You are on page 1of 50

Journal of Materials Engineering and Performance

A predictive methodology for high-cycle fatigue behaviour


of machined metallic parts.

Journal: Journal of Materials Engineering and Performance

Manuscript ID Draft

Manuscript Type: Technical Article

Date Submitted by the


n/a
Author:

Complete List of Authors: Laamouri, Adnen; ISSAT Sousse, Mechanical department


Sidhom, habib; ENSIT, mechanic
Braham, Chedly
Fo

Predictive methodology, HCF life, AISI 316L steel, surface integrity, FE


Keywords:
method, Dang Van’s criterion
rP
ee
rR
ev
iew
Page 1 of 49 Journal of Materials Engineering and Performance

1
2
3 A predictive methodology for high-cycle fatigue behaviour of machined metallic parts
4
5
Adnen Laamouri1,2, Habib Sidhom1, Chedly Braham3.
6
7 1 Laboratory of Mechanics, Materials and Processes (LMMP, LR99ES05), Higher National School of Engineers
8
9 of Tunis (ENSIT), University of Tunis, 5, Avenue Taha Husseïn, Montfleury, 1008 Tunis, Tunisia.
10
11 2 Higher Institute of Applied Science and Technology of Sousse (ISSATSO), University of Sousse, Avenue Ibn
12
13 Khaldoun, 4003 Sousse, Tunisia.
14
15 3 Laboratory of Processes and Engineering in Mechanics and Materials (PIMM, UMR CNRS 8006), ENSAM,
16
17 151, Boulevard de l’Hôpital Cedex 75013 Paris, France.
18
19  Adnen Laamouri, adnen_laamouri@yahoo.fr, Tel. +216 97 219 568
20
21
Fo

22
23 Abstract
24
This paper presents an efficient methodology for prediction of high cycle fatigue (HCF) behaviour and life of
rP

25
26
machined parts. It is based on the Dang Van’s fatigue criterion and takes into account the surface integrity effects
27
28
ee

including residual stresses, surface roughness, superficial microstructural changes and eventually damage defects.
29
30 The residual stress relaxation and the most critical stress micro-concentration acting at the potential crack
31
rR

32 nucleation site were analysed by the finite element (FE) method. The work-hardening and damaging effects are
33
34 taken into account through corrections of the fatigue threshold defined by the Dang Van’s criterion by
35
ev

36 phenomenological relationships. The predictive methodology was validated in the case of a 316L steel ground
37
38 surface. The experimental investigation by 4-point plane bending fatigue tests (R = 0.1) on notched specimens
iew

39
40 (Kt = 1.6) has showed a significant degradation of fatigue limit (-35%) for the ground state in comparison with
41
42 the electro-polished one (virgin sate). This degradation is due to the combination of three surface integrity effects:
43
44 (i) high tensile residual stresses partially relaxed, (ii) a surface roughness responsible of a multiple micro-cracks
45
46 nucleation in grinding grooves, and (iii) a significant superficial work-hardening. The proposed methodology has
47
48 permitted to predict the cyclic behaviour of both polished and ground surfaces, and to showed that the fatigue
49
50 strength degradation of the ground surface is mainly due to the effect of stabilized tensile residual stresses (IRS  -
51
52 30%), and the detrimental roughness effect appears less important (Irough  -12%). However, the beneficial work-
53
54 hardening effect seems to be relatively lower (Ihard  +5%).
55
56
57
58 Key words: Predictive methodology, HCF life, AISI 316L steel, surface integrity, FE method, Dang Van’s
59
60 criterion.

1
Journal of Materials Engineering and Performance Page 2 of 49

1
2
3 1. Introduction
4
5 Fatigue is the most common failure phenomenon for mechanical components and structures. It occurs when the
6
7 material is subjected to cyclic loading. This failure phenomenon can be occurred even for loadings under the yield
8
9 strength of the material. It has been recognized for many years that the fatigue behaviour of mechanical
10
11 components depends strongly on the surface properties resulting from manufacturing processes (machining
12
13 processes, superficial treatments, welding, etc.). Many studied [1-6] showed that metallurgical modifications (e.g.
14
15 phase transformation, plastic deformation, precipitation, etc.), compressive or tensile residual stress fields, surface
16
roughness and surface defects (e.g. micro-cracks, porosities, scales, etc.) are the principal surface properties (so-
17
18
called “surface integrity”) which influence the HCF behaviour. The relative importance of these properties and
19
20 their interaction was extensively discussed in the literature, leading to different conclusions as regards their
21
Fo

22 effectiveness on the HCF behaviour.


23
24 The machined surfaces by turning and milling, or finished by grinding (Figure 1) are particularly detrimental for
rP

25
26 fatigue strength of metallic material [7-14]. They are generally characterized by detrimental roughness, tensile
27
28 residual stress, metallurgical changes and sometimes damaging defects. So, it was shown that the micro-cracks
ee

29
30 nucleation is generally occurred at the surface and provoke excessive fatigue strength degradation.
31
rR

32 Residual stresses are the main surface integrity parameter which can affect strongly fatigue strength of machined
33
34 components [15-18]. If they are tensile stresses, they promote crack initiation and accelerate crack propagation .
35
ev

36 However, the machining stresses can be relaxed during cyclic loading until a stabilized state and their detrimental
37
38 effect will be reduced or sometimes cancelled [19]. In this case, the stabilized residual stresses should be evaluated
iew

39
and taken into account in the predictive calculations of fatigue strength.
40
41
The machined surface texture is generally characterized by parallel grooves which are more detrimental for fatigue
42
43 strength when they are oriented perpendicularly to the applied loading due to very local stress concentrations. This
44
45 accelerates the crack nucleation phase and degrades the fatigue strength, especially for notch-sensitive materials
46
47 [20-21]. Generally, hard materials such as high-carbon steels in quenched and tempered state are more sensitive
48
49 than ductile steels such as annealed low-carbon and austenitic stainless steels. The surface roughness effect is
50
51 usually considered as a stress micro-stress concentration effect. Different semi-empirical models based on
52
53 roughness parameters are proposed [22-25], but these models are notoriously inaccurate, leading to large safety
54
55 factors and overlay conservative prediction. A new approach providing more accurate prediction was presented
56
57 by Ås et al. [26-27], where FE analysis of the measured surface topography are carried out.
58
59 The microstructural changes of machined surfaces such as plastic strain, phase transformations, precipitations,
60

2
Page 3 of 49 Journal of Materials Engineering and Performance

1
2
3 grain elongation, etc., depend strongly on material properties and machining conditions. These changes lead to
4
5 superficial hardening in ductile materials such low-carbon steels [28], and on the contrary to superficial softening
6
7 in high strength materials [29-30]. Superficial hardening is beneficial for fatigue strength. It reduces the mobility
8
9 of dislocations and then retards the crack nucleation stage [23]. The full width at half maximum (FWHM) of X-
10
11 ray diffraction (XRD) peak can be used to assess qualitatively the density of defects in the material such as
12
13 dislocations arising from plastic deformation during grinding [31-32]. Deperrois et al. [33] proposed a theoretical
14
15 law relating the fatigue strength change with the FWHM in every depth of the superficial layer.
16
Several damaging defects such as thermal micro-cracks and burns could be produced when machining conditions
17
18
are not well controlled [34]. These defects are considered as potential sites of micro-cracks nucleation and
19
20 propagation, and they can lead to poor fatigue performance [35].
21
Fo

22 The contribution of each surface integrity effect on HCF behaviour is not yet well known quantitatively, in spite
23
24 of the development of several predictive models and approaches. Generally, the HCF behaviour of mechanical
rP

25
26 components can be evaluated using the local stress approach based on multiaxial fatigue criteria. Many works [4,
27
28 8, 18, 36-37] have used this approach and attempted to incorporate the effects of surface integrity for several
ee

29
30 manufacturing process (e.g. machining, surface treatments, welding, etc.). Benedetti et al. [36] postulated that
31
rR

32 fatigue endurance can be accurately predicted by a multiaxial fatigue criterion combined with a line method based
33
34 on the critical distance theory to account for the notch sensitivity, especially if stabilized residual stresses are taken
35
ev

36 into account. In previous work, Fathallah et al. [4] have developed and validated the local stress approach for HCF
37
38 behaviour of shot peened parts. The majority of surface integrity effects have been incorporated in multiaxial
iew

39
fatigue criteria such as the Crossland’s and Dang Van’s criteria. This permitted to predict the enhancement or
40
41
sometimes the degradation of fatigue strength, and also the potential crack initiation site in surface layers of shot
42
43 peened parts. This predictive approach has been validated in the case of peened T-welded joints [5] and also for
44
45 shot-peening and nitriding treatments [37].
46
47 On the other hand, there were many attempts to develop a HCF behaviour prediction methodology based on the
48
49 Finite Element simulation for peened surfaces [38 - 39]. Three steps are considered: (i) FE simulation of the initial
50
51 surface properties induced by the shot peening process, (ii) FE simulation of surface properties evolution under
52
53 cyclic loading. (iii) Prediction of both favorable and unfavorable shot peening effects on the HCF behaviour using
54
55 a multiaxial fatigue criterion. This computation methodology is carried out and validated for some materials.
56
57 In this study, we attempt to develop a methodology for fatigue lifetime, particularly applicable for machined parts,
58
59 based on the multiaxial fatigue criteria such as the Dang Van’s criterion and taking into account the surface
60

3
Journal of Materials Engineering and Performance Page 4 of 49

1
2
3 integrity effects. This methodology has the advantages to predict the cyclic behaviour of the machined surface,
4
5 and to evaluated the impacts of each surface property on HCF behaviour of material. Its validity will be studied in
6
7 the case of a AISI 316L steel ground surface. The experimental investigation was carried out on both ground and
8
9 electro-polished samples by 4-point bending fatigue tests and using surface characterization techniques before and
10
11 after fatigue tests (XRD, profilometry, micro-hardness, SEM). The prediction of the surface effects (including
12
13 stabilized residual stresses, surface roughness and superficial work-hardening and eventually damage state) was
14
15 made through the respective indicators IRS, Irough Ihard, and Idamage introduced in this work.
16
2. Predictive methodology for HCF life
17
18
2.1. Methodology flowchart
19
20 In this work, a methodology is developed to predict the HCF behaviour and life of machined materials. This
21
Fo

22 methodology is based on multiaxial fatigue criteria and takes into account the effects of surface integrity including
23
24 quasi-stabilized residual stresses, superficial hardening, surface roughness and eventually damage state. Its first
rP

25
26 application consists to calibrate and validate the adopted fatigue criterion for a treated surface under a specific
27
28 surface condition. After that, it can be used to predict and optimize the machining conditions in terms of HCF
ee

29
30 behaviour and life of the part. The flowchart of this methodology is composed of four principal steps (Figure 2):
31
rR

32  Step 1: input parameters


33
34 The principal input parameters of the methodology are defined as follows: (i) fatigue specimen geometry and
35
ev

36 dimensions, (ii) base material, (iii) machining conditions, (iv) cyclic bending (initial amplitude value 0a and stress
37
38 ratio R), and (v) a given fatigue life (104 ≤ N ≤ Nlimit).
iew

39
40  Step 2: experimental processor
41
42 This step consists to:
43
44 (i) Identify the monotonic and cyclic behaviour models of the base material, as well as two bending S-N curves
45
46 under different loading ratios (R1 and R2) in the HCF regime (104 ≤ N ≤ 107 cycles),
47
48 (ii) Characterize the initial surface integrity induced by the surface finishing process (in-depth residual stress
49
profiles, in-depth micro-hardness profile, roughness profile and damage defect nature and distribution).
50
51
This characterization can be, respectively, carried out by using the techniques of X-ray diffraction, micro-
52
53 hardness measurements, roughness profilometry and SEM examinations. It should be noted that these
54
55 surface properties could be predicted by numerical simulation of the surface machining process,
56
57 (iii) Determine the SN curve in the HCF regime for the machined specimen, in the same fatigue test conditions
58
59 as the base material (R1 or R2 loading ratio).
60

4
Page 5 of 49 Journal of Materials Engineering and Performance

1
2
3  Step 3: numerical processor
4
5 Three FE simulations are conducts in this step:
6
7 (i) FE simulation of the applied cyclic stress tensor acing at the critical zone of the polished sample. This
8
9 permits to identify the base material constants (0(N) et ) of the Dang Van’s criterion.
10
11 (ii) FE simulation of the micro-stress concentration factor (Kt) due to a roughness critical defect, and then
12
13 estimation of the fatigue micro-stress concentration factor (Kf) using a semi-empirical relationship;
14
15 (iii) FE simulation of the cyclic behaviour at the critical zone of the grinding finished sample, and determination
16
17 of the quasi-stabilized of cyclic and residual stresses.
18
19 After that, the HCF fatigue behaviour prediction process is conducted. The applied cyclic loading (a) is increased
20
21 progressively by an increment (a). For each increment, the Dang Van criterion is applied using the quasi-
Fo

22
23 stabilized cyclic stresses until minimization of the difference between the equivalent stress (  eq ( N ) ) and the
24
rP

25 criterion threshold (  s  N  ) under  = 0.1.


26
27
 Step 4: output results
28
ee

29
Two output results are given by the predictive methodology:
30
31
(i) Prediction of fatigue resistance in terms of limit loading (alim(N), R) for a given HFC life (N);
rR

32
33 (ii) Prediction of surface integrity effects by using four indicators proposed in this work, namely: IRS(N),
34
35
ev

Irough(N), Ihard(N) and Idamage(N).


36
37 2.2. Dang Van’s criterion
38
iew

39 The Dang Van’s muliaxial HCF criterion [40] is adopted in this work. It is based on the mesoscopic scale approach,
40
41 which takes into account the physical reality of the fatigue crack initiation phenomenon as a local phenomenon
42
43 that occurs in some grains that have undergone plastic slip, localized in the characteristic inter-crystalline bands.
44
45 This criterion relates to mesoscopic quantities which are calculated according to the macroscopic quantities by a
46
47 meso-macro passage. At the meso-scale, it is assumed that an elastic shakedown response at the grain scale itself
48
49 is a required condition to avoid a crack initiation after a high number of cycles (Figure 3). The basic parameters
50
51 are the maximum shear stress   n, t  on the most unfavourably oriented crystallographic plane and the hydrostatic
52
53 pressure p  t  . So, the fatigue resistance condition for a given HCF life (N) is traduced by the limitation of an
54
55
equivalent stress expressed by a linear relationship between two local stresses of the stabilized mesoscopic stress
56
57 cycle (Equation 1). This condition should be checked at the surface where fatigue micro-cracks nucleation is
58
59 always occurred.
60

5
Journal of Materials Engineering and Performance Page 6 of 49

1
2
3
 eq  N   max  max   n, t    0  N  . p  t     0  N  (1)
4 t  n 
5
6 The  0  N  and  0  N  constants define the multiaxial fatigue threshold of the base material for a given fatigue
7
8
life (N cycles).
9
10
11 The mesoscopic hydrostatic pressure p  t  is equal to macroscopic one P  t  at every instant t of the stabilized
12
13 cycle.
14
15 1
p  t   P  t   tr  t  (2)
16 3
17
18 The mesoscopic shear stress   n, t  is defined at every instant t of the stabilized cycle by the following
19
20
relationships:
21
Fo

22
  n, t   C  n, t   C m  n  (3)
23
24
Where C  n, t  represents the instantaneous macroscopic shear stress at a given material plane defined by its
rP

25
26
27 normal vector n , and where C m  n  is the mean macroscopic shear stress on the facet of normal n . In the case
28
ee

29
30 of complex loadings, C m  n  is defined by the center of the smallest circle containing the path of C  n, t  in the
31
rR

32 facet considered (Figure 4). In practice, it is necessary to scan all the facets in order to determine the critical plane
33
34 *
( n ) for which the mesoscopic shear stress is maximal.
35
ev

36 In the case of proportional loadings, the Dang Van’s criterion is simplified because the critical mesoscopic plane
37
38 corresponds to the critical macroscopic plane defined by Tresca, and it is expressed by the following equation.
iew

39
40 This formulation is admitted in the present study.
41
42  eq  N    a   0  N  .P max   0  N  (4)
43
44
2.3. Modelling of residual stress effect
45
46 The initial residual stresses are superposed to the applied cyclic stresses, and they may be relaxed during fatigue
47
48 lifetime if this superposition provokes cyclic plastic deformation of the material [15, 41]. This phenomenon is
49
50 beneficial for fatigue strength of machining surfaces, and can conducts to partial or complete relaxation of tensile
51
52 residual stresses [8, 19]. However, it is detrimental for compressive stresses induced by surface treatments such as
53
54 shot peening, deep rolling, laser peening, etc [1].
55
56 The main relaxation of residual stresses normally takes place at the first cycle, followed by further gradual
57
58 relaxation during fatigue life. The relaxation during the first cycles (quasi static loading) depends on the monotonic
59
60 yield strength of the material in tension and compression, while relaxation during successive cycles is related to

6
Page 7 of 49 Journal of Materials Engineering and Performance

1
2
3 the cyclic yield strength. They can reach a stabilized state when near fatigue limit of the part [15]. The residual
4
5 stress relaxation is a complex phenomenon which depends of the interaction of many factors such as the material
6
7 cyclic behaviour, initial state of surface layers, the applied stress amplitude, the loading mode, the number of
8
9 cycles, etc.
10
11 In order to take into account the residual stress effect in the Dang Van’s criterion, it is required to determine the
12
13 evolution of total cyclic stress state acting in the critical zone of the part. A FE modelling of the cyclic hardening
14
15 behaviour accounting for residual stress relaxation was developed in previous work and validated in the case of
16
AISI 316L steel ground surfaces [19]. It permits to simulate the cyclic hardening response (elastic-shakedown
17
18
response) of surface layers as well as the residual stress relaxation near fatigue limit of the material.
19
20
The first step of this simulation consists to introduce (by layer of 10 m) the initial residual stresses combined
21
Fo

22 with the corresponding initial work hardening as initial conditions by using two Fortran subroutines: Sigini and
23
24 Hardini, respectively (Figure 5).
rP

25
26 In the case of biaxial loadings (without shear stress), the total cyclic stresses acting at the surface where fatigue
27
28 cracks occur, is expressed by superposition of the applied and residual stresses.
ee

29
30  xx ( t )   *Rxx 0 0
31  
 t ( t )   app ( t ) +  R  
*
0  yy ( t )   *Ryy 0 (5)
rR

32
 0 0 0 
33 
34
35 Where  xx ( t ) and  yy ( t ) are normal stresses acting in the two surface directions.
ev

36
37 Once the total quasi-stabilized cyclic stresses are determined after the first cycles, it is possible to express the total
38
iew

39
equivalent stress  teq  appearing in the Dang Van’s criterion with the residual stress effect (Equation 6).
40
41
42  teq  N    app
a
  0  N  . Papp
max
 PR*  (6)
43
44
Where the hydrostatic pressure due to the quasi-stabilized residual stresses is given by:
45
46
1 *
47 PR* 
3
 Rxx   *Ryy  (7)
48
49
50 On the other hand, the applied equivalent stress  app
eq
 without this effect is expressed by:
51
52
53
 app
eq
 N    app
a
  0  N  .Papp
max
(8)
54
55 Therefore, the relaxed residual stress effect on HCF behaviour of the machined surface is estimated by the indicator
56
IRS which is expressed by the difference between the two equivalent stresses  app and  teq with respect to the
eq
57
58
59  0  N  constant (Equation 9).
60

7
Journal of Materials Engineering and Performance Page 8 of 49

1
2
3  app
eq
  teq
4 I RS  N    100 (9)
5
0  N 
6
7 2.4. Modelling of roughness effect
8
9 The roughness effect is generally described through the fatigue stress micro-concentration factor ( K̂ f ) which can
10
11
12 be deduced from the stress micro-concentration factor ( K̂ t ) [22]. To estimate K̂ t , Neuber [23] proposed a semi-
13
14 empirical relationship using the roughness parameter Rz and notch root radius (  ). A recent expression in terms
15
16 of dominant surface parameters (Ra, Ry and Rz) and notch root radius (  ) was developed by Arola [24] and
17
18 validated in the case of AISI 4130 CR steel surface machined by abrasive waterjet [25]. Ås et al. [26-27] proposed
19
20
to calculate K̂ t from FE simulations of the surface topography. The application of this method to a 6082.52-T6
21
Fo

22
23 aluminium alloy prepared with emery paper provided more accurate fatigue life prediction, since the stress
24
concentration estimation is based upon measurements rather than averaged geometrical parameters. According to
rP

25
26
27 Peterson [22], K̂ f is expressed in terms of K̂ t and the notch sensitivity q depending on the material and asperities
28
ee

29 geometry.
30
31

Kˆ f  1  q Kˆ t  1  (10)
rR

32
33
34 The notch sensitivity (q) can be defined in terms of the effective profile valley radius of the surface texture (  ).
35
ev

36 1
q (11)
37 1   /  
38
iew

39
Where  (in mm) is a material constant related to the ultimate strength (  u ) for steels by [25].
40
41
1.8
42  2070 
43   0.0254   ;  u  550 MPa  (12)
44  u 
45
46 We assume, in this work, that the ultimate strength for hardened surfaces (  us ) is proportional to the surface
47
48 Vickers micro-hardness ( HVs ). Then, the ultimate strength of the treated surface can be expressed by:
49
50 HVs
51  us   u0 . (13)
HV0
52
53
54 In many cases, the machining grooves are oriented in one direction (noted y-direction). Therefore, the micro-stress
55
56 concentration is characterized by the tensile fatigue factor K̂ fxx in the perpendicular direction (noted x-direction),
57
58
59 while the second tensile fatigue factor K̂ fyy is assumed negligeable. In this case, the micro-geometrical cyclic
60

8
Page 9 of 49 Journal of Materials Engineering and Performance

1
2
3 stress tensor due to the critical micro-notch effect, and the corresponding equivalent stress are given by Equations
4
5 14 and 15.
6
7  K̂ fxx . xx ( t ) 0 0
8  
(t )  
ˆ 0  yy ( t ) 0  (14)
9
 0 0 0 
10 
11
12 ˆ eq  N   ˆ a   0  N  .Pˆ max (15)
13
14
15    
Where ˆ a  Kˆ fxx . xxa / 2 and Pˆ max  Kˆ fxx . xxmax   yymax / 3 .
16
17
The roughness effect can be estimated by the indicator Irough(N) expressed by the difference between  app
eq
and the
18
19
20 ˆ eq with respect to  0  N  constant (Equation 16).
21
Fo

22  app
eq
 ˆ eq
23 I rough  N    100 (16)
24 0  N 
rP

25
26 2.5. Modelling of hardening effect
27
28 According to the modelling proposed by Deperrois [33], the superficial work-hardening effect is introduced in the
ee

29
30 Dang Van’s criterion through only the correction of  0  N  material constant. The  0 base material constant is
31
rR

32 supposed unchanged because it is ranged in little interval between 0.1 and 0.3 for many materials. In this case,
33
34
 0  N  become a surface constant  s ( N ) which changes with the local surface density of dislocations at the
35
ev

36
37 surface. Since the FWHM of X-ray diffraction profiles is linked qualitatively to the density of dislocations, then
38
iew

39 the surface Dang Van’s constant  s  N  was expressed by a power law of the surface parameter FWHM s . A
40
41 similar relationship was proposed by [5, 8] using the surface micro-hardness HVs .
42
43
 s ( N )  K  N  . H Vs 
n
44 (17)
45
46 Where K  N  and n are two material coefficients. K  N  can be identified from the values of the  0 ( N ) and
47
48
HV0 base material characteristics. So, Equation 17 will be expressed at the near surface region as follows:
49
50 n
51  HVs 
 s  N    0  N  .  (18)
52  H V0 
53
54 According to [4] assume, in the case of shot peened surfaces, that the n constant is equal to the work-hardening
55
56 coefficient of the Hollomon [42] law in the cyclic tensile behaviour.
57
58 The hardening effect can be estimated by the indicator Ihard expressed by the relative difference between the surface
59
60 and base material constants,  s ( N ) and  0  N  .

9
Journal of Materials Engineering and Performance Page 10 of 49

1
2
3  s  N   0  N 
4 I hard  N   (19)
5 0  N 
6
7 2.6. Modelling of damage effect
8
9 In the predictive approach of fatigue strength, the damage variable D, proposed by Lemaitre and Chaboche [44],
10
11 is used to take into account the effect of damage defects. This variable is defined at the macroscopic scale, on a
12
13 finite elementary volume, of sufficiently large size compared to the heterogeneities of the medium. On each section
14
of this volume, the area S and normal n, the traces of microcracks and microcavities cover a total area of area SD.
15
16
The damage variable relating to direction n is defined by (Figure 6):
17
18
SD
19 D ( n)  (20)
20 S
21
Fo

22 In this study, it is assumed that the distribution of surface defects is uniform and isotropic at all points of the treated
23
layer. In the presence of an isotropic damage of measure D, the section which effectively resists the forces is:
24
rP

25
26 Seff  S – S D =S 1 – D  (21)
27
28 We assume that the first layer representative volume element is damaged homogenously. The superficial damage
ee

29
30 variable, denoted DS , is supposed isotropic (independent on the direction). The effective cyclic stress tensor is
31
rR

32 then given by the following relationship:


33
𝟏
34 𝝈(𝒕) = (22)
(𝟏 ― 𝑫𝑺) 𝝈(𝒕)
35
ev

36
37 So, the effect of superficial damage results in a reduction of  0  N  parameter of the Dang Van’s criterion by the
38
iew

39 factor 1  DS  . The  damge  N  constant of the damaged surface expresses the unfavourable effect of superficial
40
41 damage on the material strength for a specific HCF life (N).
42
43
 damge  N    0  N  .1  DS  (23)
44
45
46 Therefore, the ductile damage effect can be estimated by the indicator Idamage expressed by the relative difference
47
48 between Dang Van’s constants with and without surface damage,  damage  N  and  0  N  .
49
50  damage  N    0  N 
51 I damage  N   (24)
0  N 
52
53
54 3. Experimental
55
56 The material and the experimental procedure are detailed in previous work [19]. The material studied is an AISI
57
58 316L austenitic stainless steel (equivalent to EN X2CrNiMo17-12-2). Two sets of V-notched specimens (Kt  1.6),
59
60 devoted to plane bending fatigue tests, were finished respectively by electro-polishing and severe grinding. The

10
Page 11 of 49 Journal of Materials Engineering and Performance

1
2
3 texture and the eventual defects were examined at the sample notch root using a scanning electron microscope
4
5 (SEM). The surface roughness was measured by profilometery. The microstructural changes were characterized
6
7 by cross-section Vickers micro-hardness measurements with a load of 100 g. The XRD technique was used to
8
9 measure the initial and stabilized residual stresses and also to evaluate qualitatively the superficial work-hardening
10
11 through the FWMH of XRD peak. Both sets of ground and electro-polished specimens were submitted to 4-point
12
13 plane bending fatigue tests in the HCF field (Figure 7). These tests were performed by applying controlled loads
14
15 with a censure of 2×106 cycles. The electro-polished specimens were tested using two loading ratios (R0.1 and R0.3),
16
while the ground specimens were tested with only one ratio (R0.1). The fatigue limits with 50 % failure probability
17
18
were estimated using the staircase method. The role of the surface integrity in the initiation and propagation of
19
20 fatigue micro-cracks was clarified by SEM examinations.
21
Fo

22 3.1. Initial surface integrity


23
24 The electro-polished surface appears sufficiently smooth by SEM observations (Figure 8-a). In addition, it is
rP

25
26 characterized by insignificant values of roughness parameters (Ra  0 and Rt  0), and also it is exempt of
27
28 significant levels of work-hardening and residual stresses (Table 1).
ee

29
30 The ground surface is characterized by a typical texture characterized by machining grooves in the transversal
31
rR

32 direction of specimens, and a distribution of folds and scales inherent to the grinding process and the material
33
34 ductility (Figure 8-b). It should be noted that the ground surface is completely exempt of thermal micro-cracks and
35
ev

36 burns. The characteristics of this surface are summarized in Table 1. The roughness parameters are more significant
37
38 than those of the electro-polished surface. Furthermore, the ground surface is characterized by a hardened layer of
iew

39
40 about 100 m, with a maximum hardening at the surface ( Hvs / Hv0  1.8 and FWHMs/FWHM0  2.8). The surface
41
42 residual stresses are higher tensile stresses, particularly in the transversal direction, i.e. parallel to grinding grooves.
43
44 3.2. HCF behaviour
45
46 The obtained S-N diagrams for both electro-polished and ground surfaces are given in Figure 9. They are expressed
47
48 by the nominal stress amplitude (  a ) versus the number of cycles to failure (fatigue life) in the HCF domain (105
49
50  Nf  2106 cycles). Two approximate curves with R1 = 0.1 and R2 = 0.3 are presented for the electro-polished
51
52 surface, and one curve with R1 = 0.1 for the ground surface. These curves are described by using the Basquin’s
53
54
55
 
relationship [43], that expresses a linear relation between  a  and N f (Eq. 23). The corresponding expressions

56
57 are determined by linear regression and given in Figure 9.
58
 a  A. N f 
B
59 (25)
60

11
Journal of Materials Engineering and Performance Page 12 of 49

1
2
3 For the three curves, when the number of cycles to failure increases from 105 until 106 cycles, the applied stress
4
5 level decrease following linear curves. From 106 until 2106 cycles, these curves become gradually horizontal
6
7 and tend to specific limits (fatigue limits). These fatigue limits at 2106 cycles (  f ) were determined in
8
9
accordance to the staircase method (Table 2). It was found that the fatigue limit of the electro-polished specimen
10
11 decreases of ≈ – 17 % when the loading ratio increases from R1 = 0.1 to R2 = 0.3. On the other hand, the fatigue
12
13 limit at R1 = 0.1 of the ground surface decreases greatly of about – 35 % in comparison with that of the electro-
14
15 polished surface.
16
17 All broken specimens at electro-polished or ground state are characterized by a transversal fracture which occurs
18
19 systematically at the notch root (Figure 10). SEM examinations of the notch root of a broken electro-polished
20
21 specimen (Figure 11-a) show a partially propagated secondary crack of length  150 m, situated near the fracture
Fo

22
23 surface (at a distance  70 m). In the first stage, this crack is initiated by the persistent slip band (PSB) mechanism,
24
rP

25 with a length  30 m and an orientation of  45° with respect to loading direction. This mechanism is generated
26
27 by the maximum shear stress on the most unfavourably oriented grains. In a second stage, the micro-crack grows
28
ee

29 in the transverse direction and tends to become long and unstable. Figure 11-b presents a SEM micrograph at the
30
31 notch root and near the fracture surface of a broken ground specimen. It reveals multiple cracks with sizes less
rR

32
33 than 200 m and localized in grinding grooves. These cracks are parallel to the groove direction and distributed
34
35 randomly until a distance  400 m from the fracture surface. SEM examinations of the fracture surface show that
ev

36
37 the fatigue micro-cracks were initiated in the root of a single grinding groove inducing the maximum stress micro-
38
iew

39 concentration (Figure 12-a). In addition, it appears some grinding grooves near the fracture surface, which have
40
41 approximately the same depth along the specimen width. On the other hand, the fracture surface reveals that the
42
43 fatigue-crack propagation has occurred with a less ductile manner in the ground layer ( 100 m) than in the bulk
44
45 material (Figure 12-b).
46
47 The stabilized surface residual stresses acting in both electro-polished and ground specimens are summarized in
48
Table 1. For the electro-polished surfaces, they reveal compressive longitudinal stresses and tensile transversal
49
50
stresses which don’t exceed, respectively, -140 MPa and +160 MPa at the surface. The stabilized residual stress
51
52 profiles corresponding to fatigue limit (R1 = 0.1) of the ground specimen reveal a partial stress relaxation at the
53
54 surface. This relaxation is greater in the longitudinal direction (-18%) than in the transversal direction (-7%).
55
56 4. Predictive results
57
58 4.1. Cyclic work-hardening behaviour
59
60 The applied cyclic stress tensors corresponding to fatigue limits and acting in the notch root of electro-polished

12
Page 13 of 49 Journal of Materials Engineering and Performance

1
2
3 and ground specimens (Eq. (2)) were determined by a cyclic elasto-plastic 2D FE model [19]. The cyclic behaviour
4
5 was modelled based on the strain-controlled test results, and using the nonlinear isotropic/kinematic hardening
6
7 material model proposed by Lemaitre and Chaboche [44]. In this model, the Von Misses yield criterion, a flow
8
9 rule and the nonlinear kinematic-isotropic hardening variables Xij and Rij are expressed as following:
10
11 f  J 2  ij  X ij   R  k  0 (26)
12
13
2
14 dX ij  Cd  ijp   X ij dp (27)
15 3
16
17 dR  b Q  R  dp (28)
18
19
20
In Eqs. (19)-(20), dp   2 / 3 d  ij : d  ij indicates equivalent plastic strain increment. Xij and R mean the centre
21
Fo

22 and the size of yield surface, respectively. Model coefficients k, C, , Q and b were identified in accordance with
23
24 reference [44], and given in Table 3.
rP

25
26 In this model, the initial material properties including superficial work-hardening and residual stresses are taken
27
28 into account. Furthermore, the constants of the cyclic hardening model were calibrated in order to have good
ee

29
30 agreement between numerical and experiment values of stabilized residual stresses.
31
rR

32 The cyclic stress/strain curves corresponding to fatigue limits for the electro-polished under 4-point fatigue tests
33
(R1 = 0.1 and R2 = 0.3) are presented in figure 13-a and 13-b, respectively. The cyclic behaviour of the electro-
34
35
ev

polished surface, under the stress ratio R1 = 0.1 (Figure 13-a), is characterized by a gradual plastic accumulation
36
37
conducting to an elastic-shakedown response after many loading cycles (N  50 cycles). During the shakedown
38
iew

39
process, both cyclic ratcheting and mean stress relaxation (including residual stress relaxation) occur
40
41
simultaneously with decreasing rate. The initial tensile residual stress (  Rxx = + 100 MPa) are completely relaxed
42
43
44 and transformed into significant compressive stress (  *Rxx = -180 MPa). With the stress ratio R2 = 0.3 (Figure 13-
45
46 b), the cyclic behaviour of the electro-polished surface is characterized by an elastic-shakedown response
47
48 appearing just after the first cycle. This conducted to complete relaxation of the initial residual stress (  Rxx = + 100
49
50
MPa;  *Rxx = - 20 MPa).
51
52
53 Figures 14 show, for the ground surface, the variation on the stabilized residual stress profile with different values
54
55 of the non-linear kinematic hardening constants (C and ). It was found that application of the base material
56
57 constants (Table 3) conducts to a great residual stress relaxation, the numerical profiles of stabilized residual
58
59 stresses are clearly lower than experimental ones. This imposes to calibrate the cyclic model parameters (C,  , Q
60

13
Journal of Materials Engineering and Performance Page 14 of 49

1
2
3 and b) of the ground surface. For the non-linear kinematic constants, an increasing of the C kinematic hardening
4
5 modulus (30000; 70000 MPa) and a decreasing of the  nonlinear recovery parameter (60; 20), allow to reduce
6
7 significantly the residual stress relaxation and to have numerical residual stresses more near to the experimental
8
9 ones. On the other hand, the Q and b constants of the isotropic hardening law of the base material are considered
10
11 unchanged after grinding.
12
13 The calibrated cyclic model parameters for the ground surface under 4-point bending fatigue tests (R = 0.1) are
14
15 given in Table 3. The cyclic behaviour of this surface (Figure 14-c) appears to be characterized by an immediate
16
17 elastic-shakedown response that occurred just after one loading cycle, and accompanied by a slight plastic strain (
18
19  xxp  0.001). This conducted to a partial relaxation of the initial residual stress which remain tensile stress (  Rxx =
20
21 + 520 MPa;  *Rxx = + 300 MPa).
Fo

22
23
4.2. Prediction of multiaxial fatigue strength
24
rP

25 The stress tensors present a plane stress sate at the sample notch root. The components and the maximum Von
26
27
Mises stress (  Mises
max
) of the applied stresses tensor are given in Table 4. The two stress tensors for the electro-
28
ee

29
30 polished specimen, under both bending tests (R1 = 0.1 and R2 = 0.3), are used to identify the base material constants
31
of the Dang Van’s criterion (0 = 0.137 and 0 = 171 MPa). The cyclic stress tensor applied at the notch root of
rR

32
33
fatigue specimen is calculated at the fatigue limit (for N = 2×106 cycles) of the ground specimen (Eq. 21). The
34
35
ev

applied equivalent stress was deduced (Eq. 21) and compared to the 0(N) constant by using the indicator Igrind(N)
36
37
38 (Eq. 23). It was found a degradation of multiaxial fatigue strength estimated by Igrind(N)  -37%, which is close to
iew

39
the fatigue limit degradation (≈ -35%). It is shown in the Dang Van’s diagram by the location of the loading
40
41
representative point in the resistance zone (Fig. 15-a).
42
43
44  xxa sin( wt )   xxm 0 0
 
45  app ( t )   0  yy sin( wt )   yym
a
0
46  0 0 0 

47
48
(29)
49
50
1 a 1
51  app
eq
N   xx   0  N  . xxa   yya   xxm   yym 
52 2 3 (30)
53
54  app
eq
 N   0  N 
55 I grind  N  = (31)
0  N 
56
57
58 4.3. Prediction of surface integrity effects on HCF strength
59
60 4.3.1. Stabilized residual stress effect

14
Page 15 of 49 Journal of Materials Engineering and Performance

1
2
3 The effect of quasi-stabilized residual stresses was predicted through the application of the Dang Van’s criterion
4
5 to the total stress state (Equation 24) acting at the notch root of fatigue specimen. The corresponding equivalent
6
7 stresses are given by Equation 25. This effect is then given by IRS  -30 % and appears highly detrimental for
8
9 fatigue strength. It is shown in the Dang Van’s diagram by a displacement of the loading representative point
10
11 towards the positive pressures (Figure 15-b).
12
13  xxa .sin w( t )   xxm   *Rxx 0 0
14  
 t ( t )   app ( t ) +  *R   0  .sin w( t )    
a
yy
m
yy
*
Ryy 0 (32)
15 
 0 0 0 
16
17
1 a 1
 xx   0  N  . xxa   yya   xxm   yym   *Rxx   *Ryy 
18
 teq  N  
19 2 3 (33)
20
21 4.3.2. Roughness effect
Fo

22
23 The roughness effect is manifested by multiple micro-crack nucleation in grinding grooves which localized at the
24
rP

25 notch root. It can be considered as a stress micro-concentration effect induced by the deepest and sharpest grooves
26
27 [22]. In this study, the highest stress micro-concentration in grinding grooves was simulated by the FE method.
28
ee

29 The corresponding site was modelled by a V-shaped critical micro-notch characterised by three dimensions: high
30
31 (H), width (W) and notch root radius (  ). This defect is identified by the deepest and sharpest groove from the
rR

32
33 highest measured roughness profile (Ra  1.5 m; Rt  11.7  m ) , as shown in Figure 16. The  radius was
34
35
ev

estimated using a graphical radius gage according to [25]. A best-fit circle defined by the maximum area of contact
36
37 was inscribed in the critical micro-notch root. It was found that H  10  m , W  60  m and   5 m . The stress
38
iew

39
micro-concentration due to the presence of the critical micro-notch in the sample notch root was simulated using
40
41 a 2D elastic FE model based on the plane strain assumption. This model was developed using the commercial FE
42
43 code ABAQUS [45]. The fatigue sample geometry was meshed using six-nodded triangular elements with small
44
45 size around the micro-notch, and then subjected to a static loading. The material behaviour is assumed to be
46
47 isotropic and linear elastic with a Young’s modulus of 196 GPa and a Poisson’s ratio of 0.29. This permitted to
48
49
determine K̂ txx  3.67 and to estimate K̂ fxx  1.16 using the Peterson’s model [22]. The micro-geometrical cyclic
50
51
52 stress tensor due to the critical micro-notch effect, and the corresponding equivalent stress are given by Equations
53
54 26 and 27, respectively.
55
56  K̂ fxx . xxa .sin w( t )   xxm  0 0
 
57 ˆ ( t )   0  yy .sin w( t )   yym
a
0 (34)
58  
 0 0 0
59 
60

15
Journal of Materials Engineering and Performance Page 16 of 49

1
2
3
4 ˆ eq  N  
2

1 ˆ
 1
3
ˆ .  a   m    a   m 
K fxx . xxa   0  N  .  K fxx xx xx yy yy 
(35)
5
6 So, this effect seems to be enough detrimental on fatigue strength (Irough ≈ -12%) in comparison with the total
7
8 grinding effect (Igrind  -37%). It is traduced in the Dang Van’s diagram (Figure 15.c) by a displacement of the
9
10 loading representative point towards the failure zone.
11
12 4.3.3. Damage effect
13
14 According to the experimental investigations conducted by SEM exams, it was found that as-ground surface is
15
16 absolutely free of damage effect. Therefore, this effect can be assumed negligible ( Ds  0 and I damage  0 ).
17
18 4.3.4. Superficial hardening effect
19
20
In this study, the  s  N  constant taking into account of the superficial hardening effect is identified by calibration
21
Fo

22
23 of the Dang Van’s criterion. Then, the effective cyclic stress tensor (  eff ( t )) including the residual stress and
24
rP

25 roughness effects (Equation 36) is determined, and the prediction of the Dang Van’s criterion is considered exact
26
27
28
  N     N  
eq
eff s (Equation 37). In the Dang Van’s diagram, the representative point of cyclic stresses is
ee

29
30 situated on the multiaxial threshold of the ground material (Figure 15-d). Therefore, the constants K and n are
31
rR

32 equal to 171 MPa and 0.05, respectively. So, the superficial work-hardening beneficial effect is traduced by an
33
34 increase of the material constant (  s  180 MPa) in comparison with the base material constant (  0  171 MPa) .
35
ev

36 This increase is estimated by the indicator Ihard (Equation 19). Therefore, this effect seems to be slight beneficial
37
38 (Ihard ≈ +5%) as shown in the Dang Van’s criterion by a little increase of the resistance domain of the ground
iew

39
40 material (Figure 15-d).
41
42  K̂ fxx . xx  t    *Rxx 0 0
43  
 eff ( t )  ˆ ( t )   RS
*
 0  yy ( t )   *Ryy 0 (36)
44  0 0 0 
45 
46
47
48
 effeq  N  
2

1 ˆ
 1
3
ˆ .  a   m    a   m   *   *     N 
K fxx . xxa   0  N  .  K fxx xx xx yy yy Rxx Ryy  s
(37)

49
50 4.4. Fatigue life prediction
51
52 For the experimental fatigue lives of the ground surface ( N theo ), arranged between 105 and 2106 cycles, the
53
54 prediction of theoretical fatigue lives was carried out by using the Dang Van’s criterion with calibrated
55
56 formulation. The predictive values were then compared to the four selected experimental values ( N exp = 2,5×105,
57
58 7,5×105, 106, 2×106 cycles) deduced from the Basquin’s model of the S-N curve. According to the predictive
59
60 method, the first step consists to use the S-N expressions of the electro-polished surfaces (for R1 = 0.1and R2 =

16
Page 17 of 49 Journal of Materials Engineering and Performance

1
2
3 0.3) to simulate the admissible cyclic stresses acting in the notch root of specimens for a given fatigue life ( N exp ).
4
5 These stresses are used to estimate the corresponding fatigue threshold (𝜶𝟎(𝑵𝒆𝒙𝒑); 𝜷𝟎(𝑵𝒆𝒙𝒑)) of the base material
6
7 in the Dang Van’s criterion. The second step is devoted to the simulation of admissible cyclic stress state (for
8
9 N exp ) acting in the notch root of the ground surface reported in the Dang Van’s criterion.
10
11
12 The comparison between the theoretic ( N theo ) and the experimental ( N exp ) fatigue lives is presented in Figure 18.
13
14 It can be seen that predictions are in good agreement with experiments, with all are slightly greater values within
15
16 a factor of 2. The predictive calculations are, consequently, conservative for the case of 316L steel ground surface.
17
18 5. Discussion
19
20 A predictive methodology for HCF behaviour and life of mechanical parts was proposed and applied in the case
21
Fo

of 316L steel surface at ground state. In the following sections, the validity of this methodology will be discussed
22
23
in relation with experimental observations.
24
rP

25 5.1. Residual stresses effect


26
27 The grinding residual stresses acting in the sample notch root are characterized by a partial relaxation at fatigue
28
ee

29 limit and remain high tensile stresses. This relaxation is more significant in the longitudinal direction (-18%), i.e.
30
31 the bending stress direction, than in the transversal direction which corresponds to groove direction (-7%). This
rR

32
33 is justified by the superposition of the residual stresses with the applied cyclic stresses which exceed the yield limit
34
35
ev

of the ground material, particularly in the bending direction [15]. The residual stress relaxation was simulated by
36
37 the FE method in previous work [19]. The material behaviour was described by an isotropic and non-linear cyclic
38
iew

39 model. It was found that the notch root behaviour is characterized by an elastic-shakedown response which
40
41 occurred just after the first loading cycle, and conducting to a lower plastic deformation (𝜀𝑝𝑒𝑞 ≈ 0.0011).
42
43 The predictive evaluation on multiaxial fatigue strength showed that the grinding residual stresses at stabilized
44
45 state have great influence (𝐼𝑅𝑆 ≈ ― 30 %) in comparison with the total grinding effect(𝐼𝑔𝑟𝑖𝑛𝑑 ≈ ― 37 %). This
46
47 result is in good agreement with several works showing that the grinding residual stresses have great influence for
48
49 ductile materials, unless they are enough stable during cyclic loading [17-19]. On the other hand, if only the
50
51 residual stress effect is considered, then the prediction of fatigue strength of the ground surface is clearly not
52
53 conservative((𝜎𝑒𝑞
𝑡 ― 𝛽0)/𝛽0 ≈ ― 7 %). This prediction shows the significant contribution of the surface roughness
54
55 on fatigue strength degradation of the ground specimen. It should be noted that the residual stresses relaxation is
56
57 more important than the level of applied stress is higher. Consequently, the residual stress effect is more in low
58
important in low cycle fatigue life than in high cycle fatigue one.
59
60

17
Journal of Materials Engineering and Performance Page 18 of 49

1
2
3 5.2. Surface roughness effect
4
5 It’s well established that the surface roughness effect on fatigue limit of materials depends strongly on many factors
6
7 such as surface finishing process, roughness parameters such as Ra, Rz, Rt and Ry, material characteristics, loading
8
9 direction, etc. [22]. Siebel and Gaeir [20] found that the reduction in fatigue limit is significant above a certain
10
11 critical groove depth (Rt*  2 m), and they suggested that this reduction is proportional to log (Rt). Skali and
12
13 Geslot [17] showed that the effect of grinding grooves (Rt = 11.5 m) on plane bending fatigue limit of AISI 4140
14
15 steel is  10%. Taylor and Clancy [22] revealed in the case of stress relieved AISI 4140 steel that even the relatively
16
17 low surface roughness (Ra = 0.5-1.4 m, Rmax = 7-14 m) produced by grinding reduced fatigue life (11%),
18
19 compared to polished samples (Ra = 0.1-0.3 m, Rmax = 3-5 m). The Rmax parameter is defined by the authors as
20
21 the maximum pit depth recorded on a 1.75 mm length of the surface.
Fo

22
23 In this work, the experimental investigations showed that the grinding roughness is relatively higher (Ra1.4 m,
24
rP

25 Rt10.5 m). The effect of this roughness on fatigue behaviour is characterized by multiple micro-cracks nucleated
26
27 in grinding grooves and distributed randomly until a distance  400 m from the fracture surface (at the middle of
28
ee

29 the sample notch root). This effect is due to the greatest stress micro-concentrators, i.e. the deepest and sharpest
30
31 grooves of the sample notch root [22]. The potential crack nucleation site was modelled by a V-shaped critical
rR

32
33 micro-notch, and the corresponding stress concentration factor ( Kˆ t  3.67) was simulated by the FE method and
34
35
ev

36 used to estimate the fatigue stress concentration factor ( Kˆ f  1.16) . This permitted to estimate the roughness effect
37
38 ( I rough  12% ) which appears enough significant, but lower than the residual stress effect ( I RS  30%) . This
iew

39
40
result agrees well with many works [18-21]. On the other hand, if only the roughness effect is considered then the
41
42
fatigue strength prediction of the ground surface is largely not conservative (( ˆ   0 ) /  0  25%) . However,
eq
43
44
45 it becomes conservative when roughness and residual stress effects are both considered ((  eff
eq
  0 ) /  0  5%) .
46
47 5.3. Superficial work-hardening effect
48
49 The microstructural change of the ground surface is characterized by an important increase of the micro-hardness
50
51 (80 %) at the near surface region. This superficial work-hardening is also indicated by an increase of the FWHM
52
53 parameter which is closely related to the increase of the density of dislocations [31-32]. The effect of work-
54
55 hardening on the initiation and propagation of fatigue micro-cracks was highlighted by SEM examinations of the
56
57 fracture surface. On one hand, the fatigue micro-cracks were initiated in grinding grooves (i.e. at the surface). On
58
59 the other hand, the fracture surface appears less ductile in the hardened layer ( 100 m) than in the bulk material.
60

18
Page 19 of 49 Journal of Materials Engineering and Performance

1
2
3 Consequently, this effect could retard the initiation fatigue stage and may be beneficial to fatigue strength of the
4
5 ground material [30].
6
7 The prediction of the superficial work-hardening effect on fatigue strength has been carried using the  s constant
8
9 for the ground material in the Dang Van’s criterion. Deperrois [33] proposed a theoretical law to relate s to the
10
11
FWHMs parameter by a power law. The K coefficient is generally identified using the  0 and FWHM 0 base
12
13
material characteristics, while the n constant is assumed to be equal to 0.5 according to the theory of dislocations
14
15
[33]. Fathallah and laamouri [4] proposed to use the theory of plasticity and assumed that n is equal to the hardening
16
17
coefficient deduced from the monotonic work-hardening law of the studied material (i.e. n  0.3 for 316L steel).
18
19
20 In this study, the  s constant was identified by calibration of the Dang Van’s criterion, and appears slightly
21
Fo

22 greater (  s = 180 MPa) than the base material constant (  0 = 171 MPa). Consequently, it was found that n = 0.05
23
24 is largely lower than the values defined by the theories of dislocations and of plasticity. The superficial work-
rP

25
26 hardening effect is estimated by 𝐼ℎ𝑎𝑟𝑑 ≈ +5 % which indicates that this effect is slightly beneficial in comparison
27
28 to the detrimental effects. This result shows that although the superficial work-hardening is highly increased by an
ee

29
30 abusive grinding, it improves slightly the fatigue resistance of the base material.
31
rR

32 6. Conclusion
33
34 In this study, an efficient predictive methodology of HCF behaviour and life is proposed. It permits to a better
35
ev

36 knowledge and evaluation of surface integrity effects (including stabilized tensile residual stresses, surface
37
roughness and superficial work-hardening) on fatigue performance of a machined metallic surface. The main
38
iew

39
conclusions of this study are summarized as follows:
40
41 (i) The proposed methodology has been validated In the case of ground 316L steel surface, where an important
42
43 fatigue limit degradation (-35%) in comparison with the electro-polished one (virgin state) is observed.
44
45 (ii) The predictive calculations, based on the Dang Van’s criterion, has permitted to evaluate the surface
46
47 integrity effects on fatigue strength through three indicators (IRS, Irough, Ihard and Idamage). It appears clearly
48
49 that the fatigue strength degradation is mainly due to the detrimental effect of stabilized tensile residual
50
51 stresses (IRS  -30 %), while the detrimental roughness effect appears less important (Irough  -12 %).
52
53 However, the beneficial work-hardening and damage effects seems to be lower (Ihard  +5 % and Idamage 
54
55 0).
56
57 (iii) The predictive methodology can be used to optimize the conditions of a machining process or a surface
58
59 treatment for best surface integrity effects on HCF behaviour and life of mechanical components.
60

19
Journal of Materials Engineering and Performance Page 20 of 49

1
2
3 Declarations:
4
5 a. Funding (information that explains whether and by whom the research was supported): 'Not
6
7
applicable'
8
9
b. Conflicts of interest/Competing interests (include appropriate disclosures): 'Not applicable'
10
11
12 c. Availability of data and material (data transparency): 'Not applicable'
13
14 d. Code availability (software application or custom code): 'Not applicable'
15
16 e. Ethics approval (include appropriate approvals or waivers): 'Not applicable'
17
18 f. Consent to participate (include appropriate statements): 'Not applicable'
19
20 g. Consent for publication (include appropriate statements): 'Not applicable'
21
Fo

22 h. Authors' contributions (optional: please review the submission guidelines from the journal whether
23
24 statements are mandatory): 'Not applicable'
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

20
Page 21 of 49 Journal of Materials Engineering and Performance

1
2
3 References
4
5 [1] Volker Schulze, Modern Mechanical Surface Treatment, 2006 WILEY-VCH Verlag GmbH & Co. KGaA,
6
7 Weinheim, ISBN-13: 978-3-527-31371-6
8
9 [2] Wagner L (1999) Mechanical surface treatments on titanium, aluminum and magnesium alloys. Materials
10
11 Science and Engineering A263:210–216. https://doi.org/10.1016/S0921-5093(98)01168-X
12
13 [3] Gao Y, Li X, Yang Q, Yao M (2007), Influence of surface integrity on fatigue strength of 40CrNi2Si2MoVA
14
15 steel. Materials letters. 61(2):466-469. DOI:10.1016/j.matlet.2006.04.089
16
[4] Fathallah R, Laamouri A, Sidhom H, Braham C (2004) High cycle fatigue behavior prediction of shot peened
17
18
parts. Int J Fatigue 26(10):1053–1067. https://doi.org/10.1016/j.ijfatigue.2004.03.007
19
20 [5] Sidhom N, Laamouri A, Fathallah R, Braham C, Lieurade HP (2005) Fatigue strength improvement of 5083
21
Fo

22 H11 Al-alloy T-welded joints by shot peening: experimental characterization and predictive approach. Int J Fatigue
23
24 27(7):729–747. https://doi.org/ 10.1016/j.ijfatigue.2005.02.001
rP

25
26 [6] Liang T, Changfeng Y, Dinghua Z, Junxue R, Zheng Z, Jiyin Z (2020) Evolution of surface integrity and
27
28 fatigue properties after milling, polishing, and shot peening of TC17 alloy blades. Int J Fatigue, 136,105630.
ee

29
30 https://doi.org/10.1016/j.ijfatigue.2020.105630
31
rR

32 [7] Ben Fredj N, Sidhom H (2006) Effects of the cryogenic cooling on the fatigue strength of the AISI 304 stainless
33
34 steel ground components, Cryogenics. 46:439-448. http://dx.doi.org/10.1016/j.cryogenics.2006.01.015
35
ev

36 [8] A Laamouri, F Ghanem, C Braham, H Sidhom (2019), Influences of up-milling and down-milling on surface
37
38 integrity and fatigue strength of X160CrMoV12 steel. Int J Adv Manufa Technol, 105, 1209-1228
iew

39
https://doi.org/10.1007/s00170-019-04280-2
40
41
[9] Fang Q, Zhitong C, Qiantong L, Shimin G (2019) Effects of process combinations of milling, grinding, and
42
43 polishing on the surface integrity and fatigue life of GH4169 components. Proc IMechE Part B: J Engineering
44
45 Manufacture 1–11. DOI: 10.1177/0954405419868053
46
47 [10] Jouinia N, Revel P, Thoquenne G (2020) Influence of surface integrity on fatigue life of bearing rings finished
48
49 by precision hard turning and grinding. Journal of Manufacturing Processes 57:444-451.
50
51 https://doi.org/10.1016/j.jmapro.2020.07.006
52
53 [11] Suáreza A, Veigaa F, Polvorosab R, Artazaa T, Holmbergc J,. López de Lacalleb L.N, Wretlande A (2019)
54
55 Surface integrity and fatigue of non-conventional machined Alloy 718. 48:44-50.
56
57 https://doi.org/10.1016/j.jmapro.2019.09.041
58
59 [12] Chomiennea V, Verdua C, Rech J, Valiorgue F (2013) Influence of surface integrity of 15-5PH on the fatigue
60

21
Journal of Materials Engineering and Performance Page 22 of 49

1
2
3 life. Procedia Engineering 66:274-281. doi: 10.1016/j.proeng.2013.12.082.
4
5 [13] Klocke F, Welling D, Dieckmann J (2011) Comparison of Grinding and Wire EDM Concerning Fatigue
6
7 Strength and Surface Integrity of Machined Ti6Al4V Components. Procedia Engineering, 19:184-189.
8
9 https://doi.org/10.1016/j.proeng.2011.11.099
10
11 [14] Lopes KSS, Sales WF, Palma ES (2008) Influence of Machining Parameters on Fatigue Endurance Limit of
12
13 AISI 4140 Steel. J. of the Braz. Soc. of Mech. Sci. & Eng. 30 (1). https://doi.org/10.1590/S1678-
14
15 58782008000100011
16
[15] Totten G, Howes M, Inoue T, Handbook of Residual Stress and deformation of steel, ASM International,
17
18
Materials Park, Ohio, ISBN: 0-87170-729-2, USA, 2002.
19
20 [16] Webster GA, Ezeilo AN (2001) Residual stress distribution and their influence on fatigue lifetimes, Int J
21
Fo

22 Fatigue. 23:375-383. https://doi.org/10.1016/S0142-1123(01)00133-5


23
24 [17] El-Helieby SOA, Rowe GW, A quantitative comparison between residual stresses and fatigue properties of
rP

25
26 surface-ground bearing steel (En31). Wear. 58 (1980) 155-172. https://doi.org/10.1016/0043-1648(80)90220-3
27
28 [18] Skalli N, Geslot R, Effects of grinding conditions on fatigue behaviour of 42CD4 grade steel. CIRP Annals -
ee

29
30 Manufacturing Technology. 32 (1983) 481-484.
31
rR

32 [19] Laamouri A, Sidhom H, Braham C (2013), Evaluation of residual stress relaxation and its effect on fatigue
33
34 strength of AISI 316L stainless steel ground surfaces: Experimental and numerical approaches. Int J Fatigue
35
ev

36 48:109-121. http://dx.doi.org/10.1016/j.ijfatigue.2012.10.008
37
38 [20] Novovic D, Dewes RC, Aspinwall DK, Voice W, Bowen P (2004) The effect of machined topography and
iew

39
integrity on fatigue life, International Journal of Machine Tools and Manufacture. 44:125-134.
40
41
https://doi.org/10.1016/j.ijmachtools.2003.10.018
42
43 [21] D. Taylor, O.M. Clancy, Fatigue performance of machined surfaces, Fatigue and Fracture of Engineering
44
45 Materials and Structures 14(2-3) (1991) 329-336. http://dx.doi.org/10.1111/j.1460-2695.1991.tb00662.x
46
47 [22] R.E. Peterson Stress concentration factors. New York: John Wiley and Sons, 1974.
48
49 [23] H. Neuber, In:Kerbspannungsleshre. Berlin, Germany: Springer-Verlag; 1958. p. 159-63.
50
51 [24] Arola D, Ramulu M (1999) An examination of the effects from surface texture on the strength of fiber-
52
53 reinforced plastics. J compos Mater 33(2) 101-86. https://doi.org/10.1177%2F002199839903300201
54
55 [25] Arola D, Williams CL (2002) Estimating the fatigue stress concentration factor of machined surfaces, Int. J.
56
57 Fatigue. 24:923-930. http://dx.doi.org/10.1016/S0142-1123(02)00012-9
58
59 [26] Ås SK, Skallerud B, Tveiten BW, Holme B (2005) Fatigue life prediction of machined components using
60

22
Page 23 of 49 Journal of Materials Engineering and Performance

1
2
3 finite element analysis of surface topography. Int. J. Fatigue. 27:1590-1596.
4
5 https://doi.org/10.1016/j.ijfatigue.2005.07.031
6
7 [27] Ås SK, Skallerud B, Tveiten BW (2008) Surface roughness characterization for fatigue life predictions using
8
9 finite element analysis, Int. J. Fatigue. 30:2200-2209. http://dx.doi.org/10.1016/j.ijfatigue.2008.05.020
10
11 [28] Ben Fredj N, Sidhom H, Braham C (2006) Ground surface improvement of the austenitic stainless steel AISI
12
13 304 using cryogenic cooling, Surface & Coatings Technology. 200:4846-4860.
14
15 http://dx.doi.org/10.1016/j.surfcoat.2005.04.050
16
[29] Ben Fathallah B, Ben Fredj N, Sidhom H, Braham C, Ichida Y (2009) Effects of abrasive type, cooling mode
17
18
and peripheral grinding wheel speed on the AISI D2 steel ground surface integrity. International Journal of
19
20 Machine Tools & Manufacture. 49:261-272. http://dx.doi.org/10.1016%2Fj.ijmachtools.2008.10.005
21
Fo

22 [30] Sahara H (2005) The effect on fatigue life of residual stress and surface hardening resulting from different
23
24 cutting conditions of 0.45%C steel, International Journal of Machine Tools and Manufacture. 45:131-136.
rP

25
26 http://dx.doi.org/10.1016%2Fj.ijmachtools.2004.08.002
27
28 [31] Balart MJ, Bouzina A, Edwards L, Fitzpatrick ME (2004) The onset of tensile residual stresses in grinding of
ee

29
30 hardened steels, Materials Science and Engineering. A367:132-142. https://doi.org/10.1016/j.msea.2003.10.239
31
rR

32 [32] Vashista M, Paul S (2012) Correlation between full width at half maximum (FWHM) of XRD peak with
33
34 residual stress on ground surfaces, Philosophical Magazine. 92(33):4194-4204.
35
ev

36 https://doi.org/10.1080/14786435.2012.704429
37
38 [33] Deperrois A, Castex L, Dang Van K, Merrien P, Bignonnet (1990) A High cycle fatigue life of surface
iew

39
hardened steels under multi-axial loading, in: H. Kitagawa, T. Tanaka (Eds.), Proceedings of the Fourth
40
41
International Fatigue Conference, vol. 4, Hawaii, USA, July 15-20, pp. 2417-2422.
42
43 [34] Paul S, Bandyopadhyay PP, Chattopadhyay AB (1993) Effects of cryo-cooling in grinding steels, Journal of
44
45 Materials Processing Technology. 37:791-800. https://doi.org/10.1016/0924-0136(93)90137-U
46
47 [35] Toten GE, Factors relating to heat treating operations, failure analysis and prevention, Vol 11, ASM
48
49 Handbook, ASM International, 2002.
50
51 [36] Benedetti M, Fontanari V, Santus C, Bandini M (2010) Notch fatigue behaviour of shot peened high-strength
52
53 aluminium alloys: Experiments and predictions using a critical distance method, International Journal of Fatigue
54
55 32,10:1600-1611. https://doi.org/10.1016/j.ijfatigue.2010.02.012
56
57 [37] Guechichi H, Castex L (2006) Fatigue limits prediction of surface treated materials, Journal of Materials
58
59 Processing Technology. 172:381-387. https://doi.org/10.1016/j.jmatprotec.2005.10.010
60

23
Journal of Materials Engineering and Performance Page 24 of 49

1
2
3 [38] Seddik R, Petit EJ, Ben Sghaier R, Atig A, Fathallah R (2017) Predictive design approach of high-cycle
4
5 fatigue limit of shot-peened parts, The International Journal of Advanced Manufacturing Technology, 93:2321–
6
7 2339 (2017). https://doi.org/10.1007/s00170-017-0704-4
8
9 [39] Benkhettab M, Guechichi H, Benkabouche SE (2019) Fatigue strength prediction methodology of shot-
10
11 peened materials, The International Journal of Advanced Manufacturing Technology, 104:4277–4287.
12
13 https://doi.org/10.1007/s00170-019-04089-z
14
15 [40] Dang Van K (1999) Introduction to fatigue analysis in mechanical design by the multiscale approach, in : K.
16
Dang Van, I.V. Papadoupoulos, High Cycle Metal Fatigue in the Context of Mechanical Design. CISM Courses
17
18
and Lectures N° 392, Springer-Verlag, pp 57-88.
19
20 [41] Dalaeia K, Karlssona B, Svenssona LE (2010) Stability of residual stresses created by shot peening of pearlitic
21
Fo

22 steel and their influence on fatigue behavior. Procedia Engineering 2:613–622.


23
24 http://dx.doi.org/10.1016/j.proeng.2010.03.066
rP

25
26 [42] Hertzberg RW. Deformation and fracture mechanics of engineering materials. John Wiley & Sons. Inc. 1996.
27
28 p. 17, ISBN 0-471-01214-9.
ee

29
30 [43] Usabiaga H, Muniz-Calvente M, Ramalle M, Urresti I, Fernández Canteli A (2020) Improving with
31
rR

32 Probabilistic and Scale Features the Basquin Linear and BiLinear Fatigue Models. Engineering Failure Analysis,
33
34 Vol. 116, 104728. https://doi.org/10.1016/j.engfailanal.2020.104728
35
ev

36 [44] Lemaitre J, A course on damage mechanics, 2nd revised and enlarged ed. New York: Springer-Verlag, Berlin
37
38 Heidelberg: 1996, ISBN 3 540 60980 6.
iew

39
40 [45] Hibbitt, Karlsson, Sorensen, Inc (2014). ABAQUS Standard User’s manual, version 6.14, USA.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

24
Page 25 of 49 Journal of Materials Engineering and Performance

1
2
3 List of figures:
4
5 Fig. 1. Ground surface integrity [5].
6
7 Fig. 2. Flow-chart of the proposed methodology for prediction of the effects of surface integrity on fatigue lifetime
8
9 of ductile metallic parts.
10
11 Fig. 3. Evolution of macroscopic shear stress for a crystal with isotropic hardening.
12
13 Fig. 4. Paths of the microscopic and macroscopic shear stress on a facet of normal n.
14
15 Fig. 5. Procedure for introducing of initial residual stresses in a FE model (notched fatigue specimen).
16
Fig. 6. Definition of damage variable [44].
17
18
Fig. 7. Illustration of the 4-point plane bending fatigue test applied to a V-notch specimen (Kt  1.6).
19
20
Fig. 8. Surface texture of the AISI 316L stainless steel obtained by: (a) electro-polishing and (b) grinding.
21
Fo

22 Fig. 9. S-N diagrams.


23
24 Fig. 10. Fatigue transversal fracture localized at the notched zone of the ground specimen.
rP

25
26 Fig. 11. SEM micrographs showing the mechanisms of fatigue micro-cracks nucleation. (a) Electro-polished
27
28 surface, and (b) ground surface.
ee

29
30 Fig. 12. SEM micrographs of the fracture surface for a ground specimen: (a) fatigue crack nucleation sites, (b)
31
rR

32 decrease of ductility for the ground layer in comparison with the bulk material.
33
34 Fig. 13. Cyclic elasto-plastic behaviour at fatigue limit for the electro-polished surfaces: (a) under loading ratio
35
ev

36 R0.1; (b) under loading ratio R0.3.


37
38 Fig. 14. Influence of different values of the kinematic hardening parameters on (a) the stabilized residual stress
iew

39
40 profile in the x-direction (xx) ; (b) the stabilized residual stress profile in the y-direction (yy) ; and (c) the cyclic
41
42 hardening behaviour.
43
44 Fig. 15. Dang Van’s diagrams showing the surface integrity effects on fatigue strength of ground 316L stainless
45
steel surface under 4-point bending test. (a) Total grinding effect (Igrind ≈ -37 %). (b) Detrimental effect of
46
47
stabilized tensile residual stresses (IRS ≈ -30 %). (c) Detrimental effect of surface roughness (Irough ≈ -12 %). (d)
48
49 Beneficial effect of superficial work-hardening (Ihard ≈ +5 %) and calibration of the Dang Van’s criterion for the
50
51 ground surface.
52
53 Fig. 16. Identification of the critical micro-notch from the roughness profile of maximum Rt, and estimating of
54
55 its effective root radius () using a graphical radius gage.
56
57 Fig. 17. Cyclic behaviour of the electro-polished surface under different loading levels for R0.1 and R0.3 stress
58
59 ratios.
60

1
Journal of Materials Engineering and Performance Page 26 of 49

1
2
3 Fig. 18. Difference between theoretical values of fatigue lifetime ( N theo ) and experimental values ( N exp ) .
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Fo

22
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

2
Page 27 of 49 Journal of Materials Engineering and Performance

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Fig. 1. Ground surface integrity.
20
21
Fo

22
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

3
Journal of Materials Engineering and Performance Page 28 of 49

1
2
3
4
5 Start
6
7
8 Input
9
10 A given N cycles FATIGUE SPECIMEN Machining Cyclic bending
11 Material
(104  N  Nlimit) (geometry, dimensions) conditions (0a, R)
12
13 104<N<Nlimit
14
15
16 Bending SN Cyclic Monotonic Surface integrity (RS, -HV, SN curve (for
Exp. Processor
curves (R1 ; R2) model model Roughness, damage defects) calibration)
17
18
19
20
21 Two nominal stresses FE simulation of Kt due to roughness
Fo

22 at N fatigue life and estimation of Kf


23
24 FE simulation of quasi-stabilized FE simulation of the quasi-stabilized
rP

25 stresses (residual and cyclic) stresses (residual and cyclic)


26 ia = 0a + i.a
27
28 Identification of fatigue criterion
ee

constants (,  Dang Van’s criterion i := i + 1


29
30
31 No
rR

32  eq  N    grind  N   
33  = 0.1
34 Num. Processor
Yes
35
ev

36
37 Fatigue resistance (alim(N), R) Output
38 for a given N cycles
iew

39
40
Surface integrity effects:
41
IRS(N), Irough(N), Ihard(N), Idamage(N)
42
43
44
45 End
46
47
48
49
Fig. 2. Flow-chart of the proposed methodology for prediction of the machined surface integrity effects on
50
51
fatigue lifetime of ductile metallic parts.
52
53
54
55
56
57
58
59
60

4
Page 29 of 49 Journal of Materials Engineering and Performance

1
2
3
4 C, 
5
Non-symmetrical macroscopic
6 cycle
7
8
9
10
11
12
13 Cm
14 ̂
15 min  
16
17 max
18
19 ̂ Elastically shakedown
20 Mesoscopic cycle (symmetrical)
21
Fo

22
23
24
rP

25
26 Fig. 3. Evolution of macroscopic shear stress for a crystal with isotropic hardening.
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

5
Journal of Materials Engineering and Performance Page 30 of 49

1
2
3 Macroscopic path
4
5
6
7
C(n)
8 Mesoscopic path M
C(n, t)
9
10
11
12 Cm(n)
13
14
15 O
16 [ Facet of normal n]
17
18
19
20
21 Fig. 4. Paths of the microscopic and macroscopic shear stress on a facet of normal n.
Fo

22
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

6
Page 31 of 49 Journal of Materials Engineering and Performance

1
2
3
4
5 Initial residual stresses and plastic stain of
6 ground layers
7
8
9 Introduction in each layer
10 of 10 m
11
12
13
Initial plastic strain ( 
p
14 )
15 Initial RS (  )
R
16
17  R
xx 0 0
  Identification : X  (2 / 3)C 
p
18 R  0  R
yy 0
19  0 0 0  Introduction :  p  E : R
20 
21
Fo

22
23
24 User subroutine User subroutine
« SIGINI »
rP

25 « UHARDINI »
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38 Fig. 5. Procedure for introducing of initial residual stresses in a FE model (notched fatigue specimen).
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

7
Journal of Materials Engineering and Performance Page 32 of 49

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15 Fig. 6. Definition of damage variable [44].
16
17
18
19
20
21
Fo

22
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

8
Page 33 of 49 Journal of Materials Engineering and Performance

1
2
3
4 z
5
6
7 F/2 F/2
8
9
10
11
12
13 x
14 O y
15
16
17
18
19
20
21
Fo

22 Fig. 7. Illustration of the 4-point plane bending fatigue test applied to a V-notch specimen (Kt  1.6).
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

9
Journal of Materials Engineering and Performance Page 34 of 49

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Fo

22 (a)
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
(b)
46
47
48 Fig. 8. Surface texture of the AISI 316L stainless steel obtained by: (a) electro-polishing and (b) grinding.
49
50
51
52
53
54
55
56
57
58
59
60

10
Page 35 of 49 Journal of Materials Engineering and Performance

1
2
3 250
4
5 Electropolished (4-p ; R=0.1)
6 230
7 Electropolished (4-p ; R=0.3)
8
Ground (4-p ; R=0.1)
9 210
10
Nominal stress amplitude, a (MPa)

11
12 190
13
14
15 170
16
17
18 150
19
20
21 130
Fo

22
23
24 110
rP

25
26
27 90
28
ee

29
30 70
31
rR

32
33 50
34
1E+05 1E+06 1E+07
35
ev

36 N (cycles)
37
38
iew

Fig. 9. S-N diagram.


39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

11
Journal of Materials Engineering and Performance Page 36 of 49

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19 Fig. 10. Fatigue transversal fracture localized at the notched zone of the ground specimen.
20
21
Fo

22
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

12
Page 37 of 49 Journal of Materials Engineering and Performance

1
2
3
4 Loading direction
5
6
7
8 Nucleation site
9
10
11
12 Fracture surface
13
14
15
16
17
18
19
20
21
Fo

22
23
24
rP

25
26 (a)
27
28
ee

29
30
31
rR

32
33
Fracture surface
34
35
ev

36
37
38
iew

39
40
41
42
43 Multiple micro-cracks
44
45
46
47
48
49
50
51
52
53 (b)
54
55 Fig. 11. SEM micrographs showing the mechanisms of fatigue micro-cracks nucleation. (a) Electro-polished
56
57 surface, and (b) ground surface.
58
59
60

13
Journal of Materials Engineering and Performance Page 38 of 49

1
2
3
4
5
6
7
8
9
10
11
12
Fatigue crack nucleation sites
13 Grinding grooves
14
15
16
17
18
19
20
21
Fo

22
23
24
rP

25 (a)
26
27
28
ee

29
30
31
rR

32 Fracture surface of the


33 ground layer ( 100 m)
34
35
ev

36
37
38
iew

39
40
41
Fracture surface of the
42 bulk material
43
44
45
46
47
48
49
50
51
52 (b)
53
54 Fig. 12. SEM micrographs of the fracture surface for a ground specimen: (a) fatigue crack nucleation sites, (b)
55
56 decrease of ductility for the ground layer in comparison with the bulk material.
57
58
59
60

14
Page 39 of 49 Journal of Materials Engineering and Performance

1
2
3 800

Stress, -xx (MPa)


4
Closed loop after
5 700
the first cycle
6
600 First hysteresis
7
loop
8 500
9
10 400
11
12 300
13
200
14
15 100 Initial RS
16
17 0
18 -0.001 0.000 0.001 0.002 0.003 0.004
19 -100 Plastic strain, -xx
20 -200
21
Fo

Relaxed RS
22 -300
23
24
rP

25
26 (a)
27
28
ee

29
30 800
Stress, -xx (MPa)

31
700
rR

32
33 600 Closed loop after
34 the first cycle
35 500
ev

36
37 400
38
iew

300
39
40 200
Relaxed RS

41
42 100
Initial RS
43
44 0
45 -0.0010 0.0000 0.0010 0.0020 0.0030 0.0040
-100
46
47 -200
48
49 -300 Plastic strain, -xx
50
51 (b)
52
53
54
55 Fig. 13. Cyclic elasto-plastic behaviour at fatigue limit for the electro-polished surfaces: (a) under loading ratio
56
57 R0.1; (b) under loading ratio R0.3.
58
59
60

15
Journal of Materials Engineering and Performance Page 40 of 49

1
2
3 800
4 Experimental

Stabiliezd residual stress, -xx(MPa)


5 700 C=30000 ; gamma=60 ; Q=150 ; b=1
6 C=70000 ; gamma=60 ; Q=150 ; b=1
7 600 C=70000 ; gamma=20 ; Q=150 ; b=1
8 500
9
10 400
11
12 300
13 200
14
15 100
16
17 0
18 0 20 40 60 80 100 120 140 160
-100
19 Depth, z(m)
(a)
20
800
21
Fo

Experimental
Stabiliezd residual stress, -yy(MPa)

22 700 C=30000 ; gamma=60 ; Q=150 ; b=1


23 C=70000 ; gamma=60 ; Q=150 ; b=1
24 600 C=70000 ; gamma=20 ; Q=150 ; b=1
rP

25
500
26
27 400
28
ee

29 300
30
200
31
rR

32 100
33
34 0
35 0 20 40 60 80 100 120 140 160
ev

-100
36 Depth, z(mm)
(b)
37
Stress, -xx(MPa)

1000
38
iew

900 C=30000 ; gamma=60 ; Q=150 ; b=1


39
C=70000 ; gamma=60 ; Q=150 ; b=1
40 800 C=70000 ; gamma=20 ; Q=150 ; b=1
41
42 700
43 600
44 Initial RS
500
45
46 400
Relaxed RS
47 300
48
49 200
Closed loop after
50 100 the first cycle
51 0
52 0 0.0005 0.001 0.0015 0.002
53 Plastic strain, p-xx
54 (c)
55 Fig. 14. Influence of different values of the kinematic hardening parameters on (a) the stabilized residual
56
57 stress profile in the x-direction (xx) ; (b) the stabilized residual stress profile in the y-direction (yy) ; and (c)
58
59 the cyclic hardening behaviour.
60

16
Page 41 of 49 Journal of Materials Engineering and Performance

1
2
3 Total stresses (with residual stress effect)
4 Applied stresses (without surface effects)
Applied stresses (without surface effects)
5 200 200
6 180 Failure zone
180 Failure zone
7
160 160
8 Multi-axial fatigue Multi-axial fatigue
9 140 threshold (Base material) 140 threshold (Base material)
a (MPa)

10 120 120

a (MPa)
11 100 100
12
80 80
13
14 60 60
15 40 40
Resistance zone Resistance zone
16 20 20
17 0 0
18
0 200 400 600 800 1000 1200 1400 0 200 400 600 800 1000 1200 1400
19
Pmax (MPa) Pmax (MPa)
20
21
Fo

(a) (b)
22
23
24
rP

25
Micro-geometrical stresses (with roughness effect) Effective stresses (with RS and roughness effects)
26
27 Applied stresses (without surface effects) Applied stresses (without surface effects)
28 200 200
ee

Failure zone
29 180 180 Failure zone

30 160 160
31 Multi-axial fatigue Multi-axial fatigue
140 140
rR

threshold (Base material) threshold (Base material)


32
120
a (MPa)

120
a (MPa)

33 Multi-axial fatigue
34 100 100 threshold (ground
35 80 80 material)
ev

36 60 60
37 40 40
38 Resistance zone Resistance zone
iew

20 20
39
40 0 0
41 0 200 400 600 800 1000 1200 1400 0 200 400 600 800 1000 1200 1400
42 Pmax (MPa) Pmax (MPa)
43
44 (c) (d)
45
46 Fig. 15. Dang Van’s diagrams showing the surface integrity effects on fatigue strength of ground 316L
47
48 stainless steel surface under 4-point bending test. (a) Total grinding effect (Igrind ≈ -37%). (b) Detrimental
49
50 effect of stabilized tensile residual stresses (IRS ≈ -30%). (c) Detrimental effect of surface roughness (Irough ≈ -
51
52 12%). (d) Beneficial effect of superficial work-hardening (Ihard ≈ +5%) and calibration of the Dang Van’s
53
criterion for the ground surface.
54
55
56
57
58
59
60

17
Journal of Materials Engineering and Performance Page 42 of 49

1
2
3
4 80
5 70
6

Surface height (m)


60
7
8 50
9
40
10
11 30
12
20
13
14 10
15 0
16 0 0.5 1 1.5 2
17 Length (mm)
18
19
20
21
Fo

22 80
23    m
24 70
H = 10 m
rP

25
60
26
27 50
28
ee

29 40
30
31 30
W = 60 m
rR

32
20
33
34 10
35
ev

36 0
37 0.35 0.36 0.37 0.38 0.39 0.4 0.41 0.42 0.43 0.44 0.45
38
iew

39
40
41
42 Fig. 16. Identification of the critical micro-notch from the roughness profile of maximum Rt, and estimating of
43
44 its effective root radius () using a graphical radius gage.
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

18
Page 43 of 49 Journal of Materials Engineering and Performance

1
2
3 1000
4
5
6 800
7
8
9 600
10
11
400
Stress, sxx(MPa)

12
13
14 200
15  = 240 MPa
16
17 0
18 0 0.001 0.002 0.003 0.004 0.005
19  = 300 MPa
20 -200
21  = 320 MPa  = 340 MPa
Fo

22
-400
23  = 360 MPa
24
rP

25 -600
26 strain, epxx
27
28 Fig. 17. Cyclic behaviour of the ground surface under different loading levels.
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

19
Journal of Materials Engineering and Performance Page 44 of 49

1
2
3
1E+07
4
5
6
7
8
9
10
11
N-theo (cycles)

12
13
14 1E+06
15
16
17
18
19
20
4-p bending
21
Fo

Exact prediction
22
23 Under prediction of -50 %
24 Over prediction of + 100 %
rP

25 1E+05
26 1E+05 1E+06 1E+07
27
28 N-exp (cycles)
ee

29
30 Fig. 18. Difference between theoretical values of fatigue lifetime ( N theo ) and experimental values ( N exp ) .
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

20
Page 45 of 49 Journal of Materials Engineering and Performance

1
2
3 List of tables:
4
5 Table 1. Effects of grinding and electro-polishing on surface integrity and fatigue limits of AISI 316L austenitic
6
7 stainless steel.
8
9 Table 2. Comparison fatigue strength evaluated by Basquin relation and Staircase method
10
11 Table 3. Cyclic hardening constants of Chaboche’s law for the base and ground 316L steel states.
12
13 Table 4. Applied cyclic stress tensors (at fatigue limits) in the notch root of electro-polished and ground specimens.
14
15
16
17
18
19
20
21
Fo

22
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

1
Journal of Materials Engineering and Performance Page 46 of 49

1
2
3 Table 1. Effects of grinding and electro-polishing on surface integrity and fatigue limits of AISI 316L austenitic stainless steel.
4
5 Surface Surface integrity Fatigue results
6
7 state Roughness Work-hardening Initial residual stresses Stabilized residual stresses Stress Fatigue limit at 2×106 Fatigue
8
9 Ra Rt Hvs FWMHs Hardening  Rxx (MPa)  Ryy (MPa)  Rxx
* (MPa)  Ryy
* (MPa) ratio, R cycles, D (MPa) degradation
10
11 (m) (m) (°) depth (m) rate (%)*

Fo
12
13 Electro- - 140 + 160 0.1 350 ± 16 0

rP
14 ≈0 ≈0 190 1.2 ≈0 + 80 + 80
15 polished - 45 + 130 0.3 290 ± 16 - 17

ee
16
17 Ground 1.4 10.5 342 2.8 ≈ 100 + 510 + 765 + 420 + 710 0.1 230 ± 19 - 35
18

rR
19
20

ev
21 * The 4-point bending fatigue limit (R) of the electro-polished specimen is considered as reference.
22

iew
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42 2
43
44
45
46
Page 47 of 49 Journal of Materials Engineering and Performance

1
2
3 Table 2. Comparison fatigue strength evaluated by Basquin relation and Staircase method
4
5 Surface state Fatigue test Fatigue strength, f (MPa) Fatigue strength, f (MPa)
6
7 (Basquin relation) (Staircase method)
8
9 N = 105 cycles N = 2106 cycles N = 2106 cycles
10
11 Polished Bending 4-p (R0.1) 242 170 175
12
13 Polished Bending 4-p (R0.3) 211 135 145
14
15 Ground Bending 4-p (R0.1) 181 104 115
16
17
18
19
20
21
Fo

22
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

3
Journal of Materials Engineering and Performance Page 48 of 49

1
2
3 Table 3. Cyclic hardening constants of Chaboche’s law for the base and ground 316L steel states.
4
5 Material state Cyclic hardening E  k C  Q b
6
7 conditions (GPa) (MPa)
8
9 Base material Reverse tension- 196 0.29 220 30000 60 150 1
10
11 compression (R = -1)
12
13 Ground state 4-point bending (R = 0.1) 196 0.29 220 70000 20 150 1
14
15
16
17
18
19
20
21
Fo

22
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

4
Page 49 of 49 Journal of Materials Engineering and Performance

1
2
3 Table 4. Applied cyclic stress tensors (at fatigue limits) in the notch root of electro-polished and ground specimens.
4
5 Surface Fatigue limits at 2×106 Stress ratio, Applied cyclic stress tensor,  VM
max

6 (MPa)
7 state cycles, D (MPa) R  app ( t ) (MPa)
8
9
10  290 sin( wt )  355 0 0
11
 0 87 sin( wt )  107 0 
350 0.1  574
12  0 0 0 
13 Electro-
14
polished 146 sin( wt )  270 0 0
15
 0 44 sin( wt )  81 0 
16 290 0.3  370
17
 0 0 0 
18
19
20 191 sin( wt )  233 0 0
 0 57 sin( wt )  70 0 
21 Ground 230 0.1  377
Fo

22  0 0 0 
23
24
rP

25
26
27
28
ee

29
30
31
rR

32
33
34
35
ev

36
37
38
iew

39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

You might also like