Pinto 2020 Fretting Fatigue Under Variable Amp

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

*Manuscript

Click here to view linked References

1
2
3
4
5
6
7
8
9 Fretting fatigue under variable amplitude loading
10
11
considering partial and gross slip regimes: numerical
12 analysis
13
14
15 A.L. Pintoa,c,∗, R.A. Cardosob , R. Talemic , J.A. Araújoa
16 a Department of Mechanical Engineering, University of Brasilia, 70910-900 Brasilia, DF,
17 Brazil
18 b Department of Mechanical Engineering, Federal University of Rio Grande do Norte,

19 59078-970 Natal, RN, Brazil


c Department of Materials Engineering, KU Leuven, 3001 Leuven, Belgium
20
21
22
23
24
25 Abstract
26 The main aim of this work was to numerically investigate the effect of se-
27
quences of variable amplitude shear loading on contact tractions/stresses and
28
on the development of life estimation methodologies for these more challeng-
29
30 ing loading histories. In this setting, a finite element (FE) model has been
31 implemented in order to simulate a cylinder on plane contact problem for an
32 aeronautical Al 7075-T651 alloy. High-Low (H-L) and Low-High (L-H) shear
33 amplitude loading sequences under partial and gross slip conditions were con-
34 sidered in three different case studies. For the case studies where gross slip was
35 involved, the proposed life estimation procedure was enhanced to include the
36 effect of material loss due to wear. More explicitly, such procedure was based on
37 the use of (i) the SWT(Smith-Watson-Topper) multiaxial fatigue model, (ii) the
38 Theory of Critical Distances, (iii) Miner’s cumulative damage rule and (iv) the
39 numerical update of the contact surfaces profiles (for loading blocks under gross
40 slip). The numerical analyses for the case studies under partial slip revealed
41 that, when the presence of wear is neglected in the modelling, the Miner’s rule
42 provided a divergence between the expected life for the H-L and L-H loading
43 blocks. This difference was considerably larger when gross slip took place. For
44 the case study involving the presence of gross slip in one of the shear loading
45 blocks, the calculated life tended to infinite when the damage generated by the
46
high amplitude block is greater than a certain critical value. These studies are
47
the basis for an experimental programme which will be carried out in the 4
48
49 actuators fretting fatigue rig of the University of Brasilia.
50 Keywords: Fretting Fatigue, Multiaxial Fatigue, Variable Amplitude Loading,
51 Wear.
52
53
54
∗ Corresponding author
55
56
57
58
59 Preprint submitted to Tribology International January 8, 2020
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 1. Introduction
10
11 Fretting occurs when two contacting parts are subjected to fluctuating loads,
12 which provokes relative slip of small amplitude between them. Strong stress
13 gradients and wear due to the contact loads may lead to the early initiation
14 of cracks. In this case, in presence of bulk fatigue loads, one has the so-called
15 fretting fatigue phenomenon. Under such conditions, initiated cracks may prop-
16 agate at high rates resulting in strong fatigue resistance reductions [1]. Fretting
17 fatigue takes place in many practical situations such as gears, aircraft fuse-
18 lages, connection between rotor and blades in jet engines, flanges, bolted joints
19 in pipelines and overhead conductors [2, 3, 4, 5]. Fretting fatigue studies and
20 tests have been carried under constant amplitude loads [6, 7, 8, 9]. However,
21 in practice, under actual working conditions, components such as the blade and
22 disk attachment and overhead conductors are subjected to variable amplitudes
23
over time [10, 11]. In such cases, some questions are not trivial to respond, for
24
instance: how to determine the fatigue life of a component subjected to variable
25
26 fretting loading conditions from the knowledge of simple stress (S) × life (N)
27 curves of the material? How to account the strong stress gradient inherent to
28 fretting problems? Is it necessary to consider wear in such analysis?
29 In order to predict the fatigue life of a component subjected to variable
30 amplitude loads, or to random loads, a cumulative damage law must be used.
31 Palmgren-Miner’s law, or simply Miner’s rule [12, 13], is one of the best known
32 in the technical literature. This is a linear damage accumulation rule, i.e., it
33 estimates that the order of application of the variable amplitude loading blocks
34 does not change the component fatigue resistance, as long as the total damage
35 produced by such blocks is the same at the end. According to this law, the order
36 of application of the load blocks with variable amplitude does not change the
37 fatigue resistance or the fatigue life of the material as long as the fatigue damage
38 per each block is the same. In this setting, it is important to verify Miner’s law
39 applicability evaluating its performance in classical fretting fatigue tests subject
40 to loadings with variable amplitude. Although there are a few works evaluating
41 fretting fatigue response under variable amplitude loading conditions, most of
42
them rely on the construction of fatigue resistance curves for further comparison
43
with variable amplitude loading conditions. In this case, even when Miner’s
44
45 rule is considered, simple uniaxial fatigue approximations are adopted. In this
46 setting, a brief review of previous works in this field is here presented.
47 In the early 1980s Kantimathil and Alic [14] reported fretting fatigue tests
48 conducted on the 7075-T7351 experiencing variable loading. Specimens were
49 subjected to constant amplitude loading conditions followed by tensile overloads
50 applied at each 1000 cycles. Compressive overloads were also assessed. For the
51 constant amplitude loading cycles, different load ratios were considered. From
52 their results, they have observed that the influence of the overloads were negli-
53 gible for constant amplitude cycles with high maximum bulk stresses. However,
54 for lower maximum values of the bulk stress in the constant amplitude cycles,
55 overloads seemed to have a beneficial effect in terms of fatigue resistance, which
56 may be explained by crack growth retardation due to plastic deformation at the
57
58
59 2
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 crack tip.
10 Mutoh et al. [15] published a paper in 1987 where they examined the fretting
11 fatigue resistance of a low strength medium carbon steel for two-step variable
12 amplitude loading blocks. Each block consisted of n1 cycles at a given bulk stress
13 amplitude σ1 followed by n2 cycles at σ2 . In other words, each block consisted
14 of (n1 + n2 ) fretting cycles. Different combinations of n1 and n2 per block were
15 considered in their analysis. They concluded that when the block contained a
16 large enough number of cycles and the number of blocks to failure were smaller
17 than 5 or 6, the variable amplitude loading condition lead to life reductions up
18
to 10 times. It was also shown that Miner’s linear rule did not always provide
19
good estimates, as it was the case for plain fatigue conditions. Besides, they
20
21 have highlighted that fretting fatigue life is highly influenced by the number of
22 cycles at high stress levels, where cack growth rates become significant.
23 Two years later, Mutoh et al. [16] investigated the fretting fatigue resistance
24 of a high strength steel under random loading conditions. Two load spectrum
25 were examined (A and B). In spectrum A, the maximum value of the bulk
26 stress was 45 % greater than its RMS over time, while in spectrum B it was 85
27 % greater. Overall speaking, random loading conditions reduced fretting fatigue
28 lives by a factor of 1.5. In addition, Miner’s rule led to satisfactory life estimates
29 but being generally nonconservative.
30 In the late 1990s Troshchenko et al. [17] investigated the fatigue damage
31 accumulation in Al and Ti alloys under variable loading conditions in the pres-
32 ence of stress concentration and fretting conditions. In their analysis, failure
33 is not necessarily assumed to occur when Miner’s cumulative damage reaches
34 the unity. Then, they reported that damage computations by assuming Miner’s
35 linear rule for the fretting fatigue case under variable amplitude loading were
36 equal or exceeded those from smooth specimens. The work was based on the
37
construction of S-N curves considering constant amplitude conditions for a fur-
38
ther comparison with tests carried out under variable amplitude programs.
39
40 Cortez et al. [18], in the same period, investigated the fretting fatigue behav-
41 ior of the Ti-6Al-4V under constant and variable amplitude loading. Besides,
42 they also evaluated the influence of the loading frequency on the experiments.
43 For the variable amplitude loading tests, the loading history consisted of 500
44 high-frequency small amplitude cycles at 200 Hz followed by a single large am-
45 plitude cycle at 1 Hz. The maximum load for small and large amplitude cycles
46 were the same. Different positive load ratios were considered in the small ampli-
47 tude cycles while a load ratio of about 0.1 was observed in the large amplitude
48 cycles. Two main conclusions were drawn concerning the variable amplitude
49 loading tests. Firstly, it was observed that the variable amplitude loading con-
50 ditions did not present a significant impact on the fretting fatigue resistance
51 of the Ti-6Al-4V assessed. In addition, from the constant amplitude fretting
52 fatigue loading tests, they have used Miner’s rule to estimate life at variable
53 loading conditions. In this case, Miner’s rule provided reasonably accurate re-
54 sults although slightly optimistic.
55 Kinyon and Hoeppner [19] in 2000 also presented interesting results concern-
56
ing fretting fatigue tests under variable amplitude loading. Tests were carried
57
58
59 3
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 out on a Ti alloy. Three different amplitudes of bulk stress were considered in
10 their analysis (σ1 > σ2 > σ3 ), where both two and three-step loading blocks
11 were investigated. For the two-step loading tests, the loading sequence consisted
12 of one cycle subjected to σ1 followed by another cycle subjected to σ2 . These
13 two cycles with different amplitudes were repeated up to the specimen’s failure
14 (Case 1). Similarly, another two-step loading block was considered but at this
15 time replacing σ2 by σ3 (Case 2). For the three-step loading set (Case 3), the
16 specimen was subject to one cycle at each stress amplitude σ1 , σ2 and σ3 , where
17 100 these three blocks were repeated until failure. Throughout fretting fatigue tests
18
performed with constant amplitude loads (σ1 ,σ2 and σ3 ), the authors found that
19
the application of Miner’s rule provided unsafe life estimates for the Case 1 and
20
21 conservative predictions for the Cases 2 and 3. At the end, the authors claimed
22 that Miner’s rule did not seem appropriate to model fretting fatigue lives.
23 In the 2000s Kondo et al. [20] investigated the fretting fatigue behaviour
24 under variable amplitude loading conditions for a low alloy steel. In their work,
25 the authors first obtained a fretting fatigue limit diagram (like the modified
26 Goodman line) when considering constant amplitude loadings. Different tests
27 were carried out considering the same contact indentation load but with different
28 fatigue load ratios. At this stage, for fatigue limit conditions under small mean
29 bulk loads, non-propagating cracks were observed. As the mean bulk load was
30 increased in the tests, cracks were not formed at all. In the next, the authors
31 carried fretting fatigue tests considering repeated two-step loadings. In this
32 setting, one of the loading blocks considered fatigue limit conditions by the
33 application of a fully reversed fatigue stress (σ1 ), while the other block, with
34 amplitude σ2 , was set considering positive mean stresses. It is worth noting
35 that, in this case, σ2 was chosen well below the fatigue limit obtained from the
36 fretting fatigue limit diagram with constant amplitude. Summarizing, it was
37
observed that, even though σ1 and σ2 were below the fatigue limit for constant
38
amplitude loading conditions, the repeated two-step loading of σ1 and σ2 could
39
40 lead to failure. At the end of the paper, these authors highlight that care should
41 be taken when using fatigue limit data from constant amplitude fretting fatigue
42 tests once under variable amplitude loading, failure may occur even for stresses
43 below fatigue limit conditions.
44 Massingham and Irvin [21] in 2006 start their work by highlighting the lack of
45 investigations on fretting fatigue under variable amplitude loading. They have
46 developed a finite element model to assess the influence of overloads followed
47 by the maintenance of constant amplitude loads during the simulations. As the
48 main conclusion, it is verified that in spite of the application of two or more
49 overload cycles, the position of the stick-slip zone boundary as well as shear
50 stress distributions depends only on the largest previous cycle. In other words,
51 the order that the overloads are applied is important in determining whether
52 shear stress distributions are modified, besides, it is the largest previous overload
53 that will dictate the shear stress distribution in subsequent smaller amplitude
54 cycles. The authors also suggest that these changes in stress distribution tends
55 to make linear damage rules act conservatively.
56
More recently, Gandiolle and Fouvry [22] have conducted experiments and
57
58
59 4
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 a numerical study in order to evaluate the crack propagation rate for fretting
10 fatigue under variable loading conditions. Crack length was obtained during
11 the tests by using a potential drop technique. Further, a numerical decoupled
12 approach was used to obtain the crack stress intensity factor, which allowed
13 relating the crack extension with the number of cycles by considering Paris law.
14 Even though they have not focused on the specific effects of considering variable
15 amplitude loads, the strategy permitted the estimate of important parameters
16 such as crack arrest, the crack extension and the number of cycles to failure.
17 This comprehensive review of the state of the art on variable amplitude fret-
18
ting fatigue revealed that most of the research carried out up to now was based
19
on the use of (i) S-N curves generated by fretting fatigue tests performed on one
20
21 actuator fretting fatigue devices (thus bulk and fretting loads are coupled), (ii)
22 the use of Miner’s damage rule and (iii) uniaxial fatigue analysis to estimate life.
23 This is obviously a rather simple research approach as it misses the systematic
24 investigation of a number of well know important characteristics of the fretting
25 problem such as the wear, the stress gradient, the multiaxial state of stress, the
26 varying stress ratio, contact nonlinearities, among others parameters. As there
27 is a lack of control of such parameters in these previous works, it has not been
28 possible to draw more firm conclusions even about the applicability of Miner’s
29 linear damage rule to fretting fatigue [15, 16, 17, 18, 19]. Although more de-
30 tailed studies have been proposed more recently [21, 22], they are clearly just a
31 few, which considered a limited range of materials and loading conditions.
32 In this setting, the aim of this work was to numerically investigate the effect
33 of sequences of variable amplitude shear loadings on contact tractions/stresses
34 and on the development of life estimation methodologies for these more chal-
35 lenging loading histories. In this setting, a finite element (FE) model has been
36 implemented in order to simulate a cylinder on plane contact problem for an
37
aeronautical Al 7075-T651 alloy. High-Low (H-L) and Low-High (L-H) shear
38
amplitude loading sequences under partial slip conditions or under a mix of
39
40 partial and gross slip conditions are considered in three different case studies.
41 For the case studies where gross slip was involved, the proposed life estimation
42 procedure was enhanced to include the effect of material loss due to wear. As
43 far as the authors are aware, this is the first attempt to investigate the impact of
44 sequences of variable shear amplitude loadings, which provoke a mix of different
45 contact slip regimes (partial and gross), on fretting fatigue life. Also new is the
46 investigation on the role of wear in the life estimation methodology for variable
47 amplitude loading. At last but not least, it should be remarked that another
48 important contribution of this work is the fact that it constitutes the basis for
49 the generation of experimental data (design of tests), which will allow to vali-
50 date (or not) the different hypothesis considered to construct the life estimation
51 procedure.
52
53
54 2. Multiaxial fatigue life estimation
55 Several approaches and damage parameters can be applied to problems sub-
56
jected to fretting conditions. These parameters can be classified conceptually as
57
58
59 5
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 critical plane approaches [23, 24, 25], stress invariant approaches [26, 27], fret-
10 ting specific parameters [28, 29] and continuum damage mechanics approaches
11 [30, 31]. The model used in this study was the Smith-Watson-Topper (SWT)
12 which is a multiaxial strain energy based critical plane approach [24]. This
13 parameter is better suited for materials and loading conditions where cracks
14 propagate perpendicular to the direction of maximum principal stress and strain
15 (Mode I cracks). The SWT criterion can be applied to both low and high cycle
16 fatigue. In addition, this model can also handle non-proportional loading condi-
17 tions [32]. All those features associated with its simplicity made this parameter
18
one of the most popular in determining fatigue life and planes of crack initiation
19
[33]. The SWT parameter can be defined as:
20
21 SW T = σn,max εn,a (1)
22
23 where σn,max and εn,a are the maximum normal stress and the strain amplitude
24 on the material plane that maximizes their product. It also allows estimating
25 the component life by reference to a fully reversed uniaxial test where stress
26 (Basquin) and strain (Coffin Manson) life relations can be combined yielding:
27
28 σf
29 σn,max εn,a = (2Nf )2b + σf εf (2Nf )b+c (2)
E
30
31 where σf and εf are the fatigue strength and ductility coefficients, respectively,
32 while b and c are the fatigue strength and ductility exponents, respectively.
33
34
35 3. Critical distance approach
36 The mechanical contact in the fretting problem is a stress raiser and, in
37
this case, it is necessary to consider the effects of the stress gradient caused by
38
the contact between the surfaces. In the 1930s, the development of the The-
39
40 ory of Critical Distance (TCD) was started with the works of [34, 35] in an
41 attempt of predicting fatigue failures in metallic components in the presence of
42 geometric features. Neuber proposed the Line Method (LM) and Peterson the
43 Point Method (PM). Whitney and Nuismer [36] evaluating monotonic loading
44 conditions were capable of defining the critical distance for the PM and LM
45 through a relationship between the Mode I fracture toughness, KIc , and the
46 tensile strength of smooth specimens. A decade later Tanaka presented a re-
47 lation with an identical theoretical derivation and equally valid when applied
48 to High Cycle Fatigue (HCF) [37]. Subsequently, several studies were carried
49 out with the objective of proving the feasibility of predicting the fatigue limit
50 resistance by using the TCD for HCF [38, 39, 40].
51 In general, the TCD define an effective stress, σef f , which is obtained
52 through an averaging procedure over a volume surrounding the stress raiser.
53 The method assumes that fatigue failures occurs when σef f exceeds a reference
54 material strength. For 2D analysis, the volume method can be simplified by
55 considering stress averages over an area, line (Area and Line Methods, respec-
56
tively) or even at a point located at a critical distance, L, from the stress raiser
57
58
59 6
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12 eff
13
14 L
15
16
17
18
19 Figure 1: Schematic representation of the Point Method.
20
21
22 (Point Method) [41]. The PM is used in this work, Figure 1. In this approach,
23 for HCF, the critical distance assumes the following form [39]:
24  2
25 1 ΔKth
L= (3)
26 2π Δσ−1
27 where, ΔKth is the threshold stress intensity factor range and Δσ−1 is the
28 uniaxial plain fatigue limit range both obtained at fully reversed loading con-
29
figurations.
30
On the other hand, the relation to estimate the critical distance to static
31
32 strength of notched members is given by [42, 43]:
33  2
1 KIc
34 Ls = (4)
2π σr
35
36 200 where the KIc is the plane strain fracture toughness and σr is a reference mate-
37 rial constant which may be equal to or greater than the ultimate tensile strength,
38 Sut .
39 As can be seen in Equations (3) and (4), the values of the critical distances
40 under fatigue and static loading are usually different, which leads us to assume
41 that, the critical distance, LM , at medium-high cycle fatigue regime depends on
42 the number of cycles to failure, Nf . In this work, as a simplifying procedure,
43 one assumes that LM = L for any number of cycles.
44
45
46 4. Wear model
47
In order to model wear by means of material loss, the most commonly used
48
models are the Archard’s law [44] and dissipated friction energy [45]. Several
49
50 authors have published works by using these models to account the material
51 removal due to wear when fretting takes place [46, 45, 47, 48, 49, 50]. Some
52 other were also published out of the fretting scope [51, 52, 53].
53 In this work, Archard’s law is used to model wear and according to this, the
54 volume of material removed by surface wear can be calculated as follows:
55 K
56 V = Ps (5)
H
57
58
59 7
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 where P is the normal load, s is the slip distance, H is the hardness of the
10 material (MPa) and K is a dimensionless constant. In Equation (5), the ratio
11 K/H can be replaced by the specific wear coefficient, κ. McColl et al. [46]
12 developed a local version of the Archard’s equation with the objective of using
13 it on a nodal basis in finite element analyses. With their formulation it is
14 possible to calculate the local wear depth through the following relation:
15
16 dh = κp(x)ds (6)
17
18 where, dh is the local depth of material removed for a given incremental relative
19 slip ds and p is the contact pressure. For practical reasons, in this formulation,
20 the local wear coefficient κ is assumed to be constant over the fretting cycles
21 and have the same value along all the contact region. Additionally, it can be
22 obtained from measured wear profiles of experimental tests [46].
23 Therefore, for an incremental implementation of Equation (6) in a FE frame-
24
work one has:
25
26 
n inc

27 Δhi,j = κp(xj , tk )Δs(xj , tk )ΔN (7)


28 k=1
29
30 where Δhi,j is the wear depth increment for the node j on the contact surface
31 for the ith fretting cycle simulated and ninc is the number of increments of
32 the ith fretting cycle. The contact variables p(xj , tk ) and Δs(xj , tk ) are the
33 contact pressure and the relative slip increment, respectively, for the node j at
34 the time increment tk . Note that Equation (7) is multiplied by a factor ΔN .
35 For problems involving high number of cycles such as the fretting problem, it is
36 a common practice to amplify wear computation by an accelerator factor ΔN
37 [46, 47, 54, 55, 49, 50]. In this case, if one wants to consider the total number of
38 cycles, Nt , in the analysis, one has to simulate only Nt /ΔN fretting cycles. The
39 choice of ΔN must be carried out carefully so that it minimizes computational
40 costs without introduce instabilities to the solution [46].
41
42
43 5. Application to fretting fatigue
44
45 Most of the works on fretting fatigue generally consider constant amplitude
46 loadings in the analyses. A typical loading configuration adopted to evaluate
47 fretting fatigue problems is depicted in Figure 2. In this case, firstly a normal
48 load, P , is applied to cylindrical pads ensuring contact conditions. After that,
49 this normal load is held constant while a tangential oscillatory load is applied
50 to pad in phase with an alternate bulk load, Fb , prescribed on the specimen.
51 Depending on the slip regime, i.e. for gross slip conditions, the tangential load,
52 Q, is replaced by a prescribed tangential displacement δ.
53 The main objectives of this work are (i) to carry out a numerical analysis
54 to design a set of tests which will allow one to study the effects of variable
55 amplitude loading (under partial slip and under mix gross-partial slip regime)
56
57
58
59 8
60
61
62
63
64
65
1
2
3
4
5
6
7
8
P
9 Q(t)
a) b)
10

prescribed loads
normal load, P
11 R tangential load, Q
bulk load, Fb
12
13 Fb(t)

14
15 time
16
17
18
P Q(t)
19
20
21 Figure 2: Standard fretting fatigue conditions: (a) contact configuration, (b) loading history.
22
23 a) b)
24
Load Q Block Load Q Block
25 High-Low Low-High
26
27
28
29
30 nfailure nfailure
31 n1 n2 n'1 n'2
32 D=d1 D=d2 D=d1 D=d2
33
34
35 Figure 3: Schematic representation of the loading configurations: blocks H-L (a), blocks L-H
36 (b).
37
38
39 on fretting fatigue life assessment, and (ii) to verify the effects of variable fret-
40 ting amplitude loading on fretting fatigue life assessment. In this case, fret-
41 ting fatigue simulations were performed varying the amplitude of the tangential
42 load/displacement applied to the pad while all other loads were held the same.
43 In order to exemplify that, let us consider the loading configurations depicted
44 in Figure 3. In Figure 3(a) one has a High-Low (H-L) loading sequence, where,
45 for a given number of cycles n1 and n2 the amplitudes of the tangential load
46 considered in the analysis are Q1 and Q2 , respectively, with Q1 > Q2 . On
47 the other hand, in Figure 3(b), the Low-High (L-H) loading sequence considers
48 first the application of a tangential load of amplitude Q2 followed by the one of
49 amplitude Q1 . In this case, the first block amplitude Q2 is prescribed over n1
50 cycles and this number of cycles is chosen so that the damage generated by the
51 low amplitude block is equal to the one generated in the high amplitude block
52
of the H-L loading sequence. Damage assessments are carried out by means of
53
Miner’s rule. Similarly, the number of cycles n2 is determined by equating the
54
55 damage of the high amplitude block in the L-H loading sequence with the one
56 present in the low amplitude block of the H-L loading sequence.
57
58
59 9
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 a) b) c)
10 Q
 Q  Q
11
12
13
14
15 nfailure nfailure nfailure
16 n1 n2 n1 n2 n1 n2
17 Partial Partial Gross Partial Gross Partial
18 slip slip slip slip slip (wear) slip
19
20
Figure 4: Schematic representation of the three case studies considered in the H-L loading
21 blocks: only partial slip (a), gross slip and partial slip (b), gross slip (including wear effects
22 into the modelling) and partial slip (c).
23
24
25 In this work, besides investigating the effects of variable amplitude loads
26 on fretting fatigue, wear effects by means of material loss were also taken into
27 account in the analyses. However, once the material removal is considerably
28 small under partial slip conditions [56] and apparently does not have great
29 influence on fretting fatigue life estimates [57], here, wear is only considered
30 for gross slip conditions. Figure 4 exemplifies, for the H-L loading sequence,
31 the three different case studies evaluated in this work. In the first type of
32 analysis, Figure 4(a), the H-L loading blocks are applied in partial slip conditions
33 and wear will be discarded in the life estimation methodology for both loading
34
amplitude blocks in this case. In the second case, Figure 4(b), the high block
35
is applied under gross slip regime while the low one remains under partial slip
36
37 conditions. Again, wear will be neglected in the life estimation procedure for
38 both loading blocks. In Figure 4(c), once more gross and partial slip conditions
39 are considered in the high and low blocks, respectively, however, in this case,
40 wear is considered in the life analysis when gross slip takes place. Note that,
41 in terms of loading conditions, the only thing that changes in each of these
42 analyses is the amplitude of the tangential load Q or displacement δ (gross slip
43 conditions). Following the same reasoning, a similar study was conducted for
44 the L-H loading sequence.
45 In order to estimate fretting fatigue life, two different strategies were adopted,
46 one for simulations neglecting wear and the other for simulations taking it into
47 account. In both of them, the TCD, by means of the point method, was used
48 once hot spot approaches are not suitable for contact problems, where strong
49 stress gradients are commonly found nearby contact edges. In this setting,
50 when neglecting wear, fatigue damage processes are accounted for by evaluat-
51 ing stresses and strains at a point L vertically distant from the contact edge,
52 Figure 5. In this case, such nonlocal stress and strain information over time
53
are computed along with the SWT multiaxial fatigue criterion and life can be
54
estimated.
55
56 On the other hand, when wear effects are incorporated in the life estimation
57
58
59 10
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13 Q(t)
14
15
16
P
17 Crack iniciation point
18 Body 1
19
20 (t) L Body 2
(t)
21
22
23
24
25
26 Figure 5: Procedure for estimating fatigue life when wear is neglected.
27
28
29
30
31
32
33
34
Processing phase
35
Results from one fretting Wear input data:
36 cycle simulation:
Pre-processing phase Launch - Wear coefficient (k)
37 Input data through Phython script ABAQUS - Relative slip - Cycle jumping N
38 - Contact pressure
- Geometry and mesh discretization - Coordinates
- Material and contact properties i=1
39
- Loads and boundary conditions Wear computation (Phython script which
40 - Total number of wear cycles (N t ) reads outputs results from ABAQUS):
41 - Initial cumulated damage, 0
 =  1 + 
i += 1
42 

43 Post-processing phase
False
, =  ( ,   ) ( ,  )
44 = 1
45 +
46 N = Nt Contact geometry
True N += N
47 or update modifying
48  = 1 contact nodal positions
(remeshing)
49
50
51
52 Figure 6: Flowchart with the steps for solving the fretting fatigue problem considering wear
53 effects into the life estimation procedure.
54
55
56
57
58
59 11
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
region where damage
10 L is neglected
11
12
13
14
15
16
17
18
19
20 region where damage
21 is considered
22
23
24 Figure 7: Damage accumulation procedure when considering wear.
25
26
27 procedure, the hot-spot point is constantly changing its position due to the
28 material removal and contact stresses redistribution. In this case, life cannot be
29 directly estimated, but increments of Miner’s damage rule might be cumulated
30 for each simulated fretting cycle as follows:
31

m
ΔN
32 Dm (x) = (8)
33 i=1
Nf,i (x)
34
35 where Dm is the total cumulated damage at the mth simulated fretting cycle,
36 ΔN is the jumping factor (Eq. (7)) considered in the wear computations and
37 Nf,i (x) is the estimated life by using any multiaxial fatigue criteria for a given
38 material point with coordinates x. The wear computation strategy adopted in
39 this work is depicted in Figure 6, where a main Python script is in charge of
40
running all the processes, which consist in running each fretting fatigue cycle
41
by means of the Abaqus FE software, perform wear computations with Eq. (7),
42
43 update contacting surfaces and restart the process if needed. Note that, the
44 jumping factor ΔN considered in the wear calculations makes each fretting cy-
45 cle simulated to correspond to ΔN wear cycles. As previously explained, wear
46 computations are performed only for gross slip conditions (high block configu-
47 ration). As can be seen in Figure 6, there are two criteria for terminating the
48 300 fretting wear simulation, i.e. N = Nt or Di = 1. According to Miner’s rule
49 Di = 1 points to material failure. However, as will be seen later, for gross slip
50 conditions, damage computations may reach a plateau, where in this case, the
51 simulation is terminated when a desired number of cycles Nt is reached for the
52 High block. Notice that, in Figure 6, D0 is the initial damage generated from
53 the low block, or zero if the high amplitude block is the first in loading sequence.
54 In order to update contacting surfaces, a remeshing technique was used. In this
55 work, damages computations were carried out at the centroid of the elements
56 surrounding the contact region, Figure 7. Over the fretting cycles, the position
57
58
59 12
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 of the elements centroids are constantly changing due to material loss and mesh
10 modifications, which brings the needing of extrapolating the total cumulated
11 damage from a previous wear step to the current geometry configuration in or-
12 der to apply Eq. (8). From that, failure is observed when Dm = 1 at any point
13 L distant from the contact surface. In this case, the total life experienced by
14 the component is given by m × ΔN . Once the TCD is here used to account ge-
15 ometric features, damages computations are regarded only for points at least L
16 distant from the contact surface, red dots Figure 7. For more details concerning
17 wear and damage computations the reader is referred to [57].
18
19
20 6. Results and discussions
21
22 In this section, three different analyses were conducted to evaluate the fret-
23 ting fatigue phenomenon under variable shear amplitude loading/displacemet.
24 In these three analyses, partial and gross slip conditions were investigated. For
25 some cases the wear modelling was also taken into account. In the next sub-
26 sections we present the results of our numerical analyses for each one of these
27 investigations. The geometrical model considered in this work is depicted in
28 Figure 8. A pad radius R = 70 mm was considered in the analysis. The ma-
29 terial adopted for pads and specimens was the aluminum alloy 7075-T651. Its
30 basic material monotonic and cyclic properties are provided in Table 1. Ac-
31 cording to [58], the stress intensity factor threshold range, ΔKth , and the
32 √ fully
reversed fatigue limit range, Δσ−1 , for this aluminium alloy are 3 MPa m and
33
290 MPa, respectively, which yields a critical distance parameter L = 17 μm.
34
The friction coefficient between the contacting surfaces for a cylinder on plane
35
36 contact configuration may vary from 0.7 to 0.85, according to the tests carried
37 out by the University of Seville fretting fatigue research team [59, 60]. Thus
38 here we adopted μ = 0.75. The local wear coefficient κ is 1.25 × 10−8 mm2 /N
39 according to [61].
40
41 Table 1: Mechanical properties of Al7075-T651 [62].
42 Monotonic properties
43
Young’s modulus E 71 GPa
44
45 Poisson’s modulus ν 0.33
46 Yield strength Sy 503 MPa
47 Ultimate tensile strength Sut 572 MPa
48 Cyclic properties
49 Fatigue strength coeff. σf 1231 MPa
50 Fatigue ductility coeff. εf 0.263
51 Fatigue strength exp. b -0.122
52 Fatigue ductility exp. c -0.806
53
54 As Figure 8 shows, a coarser mesh is used far from the the contact region,
55 whereas, a refined one is considered near the contacting surfaces. In the refined
56
region, linear quadrilateral elements are used with an approximate size of 8.5
57
58
59 13
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 ~8.5m
10 P
Q(t)
11
12
13
14 5.3 mm
15 Fb(t)
16
13 mm
17
18
19
20
21 40 mm
22
23
24 Figure 8: FE model, loading and geometric configurations.
25
26
μm. Outside the refinement zone, the mesh is formed by triangular linear ele-
27
28 ments. For all simulations plane strain elements are considered in the analyses.
29 Multi-point constraints are enforced on the upper surface of the pad so that ro-
30 tation is prevented. Regarding the contact conditions, it was used the Lagrange
31 multiplier approach to define the frictional contact constraints, and the contact
32 nodes located at the specimen were defined as the master ones while the slave
33 nodes were set on the pad.
34
35 6.1. 1st analysis: partial slip conditions in both loading sequences
36 In the analysis of this subsection both H-L and L-H loading sequences were
37 settled under partial slip conditions. Therefore, the effect of surface wear is not
38 accounted for in the life estimation approach here. To carry out the simulations
39 it was considered a normal load P = 300 N/mm, a bulk fatigue load varying
40
between Fb = ±1250 N/mm, a tangential load varying between Q = ±100
41
N/mm in the high block and Q = ±50 N/mm in the low block.
42
43 Figure 9 illustrates the shear stress distribution along the contact surface for
44 each one of the loading blocks. In this case, only the tangential loading history is
45 displayed, however, it is in phase with a bulk load, Fb , and those are prescribed
46 after the application of a static normal load, P . Note that, for the L-H loading
47 sequence, the stress distribution in the high block is not influenced by the first
48 loading block, Figure 9(a). On the other hand, when the low amplitude loading
49 block comes at second, the shear tractions are influenced by the initial high
50 amplitude block, which is not surprising due to the loading history dependency
51 of contact problems. Note as well that, for both cases, only one loading cycle of
52 each loading block is already enough to ensure a stable contact stress profile. In
53 other words, to obtain the stress solutions of each loading block, only one cycle
54 has to be considered.
55 Once there is no need to perform different loading cycles for each loading
56 block, one can consider the loading history depicted in Figure 10 for evaluating
57
58
59 14
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 0.8

10 a) Block L-H

q(x)/fP0
0.8
0

11
q(x)/fP0
q(x)/fp0(Q = Qa)
0

12 Q -0.8
-1 0
x/a
1 q(x)/fP0(Q = -Qa)
-0.8
13 -1 0
x/a
1
1 1
14

q(x)/fP0

q(x)/fP0
0 0
15
16 -1
-1 0 1
-1
-1 0 1
x/a x/a
17
18
19
20 N
21
22 Block H-L
23 b) 1 1
q(x)/fp0(Q = Qa)
24
q(x)/fP0

q(x)/fP0

0 0 q(x)/fP0(Q = -Qa)
25
Q
-1 -1
-1 0 1 -1 0 1
26 x/a x/a

27 0.8 0.8

28
q(x)/fP0

q(x)/fP0

0 0
29
-0.8 -0.8
30 -1 0
x/a
1 -1 0 1
x/a
31
32
33 N
34
35
36
37 Figure 9: Shear tractions profile under variable loadings: (a) L-H and (b) H-L loading se-
38 quences (partial slip conditions).
39
40 a) b)
41 normal load, P High-Low normal load, P Low-High
Applied loads

Applied loads

42 tangential load, Q tangential load, Q


43 bulk load, F bulk load, F
b b
44
45
46 Q Q
1 2
47 Q2 Q1
48 time time
49
50
51
52
53
54
55 Figure 10: Scheme of the loading configuration considered in the FE model in the 1st analysis:
56 (a) H-L and (b) L-H loading sequences.
57
58
59 15
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20 Damage, d2
21 1 0.8 0.6 0.4 0.2 0
22 1,800,000
23
24 1,600,000
25
26 1,400,000
27
28 1,200,000
Esmated Life, N

29
30 1,000,000
31
800,000
32
33
600,000
34
35 400,000
36 H-L
37 200,000 L-H
38
39 0
40 0 0.2 0.4 0.6 0.8 1
41 Damage, d1
42
43
44 Figure 11: Life estimation as a function of damage per block of shear loading (partial slip
45 regime only).
46
47
48
49
50
51
52
53
54
55
56
57
58
59 16
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 the H-L and L-H loading sequences. Denoting d1 as the damage generated
10 by the first loading block, Figure 11 depicts the estimated life as function of
11 the initial imposed damage d1 for the H-L and L-H loading sequences. Note
12 that d1 = 0 or 1 means that there is only one loading block. Note as well
13 that the upper abscissa axis denoted by d2 represents the damage prescribed on
14 the second block, which added to d1 represents material failure (d1 + d2 = 1).
15 Damage computations were performed according to the methodology presented
16 in Section 5.
17 In Figure 11 one can see the influence of the fretting loading history de-
18
pendency in the analisys. Note that, if there was not such loading history
19
dependency, the H-L and L-H curves would intersect each other for a prescribed
20
21 damage of 0.5 (i.e. d1 = d2 = 0.5). One can also see that for the same dam-
22 age combination in the low and in the high amplitude block, the H-L loading
23 sequence leads to shorter lives compared to the L-H one. This behaviour will
24 become even clearer in the next analysis.
25
26 6.2. 2nd analysis: gross and partial slip conditions in the loading sequences
27 In this subsection, a H-L and a L-H loading sequences are also considered.
28 However, in this case, the high amplitude block is set under gross slip conditions
29 while the low one remains under partial slip conditions. As a first approximation
30 and for the sake of comparison, wear is neglected even in the gross slip phase.
31 To carry out numerical experiments, the loading history depicted in Figure 12
32
was considered. A normal load P = 300 N/mm was used. The oscillatory bulk
33
fatigue load prescribed on the specimen varied between Fb = ±1250 N/mm.
34
35 For the low blocks, a tangential load ranging between Q = ±75 N/mm was
36 chosen, whereas, for the high blocks, a tangential displacement δ = ±45 μm
37 was considered.
38 Once more, the simulation of only one cycle of each loading block already
39 provides a stable solution. Figure 13 depicts life expected for each loading
40 sequence (i.e. H-L and L-H) as function of Miner’s damage parameter d1 , in
41 the first block (or in the second block regarding the upper abscissa axis). As
42 can be seen, for the H-L shear loading sequence, the larger is the damage in
43 the first cycle, the shorter is the estimated fatigue life. On the other hand,
44 for the L-H loading sequence, the larger is the damage parameter in the first
45 400 cycle, the longer is the fatigue life expected. Note that in these analyses, wear
46 effects are neglected. Notice as well that, under these circumstances, gross slip
47 conditions yields the most severe stress state. It can also be seen that, for the
48 same damage combination in the low and high loading blocks, the H-L loading
49 sequence leads to predicted lives far below the L-H one, which can be explained
50 by the loading history effect generated by the high amplitude loading block. In
51
the next section, a similar study case is investigated, however, wear effects are
52
taken into account for gross slip regimes.
53
54
55
56
57
58
59 17
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9
10
11 a) High-Low normal load, P b) Low-High normal load, P
12 tangential load, Q tangential load, Q
Applied loads/

Applied loads/
displacements

displacements
bulk load, Fb bulk load, Fb
13 tangential displacement,  tangential displacement, 
14 Displacement Load Load Displacement
15 control control control control
16
17 max max
Qmax Qmax
18 time time
19
20
21
22
23
24
25 Figure 12: 1st and 2nd loading sequences analyses: (a) H-L and (b) L-H.
26
27
28
29
30
31 Damage, d2
1 0.8 0.6 0.4 0.2 0
32
800,000
33
34
700,000
35
36
37 600,000
38
Esmated Life, N

39 500,000
40
41 400,000
42
43 300,000 H-L
44 L-H
45 200,000
46
47 100,000
48
49 0
50 0 0.2 0.4 0.6 0.8 1
51 Damage, d1
52
53
54 Figure 13: Life estimation varying d1 (partial and gross slip regimes).
55
56
57
58
59 18
60
61
62
63
64
65
1
2
3
4
5
6
7
8
a) b)
9
0 250
10 N=0 cycle
11
-0.001 N=50k cycles
12 200
N=100k cycles
13 -0.002

Pressure, MPa
N=140k cycles
14 150
y, mm

15 -0.003
16 100
N=0 cycle
17 -0.004

18 N=50k cycles
50
19 -0.005 N=100k cycles

20 N=140k cycles
-0.006 0
21 -2 -1 0 1 2 -2 -1 0 1 2
22 x, mm x, mm
23
24
25 Figure 14: Variation of the specimen contact profile (a) and contact pressure profile (b)
26 considering wear effects.
27
28
29 6.3. 3rd analysis: gross and partial slip conditions in the loading sequences (con-
30 sidering wear)
31 The analysis conducted in this subsection has the same loading conditions
32 used in the 2nd analysis. However, for both the H-L and L-H loading sequences,
33 the effect of surface wear on the high amplitude block was taken into account. As
34 previously discussed, the wear generated in partial slip regime can be neglected
35 without major consequences in life estimates since the material loss under such
36 conditions is considerably smaller [57].
37 In order to better understand the wear effects on fatigue damage mechanisms
38
under gross slip conditions, it is here first presented a few results considering
39
only the application of the high amplitude block (loading parameters detailed
40
41 in Subsection 6.2). In this setting, the idea is to verify the changes in the con-
42 tacting surfaces profiles, the variation of pressure distribution and the damage
43 accumulation over cycles. Figure 14(a) depicts the specimen contact profile for
44 different number of loading cycles. As can be seen, as wear evolves, the size of
45 the contact zones become larger. As consequence, contact pressure gets smaller
46 once prescribed loads are always the same, Figure 14(b). It can also be seen
47 that the pressure distribution tends to a flat distribution for a high number of
48 cycles.
49 In Figure 15 one has a comparison between the damage evolution as a func-
50 tion of the number of cycles for simulations considering and neglecting the effects
51 of wear into the life estimation methodology. When neglecting wear, damage
52 computations are carried out at a point L vertically distant from the contact
53 trailing edge (Figure 5). When considering wear, the damage results displayed
54 in Figure 15 were obtained finding the maximum cumulated damage for points
55 L vertically distant from the contact surface. Once there are many points L
56 distant from the contact surface, the point which presents the highest damage
57
58
59 19
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 1
10
11 0.9

12
0.8
13
14 0.7
15
16 0.6
Damage, dmax

17
18 0.5
19
20 0.4
21
0.3
22
With wear
23 0.2
24 Without wear
25 0.1
26
27 0
28 0 20,000 40,000 60,000 80,000 100,000 120,000 140,000
29 Number of cycles, N
30
31
32 Figure 15: Damage evolution considering and neglecting wear for the high amplitude shear
33 loading block only.
34 Damage, d2
35 1 0.8 0.6 0.4 0.2 0
36
37 800,000
38 H-L d1=0.51
700,000 d2=0.49
39 L-H
40
600,000
41
Esmated Life, N

42 500,000
43
44 400,000
45
46 300,000
d1=0.49
47 d2=0.51
48 200,000

49
100,000
50
51 0
52 0 0.2 0.4 0.6 0.8 1
53 Damage, d1
54
55
56 Figure 16: Life estimation varying d1 (partial slip and gross slip regime with wear).
57
58
59 20
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 value is chosen. The jumping factor of ΔN = 1000 was considered in the analy-
10 sis accounting wear. Such value was chosen after a convergence study regarding
11 both the damage computation and the surfaces worn profile for different number
12 of fretting cycles.
13 As expected, neglecting wear provides a linear relation between the number
14 of cycles and the cumulated damage. On the other hand, when regarding wear
15 effects by means of material loss, it can be seen that over the first cycles the re-
16 lation between the cumulated damage and the number of cycles is nonlinear and
17 finds a plateau near 0.51 after nearly 140000 cycles. These results exemplifies
18
the importance of considering wear for gross slip regimes.
19
Coming back to the investigation of the variable amplitude loading on fret-
20
21 ting fatigue, the H-L and L-H loading sequences presented in the last subsection
22 are now evaluated but considering wear for the high blocks. Figure 16 depicts
23 the total estimated life for the H-L and L-H loading sequences as a function of
24 the damage generated by the first block, d1 , or the second one d2 .
25 As can be seen in Figure 16, for the H-L loading sequence, finite life is
26 observed only for a limited range of damage in the first block (d1 < 0.51).
27 In this case, besides limiting the damage evolution, the large material loss in
28 the high block also considerably increases the size of the contact zone. Since
29 the normal load is always the same no matter the loading block, contact stress
30 distribution becomes lower. In this setting, despite the number of cycles in the
31 low block (partial slip conditions), stress distributions do not lead to failure.
32 A similar behaviour is observed for the L-H loading sequence, where in this
33 case, failure is only observed for d1 > 0.49 (i.e. d2 < 0.51). In this setting,
34 the high material loss introduced by the high cycle (second block) only causes
35 failure if the damage generated by the low cycle (partial slip conditions) is high
36 enough. Such behaviour can be attributed to the removal of severely damaged
37
area introduced by the gross slip regime found in the high loading block.
38
39
40 7. Conclusions
41
42 The studies presented in this work show that, for the L-H loading sequence,
43 the first loading block does not influence the stress distribution of the second
44 block. However, for the H-L loading sequence, the shear tractions in the second
45 block are influenced by the high amplitude block due to the loading history
46 dependency in contact problems. For the loading condition in the first analysis,
47 i.e. both loading sequences under partial slip regime, it can be seen that, if
48 only the low amplitude block is considered, life is about 9% larger than the
49 one observed when the low loading cycle is prescribed after applying one high
50 loading cycle. In the second analysis, i.e. loading blocks under partial and gross
51 slip regime without including wear effects in the life estimation procedure, the
52 same behaviour is observed. In this case, the loading history effect decreases life
53 estimates in around 70% when the full low amplitude cycle is compared to the
54 low amplitude cycle prescribed after one high amplitude loading cycle. Note as
55 well that, due to the loading history dependency, the H-L loading sequence leads
56
to life estimates far below the L-H one when the same damage combination in
57
58
59 21
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 the low and high loading blocks are regarded. This difference is also noted in
10 the first analysis (only partial slip conditions), but with a subtler difference.
11 Regarding the third analysis, where the loading blocks are the same as the
12 ones in the second analysis, one can see that, including wear in the investigation
13 significantly changes life estimate results. While when neglecting wear gross
14 slip produces the severest stress state, when it is taken into account the damage
15 generated by the high block (gross slip regime) is limited to a certain value. For
16 the H-L loading sequence, the number of cycles in the first block has to be less
17 than a certain value, otherwise, contact stress distribution becomes lower due
18
to the contact profile changes (Figure 14) and the second block under partial
19
slip condition does not yield failure. Similarly, for the L-H loading sequence,
20
21 the damage generated by the low amplitude, which is settled under partial slip
22 conditions, has to be high enough, otherwise, the gross slip phase will have time
23 enough to remove severely damage areas and provides infinite life.
24 To the authors’ best knowledge, it is the first time that fretting fatigue
25 cases under variable amplitude loading are assessed in terms of life estimates
26 also including wear effects. This work was aimed to develop a methodology
27 to predict fretting fatigue life under such conditions. Up to this point, it is
28 not possible to measure the accuracy of the present approach due to the lack
29 of experimental data including the literature, however, in a future work to be
30 published soon, an extensive set of fretting fatigue tests reproducing study cases
31 500 similar to the ones presented here will be carried out. After that, it is expected
32 to answer some open questions such as: (i) Miner’s linear rule is capable of
33 dealing with fretting variable amplitude loading? If not, can a nonlinear version
34 of it be more adequate? (ii) Will be the loading history effect observed in the
35 numerical model also verified in the experiments? (iii) Do wear have such a
36 great influence on fatigue lifetime such as verified in the numerical tests? (iv)
37
With the experimental data in hands, can the use of a variable critical distance
38
parameter improve the quality of the results? Therefore, those are some of
39
40 the points that one expects to shine some light after conducting an extensive
41 experimental campaign to be published soon.
42 In addition, notice that in our simulations we have always moved from the
43 high amplitude blocks to the low ones starting from a valley (Figures 10a and
44 12a). However, stress responses would be different if we had started from the
45 high amplitude block peaks (loading history effect of contact problems). In
46 this setting, we have conducted a preliminary numerical analysis considering
47 this aspect for partial slip conditions. This analysis showed that a load history
48 where the low cycle block comes just after the peak of the high one lead to a
49 life increase of the order of 10%. We expect that for gross slip condition, such
50 difference might be increased, and it is something that we will keep in mind for
51 future investigations.
52
53
54 8. Acknowledgements
55 The authors would like to acknowledge the financial support of TBE (Trans-
56
missoras Brasileira de Energia) in the context of the project entitled “Fadiga
57
58
59 22
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 de Cabos de Alumı́nio Liga (CAL) 1120 e 6201: estudo comparativo, efeito de
10 grampos AGS e de emendas pré-formadas”, This project has been funded in
11 the context of the Research and Development Program of the Brazilian Na-
12 tional Agency of Electric Energy (ANEEL). This study was also financed in
13 part by the Coordenação de Aperfeiçoamento de Pessoal de Nı́vel Superior -
14 Brasil (CAPES) - Finance Code 001.
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 23
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 References
10
11 [1] T. Lindley, Fretting fatigue in engineering alloys, International journal of
12 fatigue 19 (93) (1997) 39–49.
13
14 [2] C. Ruiz, K. C. Chen, Life assessment of dovetail joints between blades and
15 discs in aero-engines, Mechanical Engineering Publications, (1986) 187–194.
16
[3] P. A. Withey, Fatigue failure of the de havilland comet i, Engineering failure
17
analysis 4 (2) (1997) 147–154.
18
19 [4] J. Araújo, D. Nowell, Mixed high low fretting fatigue of ti6al4v: Tests and
20 modelling, Tribology international 42 (9) (2009) 1276–1285.
21
22 [5] C. Azevedo, A. Henriques, A. Pulino Filho, J. Ferreira, J. Araújo, Fret-
23 ting fatigue in overhead conductors: Rig design and failure analysis of a
24 grosbeak aluminium cable steel reinforced conductor, Engineering Failure
25 Analysis 16 (1) (2009) 136–151.
26
27 [6] D. Nowell, D. Dini, D. Hills, Recent developments in the understanding of
28 fretting fatigue, Engineering Fracture Mechanics 73 (2) (2006) 207–222.
29
30 [7] R. Rajasekaran, D. Nowell, Fretting fatigue in dovetail blade roots: Exper-
31 iment and analysis, Tribology International 39 (10) (2006) 1277–1285.
32
33 [8] C. Navarro, S. Muñoz, J. Dominguez, On the use of multiaxial fatigue
34 criteria for fretting fatigue life assessment, International Journal of fatigue
35 30 (1) (2008) 32–44.
36
[9] A. A. Fadel, D. Rosa, L. Murça, J. Fereira, J. Araújo, Effect of high mean
37
tensile stress on the fretting fatigue life of an ibis steel reinforced aluminium
38
39 conductor, International Journal of Fatigue 42 (2012) 24–34.
40 [10] C. Mary, Simulation expérimentale de l’usure du contact aube-disque de
41 compresseur sous sollicitations de fretting, Ph.D. thesis, Ecole Centrale de
42
Lyon (2009).
43
44 [11] A. Cardou, Fretting fatigue under spectrum loading application to overhead
45 electrical conductors: a literature review, Université Laval - Department of
46 Mechanical Engineering (2002) 15.
47
48 [12] A. Palmgren, Die lebensdauer von kugellargern, Zeitshrift des Vereines
49 Duetsher Ingenieure 68 (4) (1924) 339.
50
51 [13] M. A. Miner, Cumulative damage in fatigue, Journal of Applied Mechanics
52 12 (1945) 159–164.
53
54 [14] A. Kantimathi, J. Alic, The effects of periodic high loads on fretting fatigue,
55 Journal of Engineering Materials and Technology 103 (3) (1981) 223–228.
56
57
58
59 24
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 [15] Y. Mutoh, K. Tanaka, M. Kondoh, Fretting fatigue in jis s45c steel un-
10 der two-step block loading: Solid-mechanics, strength of materials, JSME
11 international journal 30 (261) (1987) 386–393.
12
13 [16] Y. Mutoh, K. Tanaka, M. Kondoh, Fretting fatigue in sup9 spring steel
14 under random loading, JSME international journal. Ser. 1, Solid mechanics,
15 strength of materials 32 (2) (1989) 274–281.
16
17 [17] V. Troshchenko, V. Dragan, S. Semenyuk, Fatigue damage accumulation
18 in aluminium and titanium alloys subjected to block program loading un-
19 der conditions of stress concentration and fretting, International journal of
20 fatigue 21 (3) (1999) 271–279.
21
22 [18] R. Cortez, S. Mall, J. R. Calcaterra, Investigation of variable amplitude
23 loading on fretting fatigue behavior of ti–6al–4v, International Journal of
24 Fatigue 21 (7) (1999) 709–717.
25
[19] S. E. Kinyon, D. W. Hoeppner, Spectrum load effects on the fretting be-
26
havior of ti-6al-4v, in: Fretting Fatigue: Current Technology and Practices,
27
28 ASTM International, 2000.
29 [20] Y. Kondo, C. Sakae, M. Kubota, H. Kitahara, K. Yanagihara, Fretting fa-
30 tigue under variable loading below fretting fatigue limit, Fatigue & Fracture
31
of Engineering Materials & Structures 29 (3) (2006) 191–199.
32
33 [21] M. Massingham, P. Irving, The effect of variable amplitude loading on
34 stress distribution within a cylindrical contact subjected to fretting fatigue,
35 Tribology international 39 (10) (2006) 1084–1091.
36
37 [22] C. Gandiolle, S. Fouvry, Fretting fatigue crack propagation rate under vari-
38 able loading conditions, Frattura ed Integrità Strutturale 10 (35) (2016)
39 232–241.
40
41 [23] W. N. Findley, A theory for the effect of mean stress on fatigue of met-
42 als under combined torsion and axial load or bending, no. 6, Engineering
43 Materials Research Laboratory, Division of Engineering, Brown . . . , 1958.
44
45 [24] K. N. Smith, A stress-strain function for the fatigue of metals, Journal of
46 materials 5 (1970) 767–778.
47
48 [25] L. Susmel, P. Lazzarin, A bi-parametric wöhler curve for high cycle mul-
49 tiaxial fatigue assessment, Fatigue & Fracture of Engineering Materials &
50 Structures 25 (1) (2002) 63–78.
51
[26] B. Crossland, Effect of large hydrostatic pressures on the torsional fatigue
52
53 strength of an alloy steel, in: Proc. Int. Conf. on Fatigue of Metals, Vol.
54 138, Institution of Mechanical Engineers London, 1956, pp. 12–12.
55 600 [27] G. Sines, Failure of materials under combined repeated stresses with super-
56
imposed static stresses, Tech. rep., California. Univ., Los Angeles (1955).
57
58
59 25
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 [28] T. Hattori, M. Nakamura, T. Watanabe, Simulation of fretting-fatigue life
10 by using stress-singularity parameters and fracture mechanics, Tribology
11 international 36 (2) (2003) 87–97.
12
13 [29] J. Ding, D. Houghton, E. Williams, S. Leen, Simple parameters to predict
14 effect of surface damage on fretting fatigue, International Journal of Fatigue
15 33 (3) (2011) 332–342.
16
17 [30] J. Lemaitre, A continuous damage mechanics model for ductile fracture,
18 Journal of engineering materials and technology 107 (1) (1985) 83–89.
19
[31] J.-L. Chaboche, Continuum damage mechanics: Part ii—damage growth,
20
crack initiation, and crack growth, Journal of applied mechanics 55 (1)
21
(1988) 65–72.
22
23 [32] D. Socie, Multiaxial fatigue damage models, Journal of Engineering Mate-
24 rials and Technology 109 (4) (1987) 293–298.
25
26 [33] N. A. Bhatti, M. A. Wahab, Fretting fatigue crack nucleation: A review,
27 Tribology International 121 (2018) 121–138.
28
29 [34] H. Neuber, Forschg ing-wes, Bd 7 (1936) 271.
30
31 [35] R. Peterson, Methods of correlating data from fatigue tests of stress concen-
32 tration specimens, Stephen Timoshenko Anniversary Volume (1938) 179.
33 [36] J. M. Whitney, R. Nuismer, Stress fracture criteria for laminated compos-
34
ites containing stress concentrations, Journal of composite materials 8 (3)
35
(1974) 253–265.
36
37 [37] K. Tanaka, Engineering formulae for fatigue strength reduction due to
38 crack-like notches, International Journal of Fracture 22 (2) (1983) R39–
39 R46.
40
41 [38] P. Lazzarin, R. Tovo, G. Meneghetti, Fatigue crack initiation and propaga-
42 tion phases near notches in metals with low notch sensitivity, International
43 Journal of Fatigue 19 (8-9) (1997) 647–657.
44
45 [39] D. Taylor, Geometrical effects in fatigue: a unifying theoretical model,
46 International Journal of Fatigue 21 (5) (1999) 413–420.
47
48 [40] D. Taylor, G. Wang, The validation of some methods of notch fatigue
49 analysis, Fatigue & fracture of engineering materials & structures 23 (5)
50 (2000) 387–394.
51
[41] J. Araújo, L. Susmel, M. Pires, F. Castro, A multiaxial stress-based critical
52
53 distance methodology to estimate fretting fatigue life, Tribology Interna-
54 tional 108 (2017) 2–6.
55 [42] D. Taylor, The theory of critical distances, Engineering Fracture Mechanics
56
75 (7) (2008) 1696–1705.
57
58
59 26
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 [43] L. Susmel, D. Taylor, On the use of the theory of critical distances to predict
10 static failures in ductile metallic materials containing different geometrical
11 features, Engineering Fracture Mechanics 75 (15) (2008) 4410–4421.
12
13 [44] J. Archard, Contact and rubbing of flat surfaces, Journal of applied physics
14 24 (8) (1953) 981–988.
15
16 [45] S. Fouvry, T. Liskiewicz, P. Kapsa, S. Hannel, E. Sauger, An energy descrip-
17 tion of wear mechanisms and its applications to oscillating sliding contacts,
18 Wear 255 (1-6) (2003) 287–298.
19
[46] I. McColl, J. Ding, S. Leen, Finite element simulation and experimental
20
validation of fretting wear, Wear 256 (11-12) (2004) 1114–1127.
21
22 [47] J. Ding, S. Leen, I. McColl, The effect of slip regime on fretting wear-
23 induced stress evolution, International journal of fatigue 26 (5) (2004) 521–
24 531.
25
26 [48] J. Madge, S. Leen, I. McColl, P. Shipway, Contact-evolution based predic-
27 tion of fretting fatigue life: effect of slip amplitude, Wear 262 (9-10) (2007)
28 1159–1170.
29
30 [49] A. Cruzado, M. Urchegui, X. Gómez, Finite element modeling and experi-
31 mental validation of fretting wear scars in thin steel wires, Wear 289 (2012)
32 26–38.
33
34 [50] S. Garcin, S. Fouvry, S. Heredia, A fem fretting map modeling: Effect of
35 surface wear on crack nucleation, Wear 330 (2015) 145–159.
36
[51] V. Hegadekatte, N. Huber, O. Kraft, Finite element based simulation of
37
38 dry sliding wear, Modelling and Simulation in Materials Science and Engi-
39 neering 13 (1) (2004) 57.
40 [52] E. Bortoleto, A. Rovani, V. Seriacopi, F. Profito, D. Zachariadis,
41
I. Machado, A. Sinatora, R. Souza, Experimental and numerical analysis
42
of dry contact in the pin on disc test, Wear 301 (1-2) (2013) 19–26.
43
44 [53] H. Lin, Simulation of wear in turbocharger wastegates (2016).
45
46 [54] J. Madge, S. Leen, P. Shipway, The critical role of fretting wear in the
47 analysis of fretting fatigue, Wear 263 (1-6) (2007) 542–551.
48
49 [55] A. M. Tobi, J. Ding, G. Bandak, S. Leen, P. Shipway, A study on the
50 interaction between fretting wear and cyclic plasticity for ti–6al–4v, Wear
51 267 (1-4) (2009) 270–282.
52
53 [56] O. Vingsbo, S. Söderberg, On fretting maps, Wear 126 (2) (1988) 131–147.
54 [57] R. Cardoso, T. Doca, D. Néron, S. Pommier, J. Araújo, Wear numerical as-
55 sessment for partial slip fretting fatigue conditions, Tribology International
56
136 (2019) 508–523.
57
58
59 27
60
61
62
63
64
65
1
2
3
4
5
6
7
8
9 [58] M. Benedetti, V. Fontanari, C. Santus, M. Bandini, Notch fatigue be-
10 haviour of shot peened high-strength aluminium alloys: Experiments and
11 predictions using a critical distance method, International Journal of Fa-
12 tigue 32 (10) (2010) 1600–1611.
13
14 [59] V. Martı́n, J. Vázquez, C. Navarro, J. Domı́nguez, Fretting-fatigue analysis
15 of shot-peened al 7075-t651 test specimens, Metals 9 (5) (2019) 586.
16
17 [60] L. Bohórquez, J. Vázquez, C. Navarro, J. Domı́nguez, On the prediction
18 of the crack initiation path in fretting fatigue, Theoretical and Applied
19 Fracture Mechanics 99 (2019) 140–146.
20
21 [61] G. Jordano, C. Navarro, J. Vázquez, J. Domı́nguez, Measuring wear in a
22 fretting test with a confocal microscope, in: Key Engineering Materials,
23 Vol. 774, Trans Tech Publ, 2018, pp. 461–466.
24
[62] J. Vázquez, C. Navarro, J. Domı́nguez, Experimental results in fretting
25
fatigue with shot and laser peened al 7075-t651 specimens, International
26
Journal of Fatigue 40 (2012) 143–153.
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 28
60
61
62
63
64
65

You might also like