You are on page 1of 7

Materials Science and Engineering A 527 (2010) 59625968

Contents lists available at ScienceDirect

Materials Science and Engineering A


journal homepage: www.elsevier.com/locate/msea

Prediction of fatigue crack growth and residual stress relaxations in shot-peened material
Jinxiang Liu a, , Huang Yuan b , Ridong Liao a
a b

School of Mechanical Engineering, Beijing Institute of Technology, Beijing 100081, China Department of Mechanical Engineering, University of Wuppertal, Wuppertal 42103, Germany

a r t i c l e

i n f o

a b s t r a c t
The fatigue crack growth, incorporating effects of shot peening, is investigated using two-dimensional nite element analysis. The cohesive zone model, which takes into account the damage evolution under cyclic load, is adopted in the nite element analysis to simulate the potential fracture surfaces. The effects of shot peening on the fatigue crack growth and the relaxation of residual stress under cyclic load are studied. It is shown that the residual stress retards the fatigue crack propagation and the retardation effect depends not only on the shot peening intensity but also on the cyclic load amplitude. The residual stress relaxes nonlinearly as the load cycles increase. 2010 Elsevier B.V. All rights reserved.

Article history: Received 16 March 2010 Received in revised form 25 May 2010 Accepted 26 May 2010

Keywords: Shot peening Fatigue crack growth Cohesive zone model Finite element method

1. Introduction Fatigue life is an important specication in the design of components subject to constant or variable amplitude loads. The basic mechanism of a fatigue failure is that a slowly spreading crack extends with each cycle of the applied mechanical load. Practically tensile stress is responsible for nucleating and propagating the crack across a certain component, while a compressive stress will only close the crack and cause no damage [1]. One effective way to extend the fatigue life of components is to reduce or eliminate the tensile stresses by inducing a constant compressive stress in the surface of the components. Shot peening is widely used as a method to create such compressive stress eld in the surface layers of machine parts [2]. Shot impacts result in local plastic deformations on the surface. Since the plastically stretched surface layer tends to expand and the adjacent elastically responding material in vicinity and below the impact location restrains the expansion, a compressive residual stress eld is formed in the near surface layers [3]. Compressive residual stress close to the surface of material may prolong fatigue life by decelerating the initiation and growth phases of the fatigue process [47]. In damage tolerant design the durability of the fatigue crack growth is a major concern. The process of shot peening affects the initiation of small cracks and the growth of cracks near surfaces. Because the compressive residual stress layer is very thin it is very

challenging to experimentally quantify the interaction between the residual stress eld and the fatigue crack development. Even in conventional FEM computations the effects of the residual stresses cannot be considered. On the other hand, the quantitative relation between the residual stress state and the crack growth is of great signicance in component design optimization. In this paper, the cohesive zone model (CZM) is used to simulate the crack growth of shot-peened specimens. The parameters of CZM are identied using Paris law in common cracked specimens and are directly applied to 2D shot-peened specimens. Fatigue crack growth and residual stress relaxation of shot-peened specimens have been studied for different load amplitudes and different shot peening intensities. The main purpose of the study is to examine the role of residual stress on fatigue crack initiation and growth. 2. Cohesive zone model and parameter identication During crack propagation, a fracture process zone exists in front of the crack tip where micro-voids initiate, grow, and nally coalesce with the main crack. In the fracture process zone, material degradation is obvious. The conventional elasticplastic constitutive relation used in bulk material is not suitable for the local material degraded region. To consider the material degradation in the fracture process zone, one simple way is to use the cohesive zone model. The fundamental idea behind the cohesive zone model can be traced back to the strip yield models of Dugdale [8] and Barenblatt [9]. In this way, the unrealistic continuum mechanics stress singularity at the crack tip could be avoided. Contrary to the

Corresponding author. Fax: +86 10 68913041. E-mail address: liujx@bit.edu.cn (J. Liu). 0921-5093/$ see front matter 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.msea.2010.05.080

J. Liu et al. / Materials Science and Engineering A 527 (2010) 59625968

5963

standard concept in continuum mechanics, the cohesive zone model allows for a displacement jump inside material and for the separation of material surfaces which nally leads to the failure of the material. No continuum elements are involved in the cohesive zone. Cohesive interfaces or elements are dened between the continuum elements instead, which open when damage occurs and loose their stiffness at failure so that the continuum elements are disconnected. In a cohesive zone model, the fracture process zone is treated with its own constitutive relationship in form of the tractionseparation law (TSL). Since the cohesive zone model is a phenomenological model there is no direct evidence for formulation of the tractionseparation law. The fracture behavior of a cracked specimen is mainly determined by the energy release rate and the effect of the TSL form is secondary. We take the popular law suggested by Needleman [10] as Tn =
max e

exp

Fig. 1. The tractionseparation relationship for the cohesive zone model under simple and cyclic load condition.

(1) present investigation follows a linear relationship with a slope equal to that of the current tractionseparation curve at zero separation. The value of current unloading/reloading stiffness is given by kn =
max e

where Tn and are the normal traction and normal separation, respectively. The model parameters are the cohesive strength, max , and the cohesive length, 0 , i.e. where the traction reaches a peak value, max . Under monotonic load conditions max and 0 are supposed to be material parameters. The area under the Tn curve is called the separation energy, , which is dened as

(5)

=
0

Tn d =

max 0 e.

(2)

The cohesive zone model has been extensively used for analysis of fracture under monotonic load cases [1014]. The results seem to conrm the cohesive zone model, even though the present applications of cohesive models are still far away from practical engineering employment in structural integrity assessments [15]. The application of cohesive zone model for fatigue cracking is, however, still at initial stage. The general treatments of the cohesive zone model for fatigue analysis is to introduce additional parameters to describe the degradation of material. The main difculty is constructing a realistic damage evolution equation for the damage variable. In the model suggested by Siegmund et al. [16,17] a damage variable, D, is introduced for cyclic load and is described by D= | | Tn
max

f max,0

H(C 0 ),

(3)

where D is the damage increment, is the separation increment, C is the accumulated separation, H stands for Heaviside function. Two additional parameters are introduced in the damage evolution Eq. (3): the cohesive endurance limit, f , and the accumulated cohesive length, . Because the material damage is irreversible, the damage increment should not be less than zero. The incremental damage, D, is proportional to the normalized incremental separation, | |/ , and weighted by a measure of current traction reduced by the endurance limit. The current value of the damage, D, is then used to calculate the current cohesive strength, max , as
max

= (1 D)

max,0 .

(4)

In the above equation max,0 denotes the initial cohesive strength. Obviously, the cohesive strength decreases with the damage indicator, D, which represents the material degradation. The TSL, taking into account the damage accumulation under cyclic load, is given by substituting the current cohesive strength, max , into Eq. (1). The denition of unloading/reloading path is indispensable for cyclic cohesive zone model. The unloading/reloading path in the

The traction and separation relationships for monotonic and cyclic loads are depicted in Fig. 1. Since max decreases with cycling, kn is not a constant even within one loading step. The cohesive zone model presented above has been implemented into the commercial nite element code ABAQUS [18] via the interface UINTER. For the mode I problems the crack propagates along the symmetric plane where the cohesive zone can be assumed. The cohesive model with the modication for fatigue is used in the computations for 2D crack growth. In the cyclic cohesive zone model, there are four parameters, i.e. the initial cohesive strength max,0 , the cohesive length 0 , the cohesive endurance limit f and the accumulated cohesive length . In general, the rst two parameters are determined by the fracture energy under monotonic load condition [13,14]. For simplication, in this study they are assumed to be material parameters and constant through the whole specimen. The small variations due to constraint effects are neglected [19]. The void growth and nucleation process, which is the physical phenomenon behind the cohesive zone model, in the specimens under nominal plane strain condition, a rather highly constrained state, are generally characterized by values of max,0 between 3 and 4 times of material yield stress y [20]. For thin sheet, a plane stress case, the cohesive strength max,0 at the crack tip is approximately 2 y [21,13]. Based on previous studies, the value of the cohesive strength max,0 is taken to be 3 y for all computations. For the crack initiation under small scale yielding, the mode I cohesive energy can be identied with the value of the J-integral 2 at crack initiation, = JIC . With JIC = (1 2 )KIC /E and Eq. (2), the value of cohesive length 0 is determined. For Inconel 718 alloy [22], taking the crack initiation value of KIC = 120 MPa m and yield stress value of y = 710 MPa under room temperature, the cohesive strength max,0 and cohesive length 0 are 2130 MPa and 0.0055 mm, respectively. To understand the interactions between crack growth and f as well as , we have conducted several 2D simulations on compact tension specimens, which is in the square geometry of size 250 mm with an initial crack of 100 mm. For the mode I problem, the crack is constrained to grow along the symmetrical plane of the specimen. Due to symmetry, only half of the specimen has to be modeled. The constitutive relation governing the deformation behavior of

5964

J. Liu et al. / Materials Science and Engineering A 527 (2010) 59625968

Fig. 2. The crack growth versus load cycles under different load conditions K, i.e. 20.6 MPa m, 24.8 MPa m, 28.9 MPa m and 41.3 MPa m. The cohesive endurance limit f and accumulated cohesive length are 0.25 max,0 and 40 , respectively.

Fig. 4. Residual stress ( yy and zz , see Fig. 5) distribution in specimens peened by the Almens intensity 0.12 and 0.25 mmA, respectively. The experimental data are taken from the published works [24,25].

the bulk material is the elasticplastic constitutive relation using the Mises yield criterion. Crack growth in the 2D specimen versus load cycles is shown in Fig. 2 under stress controlled loading condition with loading ratio R = 0. The stress intensity factor range K is used to serve as an indicator of load intensity. In present work the stress intensity factor, K, JE/(1 2 ). It should be is calculated from the J-integral by K = carefully noted that as the amplitude of the load is held constant, K gradually creeps up as a result of the steady increase in crack length. However, when crack growth is relatively small compared with the initial crack length, the change of K is negligible. For all load cases, the crack growth speed increases after crack initiation and becomes stationary rapidly. Due to mesh limits, we cannot simulate larger crack growth amount. From the gures we obtain two features which have been conrmed in experiments: (1) The crack initiation is sensitive to the load amplitudes. The initiation time decreases nonproportionally as load amplitude increases. (2) After initiation the crack grows linearly with the load cycles, i.e. the crack growth rate becomes constant for small amount of crack growth. Based on the above described parameter study and experimental data, appropriate accumulated cohesive length and cohesive endurance limit f can be determined. Fig. 3 plots experimental fatigue data [23] and numerical prediction together. For tested material, the heat treatment is water quenching from 960 C followed by tempering in 718 C for 8 h. Crack propagation test was done in room temperature and stress ratio is 0. Here, equals 300

and f equals 0.25 max,0 . The predicted line is in a good agreement with experimental data and has approximately the same slope. This comparison demonstrates that the cohesive zone model is capable of matching long-crack constant amplitude fatigue data at least as well as Paris law.

3. 2D simulations of fatigue crack propagation In shot peening life assessment, inuence of the specimen surface damage will not be considered explicitly. This means, the fatigue prediction can be performed just based on the stress and strain states. The fatigue life variations are induced by the local residual stress. In the study, no other effects of shot peening are considered, as in most other works. With the help of ABAQUS subroutine SIGINI [18], two intensities of shot peening, 0.12 and 0.25 mmA, are used. The experimental data have been taken from the papers by Hessert et al. [24] for 0.12 mmA and by Hoffmeister et al. [25] for 0.25 mmA. The residual stress of 0.12 mmA is not generated by shot peening but by ultrasonic peening. Since no other effects but only residual stress is considered in the study, ultrasonic peening, in a sense, has no difference with shot peening. The residual stresses of two shot peening intensities are shown in Fig. 4. The main differences between the two residual stress proles are the depth of compressive zone and the maximum compressive stress. The stronger peening intensity leads to the greater depth of compressive zone and the greater value of maximum compressive stress. For comparison, two cyclic load amplitudes are considered for each shot peening intensity. One half of a square plate with a single-edge-crack is modeled utilizing its geometrical symmetry. The two-dimensional FE model is shown in Fig. 5. The bottom plot shows the initial mesh and the top plot zooms in to show the tip region. The plane strain element with 4 nodes is used for continuum elements. Since the surface layer affected by the shot peening process is very thin and the stress gradient in this layer is very high, extremely ne elements are needed for not only the crack area but also the whole affected layer. Therefore, only a very small specimen can be modeled in order to keep the total number of nodes within acceptable bounds. The specimen is 2.5 mm long and 1.25 mm wide. As shown in the gure, the elements in the shot peening affected layer are 0.004 mm in the depth direction (x direction). The initial crack is 0.012 mm, i.e. three elements, and the potential fracture surface is assumed to be 0.6 mm. The symmetric line is y = 0, and the initial crack is from x = 0 on the symmetric line. The boundary and loading conditions are as in the preceding section for all of following computations.

Fig. 3. Comparison of experimental [23] and numerical crack growth rates. The cohesive endurance limit f and accumulated cohesive length are 0.25 max,0 and 300 , respectively.

J. Liu et al. / Materials Science and Engineering A 527 (2010) 59625968

5965

Fig. 8. The crack growth with the number of load cycles under different loading forces. Shot peening intensity is 0.12 mmA.

Fig. 5. The two-dimensional FEM model with initial crack length of 0.012 mm (upper: near-tip detail, lower: overall view).

Fig. 6. The crack growth with the number of load cycles under different loading forces for unpeened specimen. The crack tip position is measured from free surface.

Fatigue crack growth for unpeened specimen is also studied. Crack growth vs. load cycles for specimen without shot peening effect is shown in Fig. 6. For both load cases, the crack growth speed increases after crack initiation and becomes steady rapidly. The numerically predicted crack growth for shot-peened material reveals that fatigue crack growth may not only start from the pre-existing initial crack tip as in the case of unpeened specimen, but also start from the position deeper than the initial crack tip, as shown in Fig. 7. The conventional denition of the crack length

Fig. 7. Scheme showing the fatigue crack growth for shot-peened material.

measured from initial crack tip is no longer suitable in the shot peening case. Here, the crack tip position measured from free surface is used for describing the fatigue crack growth of shot-peened material. Crack growth vs. load cycles in a specimen with 0.12 mmA shot peening is depicted in Fig. 8. K, based on calculation of specimen without shot peening but under the same load, is used as an indicator of the loading intensity for shot peening case. Comparing the gures, it can be found, besides that shot peening greatly retards the crack initiation for both load intensities, the location where the crack growth starts is also changed by shot peening. For the 0.12 mmA shot peening intensity, the crack growth starts at the position that is 0.15 mm away from the initial pre-existing crack tip. This position is where the initial residual stress reaches its maximum positive value while the initial existing crack is within compressive stress zone. Once a crack initiates, it grows in two directions, i.e. toward the center of the specimen and toward the surface of the specimen. The crack growth speed toward the center is a little faster than that in the unpeened specimen. However, the crack growth speed toward the surface is much slower than that in the unpeened specimen because this area has the strong compressive residual stress. After a certain load cycles, besides the two newly formed crack tips, the crack also begins to grow from the initial crack tip. Thus, in specimen the crack growth occurs on three crack tips simultaneously. The crack growth in the compressive stress zone is very slow. The last cracking position in this zone is where the residual stress reaches its maximum negative value. The results indicate that the tensile residual stress can accelerate crack growth. Thus, the overall effect of shot peening seems to accelerate crack growth in two-dimensional case once the crack begins to grow, even though the compressive residual stress will decelerate crack growth. For the two load intensities shown in the gure, the characteristics of the crack initiation and growth are the same. Fig. 9 presents crack growth vs. load cycles in the specimen with the 0.25 mmA shot peening. The 0.25 mmA shot peening delays the crack initiation more than the 0.12 mmA shot peening does. Compared with the 0.12 mmA shot peening, the position where crack growth starts is not 0.15 mm but 0.3 mm away from initial pre-existing crack tip, which corresponds to the maximum tensile stress area of the 0.25 mmA shot peening. The reason for the differences of crack growth is that the 0.25 mmA shot peening creates a deeper compressive zone and larger maximum compressive residual stress. Except for above mentioned differences, the shapes of the curves are basically the same for both shot peening intensities. In general, introduction of residual stress through shot peening is intended to extend fatigue life via the delay of crack initiation. Fig. 10 summarizes the effect of shot peening on the delay of crack

5966

J. Liu et al. / Materials Science and Engineering A 527 (2010) 59625968

Fig. 9. The crack growth with the number of load cycles under different loading forces. Shot peening intensity is 0.25 mmA.

Fig. 11. Redistribution of residual stress under cyclic load before crack initiation. Shot peening intensity is 0.12 mmA, and load is K = 34.3 MPa m.

initiation, although the positions of crack initiation can be different. For all cases, i.e. unpeened, 0.12 and 0.25 mmA shot peening, with increasing load intensity, the number of load cycles to crack initiation decreases nonlinearly. Shot peening increases the number of load cycles to crack initiation when specimens are exposed to the same load intensities. The delay of crack initiation caused by shot peening is much longer at lower load than that at higher load. In addition, higher shot peening intensity results in a larger number of load cycles for crack initiation. 4. Relaxation of residual stress The residual stress, caused by shot peening, can be relaxed by cyclic loads. For specimens without initial cracks, the stress relaxation has been extensively studied. For specimens with initial cracks, the interaction between a growing fatigue crack and residual stress eld is not well understood. The lack of understanding has sometimes led to confusions and possibly inaccurate methods being employed in fatigue life prediction. In the study of the two-dimensional case, the initial crack is in the area with initial compressive stress and far away from the initial tensile stress. Fig. 11 shows the relaxation of residual stress before the crack starts to propagate. The computational condition is that the shot peening intensity is 0.12 mmA and the load intensity is 34.3 MPa m. As load cycle increases, the relaxation of residual stress is increased even without crack growth. At the same time, the cyclic load introduces a little compressive stress inside the specimen in areas where they previously do not exist. The stress

relaxation is not linear with load cycles. In the gure, the increment of loading cycles for every curve is 100 load cycles. It can be seen that the stress relaxation in the rst 100 load cycles is much larger than that in the following cycles until 400 load cycles. In fact, the stress relaxation in this stage is mainly caused by the redistribution of the residual stress under the rst load cycle. After the rst loading cycle, the residual stress is relatively stable due to relatively small damage accumulation and small change in stiffness of the cohesive zone until a certain load cycles, here, 400 load cycles. Beyond 400 load cycles, the damage accumulation becomes larger and thus results in the decreases of stiffness of the cohesive zone, and loadbearing capacity. Consequently, a large relaxation of residual stress can be observed after 400 load cycles. Because the damage accumulation is in an accelerated relation with the increasing of load cycles, the relaxation of residual stress becomes faster and faster. With load cycle increasing, the separation of potential fracture surfaces, i.e. cohesive zone, is increasing once damage begins to accumulate. The increasing separation in fact weakens the crack closure effect in the initial crack area caused by the compressive residual stress. When the crack closure effect nally vanishes, the residual stress in initial crack area disappears totally. Under the same computational condition, the relaxation of residual stress after the crack growing is given in Fig. 12. It can be seen in the gure that during the crack growth the residual stress in newly cracked area vanishes eventually. Meanwhile, in order to keep in stress equilibrium, the compressive residual stress increases in areas where crack propagation has not reached yet. After a certain load cycle, the compressive residual stress will stop increasing and begins to decrease due to weakened stiffness in its corresponding cohesive

Fig. 10. The effect of shot peening on the number of load cycles to start crack growth in 2D computations.

Fig. 12. Redistribution of residual stress under cyclic load after crack initiation. Shot peening intensity is 0.12 mmA, and load is K = 34.3 MPa m.

J. Liu et al. / Materials Science and Engineering A 527 (2010) 59625968

5967

Fig. 13. Redistribution of residual stress under cyclic load before crack initiation. Shot peening intensity is 0.12 mmA, and load is K = 37.7 MPa m.

Fig. 15. Redistribution of residual stress under cyclic load before crack initiation. Shot peening intensity is 0.25 mmA, and load is K = 34.3 MPa m.

zone. Once the crack growth is completed in the cohesive zone, the residual stress becomes zero on the whole fracture surface. Figs. 13 and 14 show the relaxation of residual stress before and after the crack growth. The shot peening intensity is the same as in Figs. 11 and 12, but the load intensity has a larger value of 37.7 MPa m. Comparing the results of the two load intensities, it is observed that higher load intensity leads to larger relaxation of residual stress for the same load cycles. Consequently, higher load intensity reduces the number of load cycles to eliminate the residual stress in the initial crack area. Except for this difference, the stress relaxations, before and after the crack growth, have the same trends for both load intensities. The relaxation of residual stress, for the studied case of 0.25 mmA shot peening intensity and 34.3 MPa m load amplitude, is plotted in Figs. 15 and 16, before and after the crack growth, respectively. It can be seen from the gures that, similar to the 0.12 mmA shot peening intensity, the stress relaxation in the rst 100 cycles is much larger than that in the following cycle increment before crack begins to propagate. Different from the 0.12 mmA shot peening intensity, however, the residual stress in the initial crack area does not release totally before crack begins to propagate for the 0.25 mmA shot peening intensity. The reason is that the crack closure effect of the residual stress for 0.25 mmA shot peening intensity is stronger than the 0.12 mmA shot peening intensity and can not be overcome under 34.3 MPa m load intensity. The stress relaxation behavior after crack begins to propagate, has the same characteristics for 0.25 and 0.12 mmA shot peening intensities. Figs. 17 and 18 show the relaxation of residual stress under the condition of the 0.25 mmA shot peening intensity and 37.7 MPa m

Fig. 16. Redistribution of residual stress under cyclic load after crack initiation. Shot peening intensity is 0.25 mmA, and load is K = 34.3 MPa m.

load amplitude before and after the crack growth, respectively. The trends of the stress relaxation are observed to be the same as the 0.12 mmA shot peening intensity, both before and after crack growth. The residual stress in the initial crack area releases completely before crack begins to grow. This means whether the residual stress will release totally in the initial crack area before crack growth depends not only on the shot peening intensity but also on the load amplitude. Furthermore, the residual stress releases to zero on newly formed surface for all cases, and this is independent of the shot peening intensity and loading ampli-

Fig. 14. Redistribution of residual stress under cyclic load after crack initiation. Shot peening intensity is 0.12 mmA, and load is K = 37.7 MPa m.

Fig. 17. Redistribution of residual stress under cyclic load before crack initiation. Shot peening intensity is 0.25 mmA, and load is K = 37.7 MPa m.

5968

J. Liu et al. / Materials Science and Engineering A 527 (2010) 59625968

not linear with load cycles. Larger residual stress relaxation can be observed before and when crack begins to propagate. Generally speaking, the residual stress relaxation is induced by material damage in the crack growth direction which is modeled by the cohesive zone model. Cyclic loads lead to decrease of cohesive zone stiffness. And when crack grows, restrain condition for residual stress is changed. These two factors result in redistribution and relaxation of residual stress. The relaxation of residual stress is a natural consequence of cohesive zone stiffness decrease and crack propagation. References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] H. Guechichi, L. Castex, J. Mater. Process. Technol. 172 (2006) 381387. M. Kobayashi, T. Matsui, Y. Murakami, Int. J. Fatigue 20 (1998) 351357. W.Z. Zhuang, G.R. Halford, Int. J. Fatigue 23 (2001) s31s37. E.R. de los Rios, A. Walley, M.T. Milan, G. Hammersley, Int. J. Fatigue 17 (1995) 493499. A. Turnbull, E.R. de los Rios, R.B. Tait, C. Laurant, J.S. Boabaid, Fatigue Fract. Eng. Mater. Struct. 21 (1998) 15131524. P.S. Song, C.C. Wen, Eng. Fract. Mech. 63 (1999) 295304. V. Sabelkin, S.A. Martinez, S. Mall, S. Sathish, M.P. Blodgett, Fatigue Fract. Eng. Mater. Struct. 28 (2005) 321332. D.S. Dugdale, J. Mech. Phys. Solids 8 (1960) 100104. G. Barenblatt, Adv. Appl. Mech. 7 (1962) 55129. A. Needleman, Int. J. Fracture 42 (1990) 2140. A. Needleman, J. Appl. Mech.-Trans. ASME 54 (1987) 525531. X.P. Xu, A. Needleman, J. Mech. Phys. Solids 42 (1994) 13971434. H. Yuan, G. Lin, A. Cornec, J. Eng. Mater. Technol.-Trans. ASME 118 (1996) 192200. I. Scheider, M. Schdel, W. Brocks, W. Schnfeld, Eng. Fract. Mech. 73 (2006) 252263. W. Brocks, A. Cornec, Comprehensive Structural Integrity, vol. 3, Elsevier, 2003, pp. 127209. K. Roe, T. Siegmund, Eng. Fract. Mech. 70 (2003) 209232. T. Siegmund, Int. J. Fatigue 26 (2004) 929939. ABAQUS Users Manual. Version 6.7, ABAQUS, Inc., 2006. C. Chen, O. Kolednik, I. Scheider, T. Siegmund, A. Tatschl, F.D. Fischer, Int. J. Fracture 120 (2003) 517536. V. Tvergaard, J. Hutchinson, J. Mech. Phys. Solids 40 (1992) 13771397. J. Newman, J. Crews, C. Bigelow, D. Dawicke, ASTM STP 1244 (1995) 2142. W.D. Klopp, Aerospace Structural Metals Handbook: Nickel Base Alloys, Purdue Research Foundation, West Lafayette, IN, 1995. D. Skinn, J. Gallagher, A. Berens, P. Huber, J. Smith, Damage Tolerant Design Handbook, 1994. R. Hessert, J. Bamberg, W. Satzger, T. Taxer, The MTU Report, Ultrasonic impact treatment for surface hardening of the aero-engine material IN718, 2008. J. Hoffmeister, V. Schulze, A. Wanner, R. Hessert, G. Koenig, Proceedings of the 10TH International Conference on Shot Peening (ICSP10), Tokyo, 2008, pp. 157162.

Fig. 18. Redistribution of residual stress under cyclic load after crack initiation. Shot peening intensity is 0.25 mmA, and load is K = 37.7 MPa m.

tude. It should be noted that the material ahead of the crack tip is not described by conventional elasticplastic constitutive relation but tractionseparation equation of cohesive zone model. The response of material closely adjacent to cohesive zone is also strongly affected by the properties of cohesive zone. The results of residual stress relaxation are very sensitive to cohesive zone model. It is not easy to quantitatively give the inuence of plasticity on the residual stress state. 5. Conclusions The application of the cohesive zone model in the study provides a detailed understanding of the fatigue crack growth in shot-peened specimens. This model provides a reasonable crack initiation and propagation in 2D simulation. The current investigation predicts that the crack initiation position and time depend not only on shot peening intensity but also on cyclic load amplitude. The crack can be started from the position deeper than the initial crack tip for shot-peened specimen, as predicted by the computations. Retardation of fatigue crack growth is more remarkable for higher shot peening intensity. The relaxation of residual stress is increased with increasing load cycles, even without crack growing. The stress relaxation is

You might also like