to stronger K MAX effect on fatigue crack growth resistance. This point will be revisited in the next section for transverse cracks. On the basis of Fig. 6a data, one could argue that a less compressive stress field would have also contributed to higher crack growth rates. Regardless of fact that the accelerated FCGR was generated by either a less tough microstructure or a less compressive stress field, or even by both effects simultaneously, the truth is that the experiment did fail to test Glinkas model. Transverse cracking Fatigue crack growth curves for cracks running perpendicularly to the weldline are presented in Fig. 13. Up to a position corresponding to the border of the tool

the microstructure of the weld nugget is responsible for the differences observed between predicted and experimentally measured figures. The nugget region is formed by fine equiaxed dynamically recrystallized grains,2 while the parent metal presents elongated grains due to the rolling process, as can be seen in Fig. 10. Therefore, it is

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

534

M . T . M I L A N et al.

Fig. 10 Typical microstructures developed in FSW aluminium alloy. FSW tool advances in the right hand side region. The nugget consists of recrystallized equiaxial grains, TMAZ of stretched grains due to rotational tool movement, and HAZ possesses similar appearance to the parent metal.

Fig. 12 Longitudinal fatigue crack growth specimen showing crack deflection towards the TMAZ/HAZ interface distant approximately 8 mm from weld centre line. Penetrant-dye testing was performed in order to facilitate crack visualization.

Fig. 11 Experimental and predicted FCGR of welded material, with the longitudinal crack propagating 5 mm from the weld centre line.

shoulder, results clearly show that the welded joint presents higher fatigue crack growth resistance than the parent material, for the same applied K value, or K MAX value since the nominal R value is the same. This is mainly

attributed to the strong negative K rx values shown in Fig. 7a, in a position still located in the parent metal. A negative K rx value reduces both the stress ratio, R, and the effective K values, so decreasing the net crack tip driving force. However, when the crack tip indeed enters the weld region, FCGR are found to be higher for the welded joint than for the parent material, although the K rx values still remain considerably negative (Fig. 7a).

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

FATIGUE BEHAVIOUR OF FRICTION STIR WELDED AA2024-T3 ALLOY

535

Fig. 13 Fatigue crack propagation curves of welded joint and parent material, with crack growing transverse to the weld line.

This fact suggests that the weld microstructure is brittle than the parent metal because at this position residual stresses approach zero. Figure 14 provides Vickers microhardness profile transverse to the weldline at the mid-thickness position of the welded joint. Although it agrees quite well with literature data for 2024-T3 Al-alloys,14 the relatively soft nugget could lead to the conclusion of a tougher microstructure if compared to the parent plate. However, recent study by Kamp et al.15 has shown that alloys containing precipitate distributions that are unstable at elevated temperature, such as aerospace aluminium alloys, are quite prone to profound microstructural changes when friction stir

welded. These include significant intermetallic particles coarsening at grain boundaries in the HAZ and nugget regions, hence leading to poor fracture properties eventually associated to only modest bulk hardness values. Furthermore, non-conservative Glinkas predictions given in Fig. 15 endorse the assumption of brittle nugget microstructure. Inasmuch as the R method takes into account the fatigue properties of the parent material, it is possible, in the final fracture region (i.e. K MAX K C ), that the low-initiation fracture toughness (K C value) of the nugget microstructure enhances the FCGR by strengthening monotonic tensile modes of fracture, resulting in somewhat underestimated FCGR predictions. This was verified as well during longitudinal cracking analysis, and it seems again that Glinkas model is not able to cope with this particular condition. At low applied K values, however, a possible microstructural effect is discarded because the crack tip is still in the parent material. Thus, in this case, it is more likely that residual stress relaxation is taking place due to the cyclic load application, which would result in non-conservative predictions. The R method is only valid if linear elastic conditions are maintained during the test, that is, if crack tip plasticity is small. If this condition is broken, the obtained K rx profile is no longer the best choice to carry out the prediction. Moreover, it deserves to be emphasized that FCGR prediction is based on initial K Ir profile, which presents high negative values as shown in Fig. 7a. In order to verify whether there was plastic relaxation or not, both K rx and corresponding rx profiles were again derived from the broken halves of a fatigued test piece. Results exhibited in Fig. 16 indicate that the material indeed suffered significant stress relaxation by plastic flow. This is in line with recent findings from Liljedahl et al.,16 in which good prediction of residual stress redistribution was achieved for growing cracks at constant stress

Fig. 14 Vickers hardness profile transversal to the friction stir weld line.

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

536

M . T . M I L A N et al.

inasmuch as the relaxation takes place progressively as loading cycles are applied to the test samples. This can be confirmed in Fig. 16b, when a single load cycle is enough to cause a detectable stress relaxation and hence a rearranged residual stress profile. Last but not least, it should be noted that the transverse crack growth case illustrates very well the limitation of the presently adopted approach in identifying the main controlling factor of FCGR and drawing conclusive facts about it. Aimed at separating the effects of the magnitude of the residual stresses and the weld microstructure on fatigue behaviour of the FSW plates, FCGR experiments employing pre-strained test pieces are being carefully planned. FCGR testing of these stressrelieved specimens certainly will clarify current uncertainties and allow the role of the secondary stresses to be fully understood.

CONCLUDING REMARKS
Fig. 15 FCGR predictions for the welded joint compared to experimental values, with crack growing transverse to the weld line.

intensity range, a scenario farther from situations in practice than that studied in this work (i.e. increasing- K testing). The maximum nominal stress initially applied to the test piece was approximately 80 MPa. By adding this value to the peak tensile residual stress observed on a position corresponding to the advancing side of the rotating tool (Fig. 7b), an effective tensile stress approaching 300 MPa is obtained (Fig. 16b). Such stress level is certainly high enough to cause plastic deformation as the elastic limit of the material is around 250 MPa. Plastic deformation reduces both the peak tensile residual stress and the compressive residual stress near the edges of the test piece because a zero-stress global balance must be maintained. Thus, the initial K rx values calculated considering the non-relaxed residual stress profile are likely to be overestimated, resulting in non-conservative prediction of FCGR. A new FCGR prediction of the welded joint based on the relaxed K rx profile is given in Fig. 17. Results indicate that even considering the residual stress relaxation, forecast values for the final fracture region within the weld region are still non-conservative, so confirming that monotonic fracture modes do prevail at higher K values. However, for K values below 17 MPa. m it is now observed that predicted values are invariably higher than experimental ones, evidencing that the relaxed K rx profile is now underestimated. Therefore, the real K rx profile during fatigue testing must range from that obtained with the intact test pieces to that derived from the test specimens halves,

Longitudinal and transverse fatigue crack growth properties of friction stir welded joints of thin rolled plates of 2024-T3 aluminium alloy have been studied under constant load amplitude conditions (increasing- K ). Corresponding residual stress intensity factors in mode I (K Ir ) profiles were determined using a fracture mechanics approach, and the inverse weight function method was utilized to obtain the residual stress profiles. Fatigue properties of the parent metal were employed alongside K Ir profiles and Glinkas R method in order to predict FCGR of the welded joints. Main conclusions that may apply to the weld region, but not to necessarily a welded component as a whole, are as follows:
1. Welded joints are more resistant to longitudinally growing fatigue cracks than the parent metal at threshold K values probably due to beneficial thermal residual stresses arresting crack propagation, but less resistant next to the final instability point probably owing to monotonic tensile fracture modes intercepting subcritical crack growth in weld microstructures. Consequently, FCGR predictions trend to be conservative at lower propagation rates and non-conservative for faster cracks; 2. Intense compressive residual stresses render welded plates more fatigue resistant to transverse cracks than unmixed parent plate. Nonetheless, once the crack tip enters the brittle weld region FCGR accelerates due to operative monotonic tensile modes of fracture, leading to non-conservative crack growth rate predictions next to K C instability. At threshold K values non-conservative predictions values result from residual

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

FATIGUE BEHAVIOUR OF FRICTION STIR WELDED AA2024-T3 ALLOY

537

Fig. 16 Comparison between K rx profiles obtained from intact (TPs 1 e 2) and already fatigued testpieces (TPs 3 e 4).

stress relaxation. Improvements on predicted growth rate values are strongly dependent on how the progressive plastic relaxation of the residual stress field is considered; 3. With regard to design approaches broadly utilized in the aviation industry, the following findings should be noted: (i) If a damage-tolerant approach (stress intensity factor, K ) is applied, a growing crack is likely to be less detrimental when advancing towards, rather than along the weldline; (ii) If a safe-life approach (engineering stress, S) is employed, a crack is more likely to initiate transversally to the weldline in a pristine structure, rather than longitudinally to it;

4. Only complementary FCGR testing of pre-strained (i.e. stress-relieved) test pieces will permit the full understanding of the role of the secondary stresses on the fatigue behaviour of longitudinal and transverse FSW joints.

Acknowledgements M. T. MILAN would like to thank Fapesp (Fundac a o para o Amparo a ao Paulo, Brazil) for ` Pesquisa do Estado de S Grants 02/09027-4 and 03/11059-4. Thanks are also to Embraer S/A (Brazil) for providing the materials tested in this work.

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

538

M . T . M I L A N et al.

10

11

Fig. 17 FCGR predictions for the welded joint (crack growth transverse to the weld line) compared to the experimental values, considering the K rx profiles obtained, respectively, from intact testpieces and broken halves of fatigue specimens.

12

13

REFERENCES
1 Itoh, Y. Z., Suruga, S. and Kashiwaya, H. (1989) Prediction of fatigue crack-growth rate in welding residual-stress field. Engng. Fract. Mech. 33, 397407. Bussu, G. and Irving, P. E. (2003) The role of residual stress and affected zone properties on fatigue crack propagation in friction stir welded 2024-T351 aluminium joints. Intl. Jnl. Fatigue 25, 7788. Milan, M. T. and Bowen, P. (2002) Effects of particle size, particle volume fraction and matrix composition on the fatigue crack growth resistance of Al alloy / Al alloy + SiC bimaterials. Proceedings of the Institute of Mechanical Engineering L. Jnl Mater. Des. Appl. 216, 245255. Milan, M. T. and Bowen, P. (2003) Experimental and predicted fatigue crack growth resistance in Al2124/Al2124 + 35%SiC biomaterial. Intl. Jnl. Fatigue 25, 649659. 14

15

16

Parker, A. P. (1984) An overview of the mechanics of fracture and fatigue in the presence of residual stress. J. Mech. Work. Tech. 10, 165174. Milan, M. T. and Bowen, P. (2004a) Fracture toughness of selectively reinforced Al2124 alloy: precrack tip in the composite side. Metal. Mater. Trans. A 35A, 13931401. Milan, M. T. and Bowen, P. (2004b) Fracture toughness of selectively reinforced Al2124 alloy: precrack tip in the aluminium alloy side. Mater. Sci. Tech. 20, 783789. Glinka, G. (1979) Effect of Residual Stresses on Fatigue Crack Growth in Steel Weldments under Constant and Variable Amplitude Loads, ASTM Special Technical Publication (STP) 677, American Society for Testing and Materials, Philadelphia, PA, USA,2 pp. 198214. Schindler, H. J. (1996) Determination of residual stress distributions from measured stress intensity factors. Intl. Jnl. Fract. 74, R23R30. Prime, M. (1999) Residual stress measurement by successive extension of a slot: the crack compliance method. Appl. Mech. Rev. 52, 7596. Schindler, H. J. and Bertschinger, P. (1997) Some step towards automation of the crack compliance method to measure residual stress distributions. In: Proceedings of the 5th International Conference on Residual Stresses ICRS-5. (Edited by T. Ericsson, M. Oden and A. Andersson), Linkoping, Sweden, Vol. 1. Fett, T. and Munz, D. (1997) Stress Intensity Factors and Weight Functions, Computational Mechanics Publications, Southampton, UK. Milan, M. T., Bose Filho, W. W., Malafaia, A. M. S., Silva, C. P. O. and Pellizer, B. C. (2006) Slot machining effects on residual stress measurements using the crack compliance method. Jnl. Test. Eval. 34, 149152. Khaled, T. (2005) An Outsider Looks at Friction Stir Welding. Report #: ANM-112N-05-06. Federal Aviation Administration, Lakewood, USA. Kamp, N., Sullivan, A., Tomasi, R. and Robson, J. D. (2006) Modelling of heterogeneous precipitate evolution during friction stir welding process. Acta Mater. 54, 20032014. Liljedahl C. D. M., Brouard, J., Zanellato, O., Lin, J., Tan, J. F., Ganguly, S., Irving, P. E., Fitzpatrick, M. E., Zhang, X. and Edwards, L. (2007) Weld residual stress effects on fatigue crack growth behaviour of aluminium alloy 2024-T3. In: Proceedings of the First International Conference on Damage Tolerance of Aircraft Structures. (Edited by R. Benedictus, J. Schijve, R. C. Alderliesten and J. J. Homan), Delft, The Netherlands.

c 2008 The Authors. Journal Compilation c 2008 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct . 31, 526538

Вам также может понравиться