You are on page 1of 10

Wear 263 (2007) 542–551

The critical role of fretting wear in the analysis of fretting fatigue


J.J. Madge ∗ , S.B. Leen, P.H. Shipway
School of Mechanical, Materials and Manufacturing Engineering, University Technology Centre in Gas Turbine Transmission Systems,
University of Nottingham, University Park, Nottingham NG7 2RD, UK
Received 15 August 2006; received in revised form 26 October 2006; accepted 1 November 2006
Available online 26 March 2007

Abstract
A finite element based approach is used to predict the effect of fretting wear on fretting fatigue life in the commonly used ‘rounded edge punch
on flat’ laboratory geometry. The method predicts and hence explains experimentally observed phenomena, such as cracking location under the
contact and dependence of fretting fatigue life on slip amplitude, which were not previously predictable. The results from the model are compared
against available test data across a range of slip amplitudes. The model predicts a significant effect of slip amplitude on fretting fatigue life and,
in particular, the well-known phenomenon of a critical range of slip amplitude for minimum life. The sensitivity of life and cracking location to
the assumed value of wear coefficient on life and cracking location is studied for a partial slip case and it is shown (i) that the location of cracking
switches from the contact edge to the stick–slip interface with increasing wear coefficient and (ii) that there is an associated critical value of wear
coefficient which gives a maximum fretting fatigue life. For gross sliding cases, failure is always predicted at the contact edge. The work shows
that inclusion of the effects of wear on fatigue damage accumulation is necessary for prediction of fretting fatigue life and cracking location.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Fretting wear; Fretting fatigue; Rounded punch on flat; Critical plane; Contact slip

1. Introduction life prediction methods will aid further optimisation of the


design of such components. Leen et al. [6] studied the per-
Many engineering assemblies contain contacting compo- formance of a laboratory-scale spline coupling under different
nents which are nominally fixed with respect to each other, yet combinations of torque, axial and rotating bending moment
in reality experience micro slip due to the loads imposed across loading. While plain fatigue failure occurred under torque over-
the contact interface. The surface damage that results from this load, fretting fatigue cracking occurred under large bending
‘fretting’ condition is known to be a key factor in the fatigue moment overloads, with the cracks occurring within the length
behaviour of these components, and can cause dramatic reduc- of engagement contact region (Fig. 1).
tions in fatigue life. The fretting process has been found to be Similar trends have been observed in ‘punch on flat’ fret-
dependent on a large number of variables [1]. Dobromirski [2] ting fatigue tests. In studies by Jin and Mall [7] and Nakazawa
concluded that the variables of coefficient of friction (COF), nor- et al. [8], failure was observed well within the contact region
mal load and slip amplitude are of primary importance. However, under certain loading conditions. Similarly, Hertzian contact
even these ‘primary’ variables are found to be interdependent. conditions have frequently produced cracking under the con-
For example, the COF is known to affect the slip distribution tact [4]; this can occur at either the contact edge or the stick–slip
[3], whilst COF is found to be dependent on normal load; fur- interface. The effect of slip amplitude has been shown to be
thermore, COF is found to change with the number of fretting significant in experimental studies, e.g. [4,9]. A second key
cycles, e.g. [4,5]. phenomenon in fretting fatigue is the occurrence of a mini-
Spline couplings and turbine root dovetails are notable exam- mum fatigue life corresponding to a certain critical range of
ples of aeroengine components where fretting occurs. Accurate slip amplitudes, normally associated with the transition from
partial slip to small amplitude gross sliding. Both of these phe-
nomena, viz. failure within the contact region away from the
∗ Corresponding author. Tel.: +44 115 8467685. contact edge, and dependence of fatigue life on slip ampli-
E-mail address: eaxjm1@nottingham.ac.uk (J.J. Madge). tude, are difficult to explain and capture using existing fretting

0043-1648/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2006.11.021
J.J. Madge et al. / Wear 263 (2007) 542–551 543

Nomenclature
b fatigue strength exponent; also specimen width
c fatigue ductility exponent
E Young’s modulus
h wear depth
h incremental wear depth
H hardness
k bulk wear coefficient
kl local wear coefficient
kcrit critical wear coefficient value
K nondimensional wear coefficient
ni number of cycles at load cycle i
N number of fretting cycles
N cycle jump size
N* number of fretting cycles analysed within FE
Nf number of cycles to failure
Nfi number of cycles to failure at load cycle i
Fig. 1. Fretting fatigue cracking location for a laboratory scaled spline coupling
P normal load applied to fretting pad under combined torque, axial and bending moment overload [6]. (N.B. z = axial
R fatigue stress load ratio position, al = total axial length of engagement).
S sliding distance
t time gave good experimental agreement over the range of stresses
V wear volume tested.
W accumulated damage fraction In this paper, the multiaxial fatigue life prediction approach
x position across contact face of Madge et al. [10] with wear effects included, is applied to the
y depth into specimen surface from contact surface ‘rounded-edge punch-on-flat’ geometry. The method is based on
a modified version of the Archard wear equation which is used to
Greek letters predict the effect of material removal due to wear, by incremen-
δ relative slip at contact surface over one fretting tally updating the finite element mesh geometry. A critical plane
cycle implementation of the Smith–Watson–Topper (SWT) parame-
δapp maximum relative displacement applied to pad in ter [12] is combined with a damage accumulation approach to
one fretting cycle predict fatigue damage during material removal and hence fail-
δapp (t) applied displacement at time t ure. The results are validated against available test data across
εf fatigue ductility coefficient a range of slip amplitudes. The effect of slip amplitude and the
ε strain range over one cycle (transformed to plane) effect of changes in wear coefficient are investigated in order to
σf fatigue strength coefficient explain the aforementioned experimental phenomena.
σ max maximum applied fatigue stress
σ min minimum applied fatigue stress
2. FE model description
σ(t) applied fatigue stress at time t
τ time during one cycle jump
Fig. 3 shows the specimen geometry modelled, based on that
used by Jin and Mall [7]. A pair of flat fretting pads with a nom-
inal contact width of 4.45 mm and edge blend radii of 2.54 mm
analysis approaches. Conventional approaches have not been are held in contact with a flat, uniaxially loaded fatigue speci-
able to predict the effect of slip amplitude, particularly the
beneficial effect of gross sliding (Fig. 2). These approaches
also generally predict failure at the trailing contact edge. How-
ever, Madge et al. [10] have recently presented a critical plane
methodology for fretting fatigue life prediction that accounts
for the material removal effect of wear by use of the Archard
wear model. When applied to the Hertzian contact geometry,
the approach was found to predict the measured dependence
of fretting fatigue life on slip amplitude more accurately than
using a similar critical plane technique which neglects wear
(Fig. 2). Hattori and Watanabe [11] have developed a stress
intensity factor based fretting fatigue lifing methodology cou- Fig. 2. Fretting fatigue life prediction by Madge et al. [10] compared to exper-
pled with the Archard model to account for wear, this also imental values by Jin and Mall [4] (Hertzian geometry).
544 J.J. Madge et al. / Wear 263 (2007) 542–551

Fig. 5. Loading history applied to the FE model.

edges of contact to improve the stress resolution in this region


Fig. 3. (a) Experimental schematic; (b) fretting fatigue specimen; (c) fretting as very high stress gradients can be expected in this region. The
pad; after [7]. smallest element in the model is around 14 ␮m × 9 ␮m.
Due to the conforming nature of the contact and the relatively
men. The material used in the experiments by Jin and Mall [7] abrupt edge of contact, the influence of pad misalignment due
is a dual phase Ti-6Al-4V alloy which consists of α (hexagonal to rig compliance was found to be an important factor. In order
close packed) and β body centred cubic phases). The material to represent the flexibility of the experimental rig, the pad is
was solution heat treated at 935 ◦ C for 1.75 h, cooled in air and modelled to a height of 11 mm away from the contact surface.
annealed at 700 ◦ C for 2 h in vacuum and cooled in argon. The Following the work of Sabelkin and Mall [13], the effect of rig
material is modelled as linear elastic with a Young’s modulus of compliance has been incorporated by setting the elastic modulus
126 GPa and a Poisson’s ratio of 0.32. Table 1 shows the fatigue of the extended section of the pad to a value of 55 GPa. The top
properties used in the fatigue lifing model. surface of the pad is prescribed to remain plane. The ‘contact
A symmetry plane parallel to the specimen axis is employed pairs’ option [14] was used in conjunction with the Lagrange
in the FE model so that only one pad needs to be modelled friction algorithm to strictly enforce exact sticking behaviour
(Fig. 4). The FE mesh created has refinement graded towards the when the resultant shear stress is less than the critical shear
stress. A COF of 1.05 is taken as representative, based on the
Table 1 approximate stabilised dry sliding COF measured by Jin and
Fatigue properties for Ti-6Al-4V Mall [7].
σf (MPa) b εf c In the first analysis step a normal load P of 208 N/mm (equiv-
alent to 1334 N for a 6.4 mm wide specimen) is applied to the
2030 −0.104 0.841 −0.688
fretting pad. In the next step the specimen is loaded by a cyclic
The above data corresponds to a definition of crack initiation as the development fatigue load σ(t) with a maximum value σ max of 550 MPa and a
of a 1 mm crack from Dowling [25].
stress ratio R of 0.03. The pad is also loaded with a prescribed
maximum displacement δapp , which is varied for each simula-
tion to cover a range of slip amplitudes including partial slip
through to gross sliding conditions. δapp (t) and σ(t) are then
applied cyclically in the FE model. Fig. 5 shows the load cycle
history for the first two load cycles.

3. Wear modelling

3.1. Wear model

The wear simulation technique used here is based on that


implemented and validated by McColl et al. [5] for Super-
CMV high strength steel fretting wear tests. The specific method
employed here is that described in Madge et al. [10] for Ti-
6Al-4V in a Hertzian fretting fatigue configuration, which
demonstrated successful validation against fatigue life test data
across a range of slip amplitudes. The material removal approach
is based on the well-known Archard equation:
V P
Fig. 4. The FE mesh used for analyses. Shaded section shows ‘compliant sec- =K (1)
tion’ used to represent rig compliance. S H
J.J. Madge et al. / Wear 263 (2007) 542–551 545

where V is the volume wear, S is the sliding distance, K is the


wear coefficient, P is the normal load and H is the hardness of
the material.
At each point along the contact surface, Eq. (1) can be written
as
h
= kl p (2)
S
where h is the wear depth (mm), kl is the dimensional local wear
coefficient defined as the wear per unit local slip per unit local
contact pressure, and p is the contact pressure. Therefore, the Fig. 6. Schematic showing the wear modelling dealt with in the ABAQUS
incremental wear depth for a node experiencing an incremental FE code, stress–strain results are then fed into the post processing software
slip δ and pressure p at time t can be expressed as: to calculate critical plane SWT values and accumulated damage.

Δh(x, t) = kl p(x, t)δ(x, t) (3) In summary, Eq. (6) is programmed into the UMESHMOTION
where h(x,t), p(x,t) and δ(x,t) are the incremental wear depth, subroutine within ABAQUS to effect incremental modification
contact pressure and relative slip at point x and time t, respec- of the contact geometry to model the material removal effect of
tively. wear. Each FE fretting cycle was analysed in 100 increments,
Direct calculation of the desired wear coefficient, kl , requires resulting in the geometry being updated 100 times during each
knowledge of the local contact pressures and slips. As these N cycle jump.
are not readily measurable, a modified form of Eq. (3) is used Fig. 6 summarises the approach employed. The evolution of
to determine a bulk wear coefficient, k, for which a value of contact pressures, slips, stresses and damage parameters can then
2.75 × 10−8 MPa−1 has been estimated as representative of the be presented in terms of a ‘scalable’ number of cycles, N* , cor-
Ti-6Al-4V fretting contact pair of the present work (see [10] for responding to the reference wear coefficient kref , which can be
details), using measured wear scar data from Magaziner et al. easily converted to specific numbers of cycles, N corresponding
[15]. to arbitrary wear coefficients, k, as follows:
kref ∗
3.2. Variation of k N= N (7)
k

For computational efficiency, a cycle jump technique is In the present study, a kref value of 4 × 10−5 MPa−1 has been
employed, as described in [10], where the assumption is made employed to expedite the wear simulations and thus reduce com-
that wear is constant over a certain number of cycles. By multi- putational time.
plying the incremental wear by a cycle jumping factor N, one
FE cycle simulation is used to model the effect of N actual 4. Fatigue damage model
wear cycles. Eq. (3) thus becomes:
4.1. Multiaxial fatigue methodology
h(x, τ) = kNp(x, τ)δ(x, τ) (4)
The Smith–Watson–Topper parameter [12] is based upon a
where τ is the time within one ‘cycle’ corresponding to N wear
combined high- and low-cycle strain–life equation, with a maxi-
cycles and the bulk wear coefficient k is employed in place of
mum stress term to account for the mean stress effect, as follows:
kl . The modified Archard equation given in Eq. (4) shows that
since k and N appear as a product on the right-hand side, this ε σ2
product can be represented as one constant kref which can be SWT = σmax = f (2Nf )2b + σf εf (2Nf )b+c (8)
2 E
treated as a reference wear coefficient, as follows:
where σ max is the peak normal stress on a plane, ε the maxi-
kN ≡ kref (5) mum normal strain range on the same plane within one fatigue
cycle, σf is the fatigue strength coefficient, E is Young’s mod-
so that Eq. (4) can now be written as: ulus, Nf is the number of cycles to crack initiation, b is the
h(x, τ) = kref p(x, τ)δ(x, τ) (6) fatigue strength exponent, εf is the fatigue ductility coefficient
and c is the fatigue ductility exponent. For most fatigue life sit-
Subject to the assumption that the solution is approximately uations, Fatemi and Socie [16] have proposed the use of one
independent of the cycle jump size N (which has been found of two multiaxial fatigue damage criteria, one for shear crack-
to be valid by the authors for the range of cases considered in ing failure modes and one for tensile cracking failure modes.
this paper), the effect of varying the wear coefficient k, within Such parameters should take account of both crack initiation
a reasonable range of values, can be investigated by varying the and propagation. For shear cracking failure modes, the param-
cycle jump size, as a post-processing exercise to a simulation eter should include shear strain amplitude for crack initiation,
using a fixed value of kref . The implementation of Eq. (6) within and maximum normal stress on the plane of maximum shear
ABAQUS has been described previously by Madge et al. [10]. strain amplitude for crack propagation. For tensile cracking, the
546 J.J. Madge et al. / Wear 263 (2007) 542–551

parameter should include maximum principal strain amplitude and spatial position (element), which in turn is used with Eq. (8)
for initiation, and maximum normal stress on the plane of max- to furnish a number of cycles to failure, Nf for each fretting cycle.
imum principal strain amplitude for propagation. Thus, Socie The FE-implementation (without wear effects) was successfully
[17] suggests that knowledge of the cracking behaviour for the validated by comparing the predicted lives for a Hertzian fretting
material and loading state of interest allows selection of the most contact against corresponding FE-based predictions by Sum et
appropriate fatigue criterion. On this basis, Araujo and Nowell al. [23], and Araujo and Nowell [18] for partial slip conditions.
[18] have suggested that the SWT parameter is appropriate for
fretting fatigue prediction of Ti-6Al-4V. 4.2. Damage accumulation model
The critical plane approach [19] is based on physical obser-
vations that fatigue cracks initiate and grow within a material on A cumulative damage approach is required to take account of
certain planes, where the growth and orientation depends on the the evolution of stress and strain cycles with increasing number
normal stresses and strains on these planes. Dominant fatigue of fretting cycles due to material removal. Damage accumulation
damage occurs on these planes, which are called critical planes. is generally a non-linear interaction concerning both the nature
The approach has evolved from the work of Brown and Miller of each individual loading cycle and the order in which dif-
on multiaxial fatigue [20] who first suggested the combined use ferent cycles are applied [24]. The most commonly used fatigue
of maximum shear strain range and tensile strain normal to the damage accumulation model is the Miner–Palmgren (M-P) rule,
plane of maximum shear and is favoured due to its basis on phys- which assumes a simple linear accumulation of damage, ignor-
ical interpretation of the mechanisms of fatigue crack growth. ing the effect of order of the loading cycles, as follows:
The approach of Fatemi and Socie [16], i.e. using normal stress N
instead of strain as the second parameter, overcomes limita-  ni
W= (9)
tions of the Brown and Miller [20] approach with respect to the Nfi
i=1
increased damage associated with non-proportional or out-of-
phase loading attributed to the extra cyclic hardening developed where W is the accumulated damage, ni is the number of cycles
due to rotation of the principal axes. Practically, the critical plane experienced at loading cycle i, Nfi is the critical plane SWT
approach amounts to identification of the orientation of the plane predicted number of cycles to failure at loading cycle i (found
which maximises the relevant fatigue damage parameter, at each using Eq. (8)) and N is the number of different load cycles expe-
spatial location of interest, covering the complete loading cycle. rienced. In the case of fretting wear, each cycle differs from
Szolwinski and Farris [19] were the first to suggest the criti- the next, so that the cumulative damage for each individual
cal plane method (with the SWT parameter) as a technique to fretting cycle should be considered. When the cycle jumping
deal with the multiaxial stress and strain cycles in fretting con- technique described above is combined with Eq. (9), the damage
tacts. Numerous other authors have since used the critical plane rule becomes:
N
t /N
SWT parameter for fretting fatigue of Ti-6Al-4V, e.g. [20,21]. N
A study by Lykins et al. [22] which compared the performance W= (10)
Nfi
i=1
of SWT against a maximum shear stress range (MSSR) damage
parameter for prediction of Ti-6Al-4V fretting fatigue life, crack Given the adoption of a critical plane approach to calculate
location and crack orientation, without including the effects of fatigue damage, it is necessary to make an assumption about how
wear, concluded that both parameters were equally accurate for damage on a given plane interacts with damage on other planes
life and crack location prediction, but that MSSR was superior from cycle to cycle. One approach is to calculate the critical
for crack orientation prediction. Although SWT is employed in plane SWT for each wear increment and simply accumulate
this paper, primarily as a vehicle to illustrate the impact of fret- this value from wear increment to increment, irrespective of the
ting wear modelling on life predictions, it would not be difficult fact that the associated orientations of the critical plane may
to supplant SWT with MSSR in the present methodology, for change with wear, i.e. assume the critical plane SWT damage
example. to have an isotropic damaging effect. An alternative approach is
The FE implementation of the critical plane SWT parame- to calculate SWT values for all plane orientations for each wear
ter here follows the method described by Sum et al. [23]. In increment and to accumulate the incremental damage on each
practice this is achieved by transforming the time histories of plane. The latter approach is significantly more computationally
element centroidal stresses and strain ranges onto planes at 5◦ intensive, requiring storage of damage on 36 different planes
intervals over a 180◦ range using the two-dimensional trans- at each element centroid (note that a rectangular grid of 1700
formation (Mohr’s circle) equations for stress and strain. The elements is monitored for damage here) for each wear increment
maximum normal stress σ max with respect to time, and the cor- (typically about N* = 100 wear increments). In this study, the
responding strain range ε are determined for each of the 36 former assumption is adopted, so that damage is treated as a
planes in each element. ε is the difference between the max- scalar quantity for each element. Failure is assumed to have
imum and minimum values of strain normal to the candidate occurred when W reaches a value of 1 at any point.
plane over the complete loading cycle. Thus an array of SWT Due to the use of adaptive meshing to update the mesh for
values is obtained for each candidate plane in each element for wear-induced removal of material, specific elements and nodes
each fretting cycle analysed. These values are then employed to are no longer linked uniquely to actual material points through-
establish the maximum SWT with respect to plane orientation out the analysis. Consequently, a ‘material point mesh’ (MPM)
J.J. Madge et al. / Wear 263 (2007) 542–551 547

is created as the global reference for damage accumulation; the


nodes of the MPM have fixed coordinates throughout the analy-
sis. Cyclic damage is calculated at the centroid of each element,
and linearly interpolated back to the MPM for accumulation. In
this way, nodes on the MPM corresponding to removed material
(due to wear) do not accumulate any further fatigue damage.

5. Results and discussion

5.1. Effect of slip amplitude

The predicted effect of slip amplitude on fretting fatigue life


is shown in Fig. 7, along with experimental test data for the
same conditions as Jin and Mall [7]. The predictions show a
reasonable correlation with the relatively sparse experimental
Fig. 8. Contact pressure distributions for points of maximum and minimum
results. The effect of slip amplitude on fatigue life is similar fatigue stress (δapp = 40 ␮m).
to that reported by Madge et al. [10] for the Hertzian fretting
fatigue contact and generally accepted as representative of fret- tial slip case. Whether or not the experimental rig displays this
ting fatigue [3]. Specifically, a minimum in life is predicted at behaviour is not clear; however, it is interesting to note that the
the critical range of δapp amplitudes corresponding to the transi- pad rotates in this manner, even though it is modelled with an
tion from partial slip to gross sliding conditions (approximately encastre boundary condition and a compliant section modulus
50 ␮m ≤ δapp ≤ 85 ␮m), with comparatively long lives predicted (55 GPa) comparable to some engineering metals. This suggests
as δapp is either reduced or increased outside of this critical range. that reducing the system stiffness has a detrimental effect on life
At low (partial) slip amplitudes, the shear traction has a negligi- by compromising contact alignment.
ble effect so that comparatively long lives are predicted. As δapp
is increased from zero to the critical range, i.e. for δapp < 50 ␮m, 5.3. Gross sliding results
the predicted life reduces as the shear force, and in turn the
local fatigue stresses increase. At the onset of gross sliding, the Fig. 9 shows the effect of wear on the contact pressure dis-
shear tractions have saturated and a minimum in life is observed. tribution for a gross sliding case (δapp = 90 ␮m). The two salient
However, as δapp is increased still further the effect of wear is to predicted effects of wear in this case are (i) the significant reduc-
increase fatigue life, as discussed in more detail below. tion in the edge of contact pressure peak and (ii) the significant
increase in the contact area (and hence the load distribution
5.2. Effect of rig compliance area) by approximately 50%. Fig. 10 shows the evolution of
critical plane SWT distribution across the contact surface with
The presence of a complaint section in the model creates an increasing numbers of fretting cycles. For the unworn case and
appreciable amount of load transfer throughout the fretting cycle for early numbers of cycles, e.g. N* = 10 (which corresponds to
such that the pad is able to rotate under shear loading. At larger N = 60,000 cycles for k = 2.75 × 10−8 MPa−1 ), there is a pro-
slip amplitudes, the shear force is large enough to cause suffi- nounced peak at the left-hand edge of contact (viz. the trailing
cient deflection for the trailing edge of the pad to lift free, whilst edge at the instant of maximum substrate fatigue stress). How-
the leading edge is forced further into contact. This is illustrated ever, it is clear that the predicted effect of wear here is to cause a
in Fig. 8, which shows the pressure distributions at maximum
and minimum substrate fatigue stress for the δapp = 40 ␮m par-

Fig. 9. Evolution of contact pressure distribution for the δapp = 90 ␮m gross


Fig. 7. Prediction of fretting fatigue life at different slip amplitudes as compared sliding case. Note: for a typical value of k = 2.75 × 10−8 MPa−1 , N* = 100 cor-
to experimental results of Jin and Mall [7]. responds to N = 600,000 cycles.
548 J.J. Madge et al. / Wear 263 (2007) 542–551

Fig. 10. Evolution of SWT distribution across contact surface for the Fig. 12. The N* = 100 slip distribution for the δapp = 40 ␮m partial slip case.
δapp = 90 ␮m gross sliding case. Note: for a typical value of k = 2.75 ×
10−8 MPa−1 , N* = 70 corresponds to N = 420,000 cycles. comparing Figs. 11 and 12 (which shows the slip distribution
for the δapp = 40 ␮m case) that this pressure peak in fact devel-
dramatic reduction in the SWT value at this location. The asso- ops at the stick–slip interface. The discontinuity between wear
ciated rapid drop in cyclic fatigue damage is responsible for the in the slip region and no wear in the stick region generates a
increase in life observed at higher slip amplitudes in Fig. 7. Fail- geometric and loading discontinuity which leads to a stress con-
ure is predicted at the trailing edge of contact for all gross sliding centration, with load being gradually concentrated on the stick
cases. region.
Fig. 13 shows the predicted evolution of the SWT param-
5.4. Partial slip results eter across the contact surface for the latter partial slip case.
Again, consistent with the attenuation in edge-of-contact pres-
Fig. 11 shows the predicted evolution of contact pressure sure peak (and with the gross sliding case) during the early
distribution with increasing number of wear cycles for the cycles, an initial attenuation in edge-of-contact SWT value is
δapp = 40 ␮m partial slip case. As for the gross sliding case, predicted. However, the subsequent predicted development of
wear is predicted to result in an initial beneficial effect in the pressure peak at the stick–slip interface leads directly to
reducing the N* = 0 edge of contact pressure peak. Thus, for an associated predicted increase in SWT value. The implica-
k = 2.75 × 10−8 MPa−1 , for example, Fig. 11 shows that the tion of this is that the issue of cracking location is predicted
peak pressure value has dropped to about one quarter of its to become more complex for partial slip cases than for gross
initial value after about 132,000 cycles (N* = 22). For higher sliding; specifically, for partial slip cases, it is predicted that
wear coefficients, this number of cycles will reduce proportion- failure can occur at either the contact edge or the stick–slip
ally, e.g. for k = 2.75 × 10−7 MPa−1 , the same attenuation will interface, depending on the competition between rate of fatigue
have occurred within about 13,200 cycles. However, a critical damage accumulation and wear-induced evolution of fretting
phenomenon of the partial slip case, as compared to the gross stresses across the contact, as discussed below. A similar phe-
sliding case, is the development of a pressure peak well away
from the contact edges as wear advances. It can be seen by

Fig. 11. Evolution of contact pressure distribution with wear for the δapp = 40 ␮m Fig. 13. Evolution of SWT distribution across the contact surface for a partial
partial slip case. Note: for a typical value of k = 2.75 × 10−8 MPa−1 , N* = 100 slip case, δapp = 40 ␮m. Note: for a typical value of k = 2.75 × 10−8 MPa−1 ,
corresponds to 600,000 cycles. N* = 100 relates to 600,000 cycles.
J.J. Madge et al. / Wear 263 (2007) 542–551 549

nomenon has been previously predicted for the Hertzian fretting


problem [10]. Even though wear is predicted to give an atten-
uation of edge of contact SWT peak for partial slip cases (as
for gross sliding), this is offset by an accentuation of SWT at
the stick–slip interface, which cannot be predicted by models
which do not include the effects of material removal. However,
the wear effect on partial slip life is somewhat secondary to the
effect of the shear traction increase, so that ‘no-wear’ models
can achieve acceptable life prediction accuracy for partial slip
cases.

5.5. Effect of wear coefficient on fretting fatigue behaviour

As mentioned above, it is of interest to study the predicted


interaction between wear and fatigue life and this can be eas-
Fig. 15. Effect of wear coefficient on predicted fretting fatigue life (for the
ily achieved using the present methodology. For example, if
δapp = 40 ␮m partial slip case).
we consider the evolution of contact pressure and SWT dis-
tributions of Figs. 11 and 13, it can be seen that the time (N* )
at which the pressure peak and the associated SWT damage give stick–slip interface cracking. A specific estimation of the
peak at the stick–slip interface starts to develop is dependent wear coefficient of k ≈8 × 10−7 MPa−1 has been obtained for
on the wear coefficient. If wear is retarded (i.e. by use of a this case, following the method outlined by McColl et al. [5].
low wear coefficient), the initial geometry and its associated The present analyses predict failure to occur at the stick–slip
pressure distribution (pressure and SWT peaks at either edge interface under partial slip conditions for this k value. It is
of contact) will dominate fatigue damage accumulation so that worth pointing out that the value of kcrit will change for dif-
failure is predicted to occur at the contact edges. However, if ferent fretting conditions, e.g. applied displacement, normal
wear evolves more rapidly (i.e. high wear coefficient), the pres- load.
sure and SWT peaks at the edge of contact are reduced, but the These results show that the interaction between fatigue and
stick–slip interface SWT peaks then dominate damage accumu- fretting wear is critical to the prediction of failure cause, location
lation and hence failure is predicted at the stick–slip interface. and number of cycles. Essentially, if the component can sur-
The effect on predicted fatigue life of varying the wear coeffi- vive the initially severe loading conditions, viz. edge-of-contact
cient is presented here for the partial slip case of δapp = 40 ␮m. pressure peaks and associated contact stress concentrations, so
Figs. 14 and 15 show the effect of k on predicted failure posi- that wear-induced material removal can alter the pressure dis-
tion and life, respectively. The failure position is found to switch tribution, failure will occur at the stick–slip boundary, due to a
from the edge of contact for k < kcrit ≈ 5.0 × 10−8 MPa−1 to the completely different stress distribution to that given by the initial
stick–slip interface for k ≥ kcrit ≈ 5.0 × 10−8 MPa−1 . Referring geometry.
to Fig. 15, k is predicted to have relatively little effect on life The reason for this peak in life at kcrit (Fig. 15) lies in
for k < 0.5 × kcrit ; the life is predicted to suddenly increase by a the material utilisation. Fig. 16 shows the δapp = 40 ␮m par-
factor of about 5 as k approaches kcrit . Finally, as k is increased tial slip damage contour plots corresponding to failure (W = 1)
further, for k > kcrit , the life is found to gradually decay back for three values of wear coefficient, namely k = 0.55 × kcrit
to approximately the same values as for k < kcrit . Jin and Mall (k = 2.75 × 10−8 MPa−1 ), k ≈ kcrit (k = 5.5 × 10−8 MPa−1 ) and
[7] have presented a wear scar for the partial slip case of the k = 1.6 × kcrit (k = 8 × 10−8 MPa−1 ). In the k = 0.55 × kcrit case
punch on flat geometry analysed here, which was observed to (Fig. 16a) damage is highly localised at the trailing edge of
contact. Conversely, in the k = 1.6 × kcrit case (Fig. 16c), dam-
age is highly localised to the stick/slip boundary. However,
in the k ≈ kcrit case (Fig. 16b), two regions of high damage
are found, one near the edge of contact, and a second at the
stick/slip interface. The material utilisation is high in this case
because a greater amount of fatigue damage has been sustained
by spreading it over a larger area. Fig. 15 indicates that there is an
optimum wear coefficient (kcrit ) for maximum life under given
fretting conditions. These results furthermore indicate that the
present wear-fatigue methodology has the capability to model,
and therefore ultimately design for, commonly-observed phe-
nomena from industrial practice such as ‘bedding-in’, whereby
beneficial effects can be derived from using a mixed material
combination with one wear-‘soft’ material and one wear-‘hard’
Fig. 14. Effect of wear coefficient on predicted failure position (δapp = 40 ␮m). material.
550 J.J. Madge et al. / Wear 263 (2007) 542–551

interface, with the sudden shift in position predicted at a single


(critical) value of wear coefficient.
• The same critical value of wear coefficient was shown to
give an optimum (maximum) fatigue life, for a given set of
fatigue constants and fretting conditions. This method pro-
vides a framework for designing for optimum fretting fatigue
life.
• The work has demonstrated that prediction of fretting wear
is a critical aspect for prediction of fretting fatigue and can
have a major effect on the predicted evolution of stresses
and damage parameters (e.g. SWT), and hence on predicted
fatigue lives and failure position. This fundamental and impor-
tant effect for fretting fatigue cannot be predicted by models
which do not include the effects of material removal due to
wear.

Acknowledgements

The authors wish to thank Rolls-Royce plc, Aerospace Group,


for their financial support of the research, which was carried
out at the University Technology Centre in Gas Turbine Trans-
mission Systems at the University of Nottingham. The views
expressed in this paper are those of the authors and not neces-
sarily those of Rolls-Royce plc, Aerospace Group.

Fig. 16. Showing the damage contours at failure (W = 1) for different References
values of wear coefficient: (a) k = 2.75 × 10−8 MPa−1 < kcrit ; (b) k = 5.5 ×
10−8 MPa−1 ≈ kcrit ; (c) k = 8 × 10−8 MPa−1 > kcrit .
[1] J.A. Collins, S.M. Macro, The effect of stress direction during fretting
on subsequent fatigue life, Proc. Am. Soc. Test. Mater. 64 (1964) 547–
6. Conclusions 560.
[2] J.M. Dobromirski, in: R.B. Waterhouse, M.H. Attia (Eds.), Variables in
the Fretting Process: are There 50 of Them? Standardisation of Fretting
In this paper, a finite element based method has been applied Fatigue Test Methods and Equipment, ASTM, 1992, pp. 60–66.
to predict and characterise the role of fretting wear on fretting [3] D.A. Hills, D. Nowell, The Mechanics of Fretting Fatigue, Dordrecht,
fatigue for a rounded punch-on-flat Ti-6Al-4V fretting test con- Kluwer, 1994.
figuration, with comparisons against published test data from [4] O. Jin, S. Mall, Effects of slip on fretting behaviour: experiments and
analyses, Wear 256 (2004) 671–684.
Mall and co-workers. The key findings are as follows:
[5] I.R. McColl, K. Ding, S.B. Leen, Finite element simulation and exper-
imental validation of fretting wear, Wear 256 (11–12) (2003) 1114–
• A significant effect of slip amplitude on fretting fatigue life 1127.
was predicted, viz. a critical range of slip amplitude for mini- [6] S.B. Leen, C.H.H. Ratsimba, I.R. McColl, E.J. Williams, T.R. Hyde, An
investigation of the fatigue and fretting performance of a representative
mum life corresponding broadly to the transition from partial aeroengine splined coupling, J. Strain Anal. 37 (6) (2002) 565–585 (special
slip to gross sliding, which is consistent with (i) a generally issue on fretting fatigue).
accepted phenomenon of fretting fatigue and (ii) correspond- [7] O. Jin, S. Mall, Influence of contact configuration on fretting fatigue behav-
ing FE-based life predictions (with wear) [10] and test data ior of Ti-6Al-4V under independent pad displacment condition, Int. J.
for a Hertzian fretting test configuration. Fatigue 24 (2002) 1243–1253.
[8] K. Nakazawa, N. Maruyama, T. Hanawa, Effect of contact pressure on
• The life predictions were shown to exhibit good quantitative fretting fatigue of austenitic stainless steel, Tribol. Int. 36 (2003) 79–
agreement with available test data in the region of minimum 85.
fatigue life. [9] O. Vingsbo, S. Soderberg, On fretting wears, Wear 126 (1988) 131–147.
• Prediction of failure position: It was shown that for gross [10] J.M. Madge, S.B. Leen, I.R. McColl, P.H. Shipway, Contact evolution based
sliding cases, failure is invariably predicted at the edge of prediction of fretting fatigue life: effect of slip amplitude, Wear 262 (2007)
1159–1170.
contact, consistent with experimental observations. However, [11] T. Hattori, T. Watanabe, Fretting fatigue strength estimation considering
for partial slip cases, it was shown that due to the competi- the fretting wear process, Tribol. Int. 39 (2006) 1100–1105.
tion between wear rate and fatigue damage accumulation and [12] K.N. Smith, P. Watson, T.H. Topper, A stress-strain function for the fatigue
the effects of edge-of-contact versus stick–slip interface SWT of metals, J. Mater. 15 (1970) 767–778.
concentrations, failure location depended on the wear coef- [13] V. Sabelkin, S. Mall, Investigation into relative slip during fretting fatigue
under partial slip contact condition, Fatigue Fracture Eng. Mater. Struct.
ficient (for assumed constant fatigue properties). Thus, for 28 (2005) 809–824.
low wear coefficients, partial slip failure is predicted at the [14] ABAQUS User’s and Theory Manuals, Version 6.5, ABAQUS Inc., Paw-
edge-of-contact and for high wear coefficients at the stick–slip tucket, Rhode Island, 2005.
J.J. Madge et al. / Wear 263 (2007) 542–551 551

[15] R. Magaziner, O. Jin, S. Mall, Slip regime explanation of observed size [21] V. Fridrici, S. Fouvry, P. Kapsa, P. Peruchaut, Prediction of cracking in
effects in fretting, Wear 257 (2004) 190–197. Ti-6Al-4V alloy under fretting-wear: use of the SWT criterion, Wear 259
[16] A. Fatemi, D.F. Socie, A critical plane approach to multiaxial fatigue dam- (1–6) (2005) 300–308.
age including out-of-phase loading, Fatigue Fracture Eng. Mater. Struct. [22] C.D. Lykins, S. Mall, V.K. Jain, Combined experimental-numerical investi-
11 (3) (1988) 149–165. gation of fretting fatigue crack initiation, Int. J. Fatigue 23 (2001) 703–711.
[17] D. Socie, Multiaxial fatigue damage models, J. Eng. Mater. Technol. Trans. [23] W.S. Sum, E.J. Williams, S.B. Leen, Finite element, critical plane, fatigue
ASME 109 (1987) 293–298. life prediction of simple and complex contact configurations, Int. J. Fatigue
[18] J.A. Araujo, D. Nowell, The effect of rapidly varying contact stress fields 27 (2005) 403–416.
on fretting fatigue, Int. J. Fatigue 24 (2002) 763–775. [24] S.M. Marko, W.L. Starkey, A concept of fatigue damage, Trans. ASME 76
[19] M.P. Szolwinski, T.N. Farris, Mechanics of fretting fatigue crack formation, (4) (1954) 627–632.
Wear 198 (1996) 93–107. [25] N.E. Dowling, Mechanical Behaviour of Materials: Engineering Methods
[20] M.W. Brown, K.J. Miller, A theory for fatigue failure under multiaxial for Deformation, in: Fracture and Fatigue, second ed., Prentice Hall, NJ,
stress-strain conditions, Proc. Inst. Mech. Eng. 187 (1973) 745–755. 1998.

You might also like