You are on page 1of 10

International Journal of Machine Tools & Manufacture 44 (2004) 125134

www.elsevier.com/locate/ijmatool
The eect of machined topography and integrity on fatigue life
D. Novovic
a,
, R.C. Dewes
a
, D.K. Aspinwall
a
, W. Voice
b
, P. Bowen
a
a
School of Engineering, University of Birmingham, Edgbaston, Birmingham, B15 2TT, UK
b
Rolls-Royce Plc, P.O. Box 31, Derby, DE24 8BJ, UK
Received 6 August 2003; accepted 17 October 2003
Abstract
The paper reviews published data which address the eect of machining (conventional and non-conventional processes) and the
resulting workpiece surface topography/integrity on fatigue performance, for a variety of workpiece materials. The eect of post-
machining surface treatments, such as shot peening, are also detailed. The inuence of amplitude height parameters (Ra, Rt),
amplitude distribution (Rsk) and shape (Rku) parameters, as well as spatial (Std, Sal) and hybrid (Ssc) measures, are considered.
There is some disagreement in the literature about the correlation between workpiece surface roughness and fatigue life. In
most cases, it has been reported that lower roughness results in longer fatigue life, but that for roughness values in the range 2.5
5 lm Ra it is primarily dependent on workpiece residual stress and surface microstructure, rather than roughness. In the absence
of residual stress, machined surface roughness in excess of 0.1 lm Ra has a strong inuence on fatigue life. Temperatures above
400
v
C reduce the eects of both residual stress and surface roughness on fatigue, due to stress relieving and the change in crack
initiation from the surfaces to internal sites. The presence of inclusions an order of magnitude larger than the machined surface
roughness generally overrides the eect of surface topography.
# 2003 Elsevier Ltd. All rights reserved.
Keywords: Surface topography; Surface integrity; Fatigue life
1. Introduction
Fatigue is of great concern for components subject to
cyclical stresses, particularly where safety is paramount,
for example aeroengine parts. It can also contribute to
the failure of components such as moulds/dies, gears,
bearings and shafts, and therefore have a detrimental
eect on life cycle/operating costs. It has long been
recognised that fatigue cracks generally initiate from
free surfaces and that performance is therefore reliant
on the surface topography/integrity produced by
machining. Koster and Field [1] suggested that the
main mechanical property aected by machining is high
cycle fatigue (HCF) strength, the actual endurance limit
being dependent on the particular process used and the
severity of operation. Whilst it is known that fatigue
life is heavily inuenced by residual stresses, the metal-
lurgical condition of the material (microstructure and
microhardness) and the presence of notch-like surface
irregularities induced by machining play a key role.
Many early fatigue models had limited predictive abil-
ity, and fatigue performance was often attributed
entirely to amplitude workpiece surface roughness
parameters, such as Ra. Recent research to establish
functionality with respect to surface measurement
descriptors, has highlighted the inadequacy of solely
using amplitude measures for relating fatigue behav-
iour. The inuence of post surface treatments, such as
shot peening and roller burnishing is also critical. Other
variables which aect component fatigue life, but are
outside the scope of this paper, are test conditions,
stress level, mode of stressing, environment, specimen
size, the condition of the material, etc. [24].
2. Eect of surface topography on fatigue life
2.1. Conventional amplitude parameters
Some of the earliest published work on the eect of
workpiece surface topography on fatigue life was

Corresponding author. Tel./fax: +44-121-414-3541.


E-mail address: dxn198@bham.ac.uk (D. Novovic).
0890-6955/$ - see front matter #2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijmachtools.2003.10.018
undertaken by Thomas [5] and coincided with develop-
ment/use of the rst metrology instruments in the early
1930s. In many early fatigue models [57], performance
was often entirely attributed to amplitude surface
roughness parameters, in particular the arithmetic aver-
age, Ra. The ISO standard [8] provide denitions of
this and other 2D surface topography parameters. Noll
and Erickson [6] observed that the fatigue strength of
ground components made from carbon and alloy
steels, subject to low applied stresses, was 20 times
greater than that of as-forged material. This was attrib-
uted solely to dierences in surface roughness values.
Similarly, Fluck [7] noted high scatter in fatigue test
results at low applied stresses with carbon (AISI 1035)
and Ni-Cr (AISI 3130) steels, when investigating the
inuence of workpiece surface roughness and its direc-
tionality. Polishing of specimens with circumferential
grooves produced by a turning operation, was found to
extend life, however, regardless of size, longitudinal
scratches showed no eect on fatigue life (rotating bend
mode of testing). Somewhat surprisingly, the high fati-
gue life values exhibited by ground specimens were
attributed to inaccuracies in the surface measurement
instrument, rather than the inuence of surface integ-
rity.
There is a suggestion that microcracks in a specimen
with low surface roughness (such as a polished sur-
face), initiate from persistent slip bands, as explained
by Woods concept in Fig. 1, or at grain boundaries
[24]. However, where micro-notches induced by a
machining process (e.g. grinding, turning and milling)
already exist, the associated stress concentration will
generate a localised plastic strain eld when a stress is
applied. This zone of plasticity will occur in an individ-
ual grain in the form of a slip band, which will dene
the path of the shear crack [2].
Siebel and Gaier [9] found that the maximum depth
of surface irregularities following machining was the
most signicant parameter from a fatigue perspective.
They compared fatigue strength with the maximum
depth of a surface groove (Rt) in steels (tempered
spring, Cr-Mo, annealed medium carbon) and a variety
of non-ferrous alloys. They concluded that above a cer-
tain critical groove depth (Ro), the reduction in fatigue
endurance limit was proportional to log Rt. The Ro
value was found to be 12 lm for tempered steels and
brass, and 46 lm for annealed steel, Al-Cu and
Al-Mg alloys, as shown in Fig. 2. Stress-relieving of
steel specimens (500
v
C in a vacuum for an unspecied
time) resulted in a fatigue life independent of machined
surface residual stress, except with spring steel, where
tempering occurred. The authors also stated that the
inuence of surface roughness on fatigue life, could be
eliminated by employing a post-machining mechanical
or thermal surface treatment, in order to produce a
compressive or neutral residual stress at/near the sur-
face and suppress fatigue crack initiation.
The eect of surface roughness on fatigue strength
varies from one material to another. Fig. 3 presents a
summary of experimental work for several workpiece
materials (iron, nickel and titanium based alloys), sub-
ject to a range of machining processes (nish grinding,
milling and turning). The resulting surface roughness
varied signicantly within each group of specimens (for
example, Ra 827 linch (0.20.7 lm) for longitudinal
surface grinding of AISI 4340 steel), however, residual
stresses were almost identical for each group (for
example a value of ~138 to 207 MPa in com-
pression for gentle grinding of AISI 4340 steel) [10]. It
has been found that under typical machining con-
ditions using sharp tools, the endurance limit of
Ti 6-6-2 and Inconel 718 was not dependent on surface
roughness, however, this was the case with steels.
Many machined components subject to fatigue are
exposed to high in-service temperatures, which causes
further complications during analysis, because the
eect of irregularities on the workpiece surface alters
with temperature. Maiya and Busch [11] investigated
the eect of workpiece surface roughness on the fatigue
Fig. 1. Woods concept of micro-deformation leading to formation
of fatigue crack: (a) static; (b) fatigue deformation leading to surface
notch (intrusion); (c) fatigue deformation leading to slip-band
extrusion [3].
Fig. 2. Fatigue strength in repeated tension of various steels as a
function of surface roughness [9].
126 D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134
life of AISI 304 stainless steel at elevated temperature
(593
v
C). Their results suggested that fatigue life
decreased with an increase in surface roughness, and
that roughness signicantly aected the fatigue
initiation component, No(Rq), which was described by:
NoRq 1012Rq
0:21
1
where No is the initiation component of total fatigue
life and Rq is the root-mean-square roughness (lm).
Wareing and Vaughan [12] studied fatigue resistance
at elevated temperatures and found that the fatigue life
of ne turned, ne ground and electropolished AISI
316 stainless steel at 400
v
C was dominated by crack
propagation/growth rather than initiation, due to the
dierences in fatigue crack initiation mechanisms for
smooth and machined surfaces. A four-fold improve-
ment in fatigue life for electropolished specimens (Ra
<1 lm) was explained by the fact that the growth rate
of cracks with semicircular fronts was only 25% that of
sharp cracks found in turned (Ra 1.53 lm) and
ground (Ra 24 lm) specimens. However, at tempera-
tures in excess of 600
v
C, fatigue crack initiation
moved from a transgranular to predominantly inter-
granular mode, hence fatigue life was found to be inde-
pendent of surface roughness.
It has been suggested that a full understanding of the
eects of surface topography on fatigue performance
can only be achieved by employing stress-free speci-
mens. Unfortunately, stress relieving of components
after machining can cause their metallurgical condition
to be altered. Taylor and Clancy [13] examined the fati-
gue performance of stress relieved En19 steel (equiva-
lent to AISI 4140) and found that even the relatively
low workpiece surface roughness (Ra 0.51.4 lm, Rt 7
14 lm) produced by moderate grinding reduced fatigue
life, compared to polished samples (Ra 0.10.3 lm, Rt
35 lm), as shown in Table 1. The ground surfaces,
however, exhibited lower fatigue resistance than milled
surfaces with a similar roughness level (Ra, Rt). This
was explained by the curved feed marks caused by mill-
ing, which led to the early initiation of cracks in orien-
tations perpendicular to the specimen axis. It was
proposed that surface parameters, such as Rt and Rz,
are more appropriate indicators of fatigue performance
than Ra, as they reect the worst defects present in
the workpiece surface.
Suhr [2,14] suggested that fatigue cracks in stress-free
specimens could grow from short cracks (<50 lm in
depth), as dened by Newman et al. [15]. This is below
the threshold found in standard fracture mechanics
equations. Suhr argued that such small cracks were not
detected by standard methods of surface roughness
measurement or random cross-sectional examination of
surfaces, but only by detailed analysis of fracture sur-
faces. Suhrs work demonstrated that surface prep-
aration had a major eect on fatigue performance in
the absence of residual stress.
2.2. Functional parameters
The relatively large scatter in fatigue results reported
in many studies (typically 20%) for the same Ra value,
questions the appropriateness of relying solely on this
parameter. More recent models have recognised that
measures relating to maximum surface irregularities are
Fig. 3. Eect of surface nish on the fatigue strength of various materials [10].
Table 1
Eect of processing and roughness data on fatigue endurance limit
for stress relieved samples [13]
Surface Ra [lm] Rt [lm] Fatigue limit
[MPa]
Polished 0.10.3 35 775
Ground 0.51.4 714 690
Milled (ne) 1.02.2 1115 775
Milled (coarse) 1.11.8 2634 733
Shaped (ne) 2733 210280 620
Shaped (coarse) 3544 360390 520
D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134 127
a better indicator of fatigue performance. There have
also been limited attempts to determine other descrip-
tors for surfaces subject to fatigue, however, the most
widely used of the 57 2D parameters in current ISO
and British standards [8,16,17] and the 17 or so 3D
parameters [1723] continues to be Ra [23]. In an
attempt by Bayomi and Abd El-Latif [24] to correlate
the various surface roughness parameters with the fati-
gue endurance limit of an aluminium alloy, Ra and Rq
rather than spatial parameters, were found to have a
controlling inuence on endurance limit. The eect of
the hybrid parameters Dq (root mean square slope)
and kq (spacing between local peaks and valleys) on
fatigue endurance limit was found to be less than that
of the amplitude parameters.
It has been suggested that the shape of grooves, not
just the maximum height of the prole Rt, are critical,
as both factors can lead to stress concentrations [25],
which, under fatigue conditions, can initiate cracking.
Statistical parameters, particularly kurtosis, Rku, which
is derived from the fourth moment of the amplitude
distribution function [8,21,26], have consequently been
identied as likely indicators of fatigue strength. In an
attempt to relate appropriate statistical parameters of
machined surfaces to applications, Stout [27] suggested
that machining processes lead to the production of
non-Gaussian surfaces (skewness, Rsk 6 0 and kur-
tosis, Rku 6 3), see Table 2 and Fig. 4, and that Ra
and kurtosis play a major role in controlling fatigue
performance.
The functionality of machined surfaces is often based
on the experience and knowledge of an individual com-
pany. An example is the use of roughness parameters
based on the motif method, employed by companies,
such as the automotive manufacturer Renault. Here
the maximum depth of a prole irregularity (Rx) is
used as the primary measure of fatigue performance,
with mean depth (R) and mean spacing (AR) of rough-
ness motifs as secondary parameters, as dened in ISO
12085:1998 [28].
Griths [18] conducted a comprehensive survey over
many years concerning the relationship between surface
integrity and functionality. In addition to amplitude
height and distribution parameters, the 3D surface spa-
tial parameters, texture direction (Std) and the fastest
decay autocorrelation length (Sal), were recognised as
relevant to fatigue strength, see Table 3. Stout et al.
[20], suggested that surface volume parameters (Sci-
core uid and Svi-valley uid retention index), together
with a hybrid parameter Ssc (arithmetic mean summit
curvature of the surface) should also be included.
In recent years, there has been increasing research
activity on providing correlation between surface char-
acteristics and functionality. This has mainly been sti-
mulated by the development of 3D surface
characterisation equipment and new techniques for sur-
face description [23,29,30]. It is anticipated that a new
ISO standard on 3D surface texture parameters, which
is in the nal stage of review, will include not just sur-
face characterisation, but also functional aspects.
3. Eect of workpiece surface integrity on fatigue
life
It has been suggested that when surface roughness is
2.55 lm Ra, surface residual stress is often a better
indicator of fatigue performance than surface topogra-
phy [10].
This eect diminishes with an increase in tempera-
ture due to the relaxation of residual stress with ther-
mal exposure [10]. Koster and Field [1] found that a
wide range of fatigue strengths could be achieved in
conventional alpha/beta titanium alloy Ti-6Al-4V, as a
result of dierent machining processes, see Fig. 5. They
insisted that these variations occurred not as a conse-
quence of changes in surface roughness, but because of
alterations in integrity induced in the surface of the
material, as a result of the machining operation.
El-Helieby et al. [31] studied the fatigue resistance
of ground En31 steel (equivalent to AISI E52100)
and concluded that residual stress was much more
Table 2
Typical skewness and kurtosis values for various machined surfaces
[27]
Process Typical skewness Typical kurtosis
Turning +0.2 to +0.1 2 to 4
Milling +0.2 to 0.6 2 to 10
Grinding +0.0 to 0.8 2 to 6
Reaming 0.5 to 1.0 3 to 8
Honing 0.5 to 1.0 3 to 10
EDM 0.0 to +1.2 2.5 to 4
Sand blasting 0.0 to +1.4 2.5 to 3
Normally distributed surface 0 3
Fig. 4. Illustration of skewness and kurtosis [27].
128 D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134
important than surface roughness. High temperatures
generated during abusive and conventional grinding
resulted in tensile residual stress, which had a detri-
mental eect on fatigue life, despite lower workpiece
surface roughness over that generally achieved with
other machining processes. An approximately linear
relationship between surface residual stress and fatigue
strength was found. It has also been suggested by
Leverant et al. [25] that the relatively sharp surface
notches characteristic of grinding, are responsible for
the inferior fatigue performance of ne ground samples
(0.6 lm Ra), when compared with samples produced
by other processes having signicantly higher surface
roughness. Results of this study, conducted on
Ti-6Al-4V, also highlighted the importance of surface
residual stress in controlling the development of micro-
cracks and therefore overall fatigue life at both room
and elevated temperatures. Similar ndings have been
reported by other authors [32,33].
Many publications [18,25,34,35] emphasise that tem-
peratures during grinding must be carefully controlled,
in order to avoid impaired surface integrity (untem-
pered or overtempered martensite, microcracking,
overaging, microhardness changes, plastically deformed
debris, side ow and tensile residual stress), as shown
in Fig. 6 [1]. Koster [10] stated that the residual stress
which controls fatigue behaviour is frequently sub-
surface, not outer surface stress, as often suggested.
Non-conventional machining processes have a varied
eect on workpiece surface integrity and therefore fati-
gue performance, depending on the material removal
mechanism employed. Electrical discharge machining
(EDM) is a thermal process and as a result, the work-
piece surface is often characterised by a heat aected
zone with an upper hard, brittle, recast layer, contain-
ing cracks and microcracks (typically running normal
to the machined surface). This region is also subject to
a tensile residual stress regime resulting from thermal
contraction [18,3638]. The combination of these fea-
tures generally leads to poor fatigue performance
[1,37,38]. A similar situation is often observed follow-
ing laser-beam machining (LBM), where thermal con-
traction of the workpiece material due to rapid
solidication also induces unfavourable residual stres-
ses [36]. In order to compensate for the detrimental
eects of such processes on surface integrity, post-
machining heat or mechanical treatments are often
performed and the eect of these is discussed later in
the paper.
Table 3
Functional performance and roughness parameters [18]
Typical parameters Heights Distribution and shape Slopes and curvature Lengths and peak space Lay and lead
Ra, Rq, Rt, Sa, Sq Rsk, Rku, Ssk, Sku RDq, SDa Rsm, HSC, Sal Std, Sal
Fatigue u . v v u
u, Much evidence; ., some evidence; v, little or circumstantial evidence.
Fig. 5. High cycle fatigue behaviour of Ti-6Al-4V (32 HRC) [1].
Fig. 6. Residual stress in AISI 4340 steel after surface grinding [1].
D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134 129
Abrasive water-jet machining (AWJM) relies princi-
pally on a mechanical mechanism of material removal,
as a result of impacts from individual grits suspended
within the uid. Understandably therefore, the process
generally induces compressive residual stresses, which
would be expected to contribute to good fatigue per-
formance. Unfortunately, the machined surface often
contains embedded grits [3840], typically 1050 lm in
diameter, which, if combined with a high workpiece
surface roughness, act as stress concentrations and sig-
nicantly degrade fatigue performance [38].
Extensive data on machinability/surface integrity/
fatigue performance have been published by research-
ers at The University of Birmingham. This has
involved workpiece materials, such as hardened steels
and aerospace alloys, machined using a variety of pro-
cesses [37,4147]. When assessing the fatigue strength
of a gamma titanium aluminide (c-TiAl) following
EDM and polishing, Trail and Bowen [37] found little
dierence in performance, although the EDMed sur-
faces were generally characterised by tensile residual
stresses and severe cracking. Similarly, research by Sal-
mani [46] indicated that for the same workpiece
material, the eect of workpiece surface integrity on
fatigue life depended principally on microstructure.
Mantle and Aspinwall [43] found little dierence in the
fatigue life of turned c-TiAl samples produced under
various operating conditions (range of workpiece sur-
face roughness 0.461.04 lm Ra) and polished samples
(0.12 lm Ra), even when machining with a worn tool
which caused signicant levels of cracking and subsur-
face damage, see Fig. 7. Sharman et al. [45] reported a
signicantly higher fatigue run-out strength of c-TiAl
following turning than after electrochemical machining
(ECM) or electrodischarge texturing (EDT). The pres-
ence of cracks up to 5 lm deep in the turned surfaces
was compensated by highly compressive residual stres-
ses, which reduced the applied stress. Similarly, turning
was found to be more favourable than grinding with
respect to the fatigue performance of a hardened bear-
ing steel [41]. High speed milling with the correct oper-
ating parameters, can also enhance fatigue performance
[42,44,47,48]. The eect is associated with plastic defor-
mation/strain hardening and relatively low tempera-
tures, which result in compressive residual stresses at/
near the machined surface.
Gu ngo r and Edwards [49] studied the inuence of
surface texture on fatigue life for a forged Al-Mg-Si
alloy. Specimens with three dierent Ra roughness
values (0.05 lm for polished surfaces, 0.54 lm for as-
forged surfaces solution treated in molten salt, and
0.82 lm for as-forged surfaces solution treated in air),
were tested. It was concluded that contaminant parti-
cles 2080 lm in diameter (present exclusively at the
surface and containing a variety of non-metallic mate-
rials including silica, titanium and graphite) on the as-
forged specimens, acted as nucleation sites from which
cracks propagated, as shown in Fig. 8. This resulted in
a 25% lower fatigue performance for the forged com-
ponents compared to polished surfaces. Other studies
also suggest that inclusions act as stress concentration
sites, and if their size is an order of magnitude higher
than the Ra value, they outweigh any eect due to sur-
face roughness. A reduction in fatigue limit of 50% due
to the presence of inclusions has been reported [14].
4. Eect of post-machining surface treatments
Mechanical surface treatments such as shot or glass
bead peening and roller burnishing are often utilised
after machining in order to enhance fatigue perform-
ance [5063]. These lead to the development of com-
pressive residual stresses, work hardening (high
dislocation densities near the workpiece surface) and
alteration of workpiece surface roughness/surface top-
ography. The individual contributions of these para-
meters to total fatigue life is detailed in [51] and is
summarised in Table 4. Additional eects that can
inuence the fatigue performance of mechanically
treated surfaces, result from stress-induced phase trans-
Fig. 7. Fatigue life of turned gamma titanium aluminide (45-2-2+0.8) [43].
130 D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134
formations and deformation-induced changes in the
crystallographic texture in near-surface regions [10,52].
The inuence of post-machining treatments on fati-
gue life depends primarily on the value and depth of
the induced residual stress and the alteration in surface
roughness. Wagner [53] suggested that high-cycle fati-
gue cracks in mechanically surface treated specimens,
often nucleate below the surface at the position of
maximum tensile residual stress, which balances the
benecial compressive stress eld near the surface, see
Fig. 9.
In shot peening, the depth of the compressive layer
is typically in the range 100300 lm, while glass
bead peening generally produces compressive layers
2575 lm deep, depending on peening parameters and
material condition (hardness, strength, microstructure,
etc.) [53,62]. It has been found that an increase in
Almen intensity (a standard measure of the extent of
inuence from a shot peening operation), leads to
degradation of shot peened components, which can
cancel out the benets and result in a decrease in fati-
gue life [10,53,55]. The eect of Almen intensity on
fatigue life for various materials can be seen in Fig. 10.
Its unfavourable eect can in some cases be reduced
through post-shot peening mechanical polishing/elec-
tropolishing. This changes the controlling factor on
fatigue strength from microcrack propagation (rough
surfaces) to crack initiation (smooth surfaces).
It is widely recognised that the benets of mechani-
cally induced compressive residual stresses are reduced
Fig. 8. Fatigue crack nucleation from an inclusion [49].
Table 4
The eects of surface layer properties on fatigue life [51]
Crack nucleation Crack propagation
Surface roughness Accelerates No eect
Cold work Retards Accelerates
Residual compressive stress Minor or no eect Retards
Fig. 9. Crack initiation at the position of maximum tensile residual
stress in a shot peened specimen: (a) fractured surface; and (b)
residual stress depth prole [51].
Fig. 10. Fatigue life as a function of Almen intensity for various
materials [10].
D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134 131
if peened components are exposed to temperatures in
excess of 300
v
C [52,54,56,60,61]. In Ti-6Al-4V alloy,
compressive stresses were reported to be completely
relieved after annealing for 1 h at 600
v
C [52]. How-
ever, the presence of the work hardened layer may still
benet fatigue life through the retardation of crack
initiation due to the higher strength. A higher tempera-
ture is likely to counteract this due to recrystallisation
[51]. Mechanical polishing after peening has been
found to compensate for this eect [52]. The benets of
compressive residual stresses can also be reduced by
cyclic plastic deformation during either fatigue testing
or fatigue loading in service [5659]. It has been found
that the rate of residual stress relaxation can be rapid
in the early stages of fatigue cycling, with a reduction
of more than 50% during the rst few cycles [59].
Processes such roller burnishing and laser shock
peening may further enhance fatigue life due to the
ability to generate signicantly reduced surface rough-
ness and greater depth and magnitude of near-surface
compressive stresses, than can be achieved with shot/
glass bead peening [53,57,62,63]. Moreover, surfaces
treated with these processes have greater thermal
residual stress relaxation resistance, due to lower cold
working and lower propensity for annealing or recrys-
tallisation [59]. An example of the rise in fatigue
strength for roller burnished Beta C titanium alloy
relative to the shot peened condition is shown in
Fig. 11. It has been suggested that mechanical surface
treatments, in combination with appropriate thermal
treatments, could lead to an additional rise in fatigue
resistance, as observed in titanium alloys [51].
5. Conclusions
. The importance of both surface topography and
integrity parameters is well recognised, with many
experimental and analytical models relating to fati-
gue. Unfortunately, no reliable model which can
predict fatigue strength with respect to all factors
appears to exist at present.
. Up to ~90% of high cycle fatigue life involves crack
initiation. In the absence of residual stress,
machined surface roughness values in excess of 0.1
lm Ra have a strong inuence on fatigue life. With
surfaces of <0.1 lm Ra, this eect diminishes as
cracks initiate due to persistent slip bands or at
grain boundaries.
. The presence of inclusions an order of magnitude
larger than the machined surface roughness, gener-
ally overrides the eect of surface topography.
. Amplitude function distribution and shape para-
meters (e.g. skewness and kurtosis), together with
surface texture parameters (e.g. lay and lead) are
gaining increased recognition as descriptors indica-
tive of fatigue strength.
. When workpiece surface roughness values are
between 2.55 lm Ra, residual stress is a signicant
factor relative to fatigue life. This eect is largely
dependent on material properties and strain harden-
ing.
. Temperatures in excess of 400
v
C reduce the eects
of both residual stress and surface roughness on fati-
gue, due to stress relieving and the change in crack
initiation from the surface to internal sites.
. If the correct operating parameters are employed,
machining processes such as high speed milling,
turning and grinding (reciprocating and creep feed)
can provide enhanced fatigue performance over
polished surfaces. The eect of higher surface rough-
ness is diminished by the inducement of compressive
residual stresses at and near the machined surface
through strain hardening and low temperatures.
. Mechanical and thermal surface treatments can be
benecial in improving workpiece fatigue perform-
ance, although the eect largely depends on work-
piece material properties.
Acknowledgements
The authors would like to thank the School of
Engineering, Universities UK and Rolls-Royce Plc. for
the provision of funding. We would also like to
acknowledge the invaluable advice given by Dr Brian
Griths, Brunel University, Prof. Liam Blunt, Univer-
sity of Hudderseld and Profs. John Knott and Mike
Loretto, University of Birmingham.
Fig. 11. Eect of mechanical surface treatment on SN curves
(R 1) for Beta C titanium alloy [53].
132 D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134
References
[1] W.P. Koster, M. Field, Eects of machining variables on the
surface and structural integrity of titanium, Proceedings of the
North American Manufacturing Research Conference, SME,
vol. 2, 1973, pp. 6787.
[2] R.W. Suhr, High cycle fatigue, The Institute of Metals, London,
1988, pp. 226287, ISBN 0-901-46251-9.
[3] G.E. Dieter, Mechanical Metallurgy, SI Metric Edition, McGraw-
Hill Book Company, London, 1988, ISBN 0-07-100406-8.
[4] S. Suresh, Fatigue of Materials, Cambridge University Press,
Cambridge, 2001, ISBN 0521570468.
[5] W.N. Thomas, Eect of scratches and of various workshop n-
ishes upon the fatigue strength of steel, Engineering 116 (1923)
449485.
[6] G.C. Noll, G.C. Erickson, Allowable stresses for steel members
of nite life, Proceedings of the Society of Experts for Stress
Analysis, 5(2), 1948, pp. 132143.
[7] P.G. Fluck, The inuence of surface roughness on the fatigue
life and scatter of test results of two steels, Proceedings of Amer-
ican Society for Testing Materials, vol. 51, 1951, pp. 584592.
[8] ISO 4287:2000 Geometrical Product Specication (GPS), Surface
texture: Prole method, Terms, denitions and surface texture
parameters.
[9] E. Siebel, M. Gaier, Inuence of surface roughness on the fati-
gue strength of steels and non-ferrous alloys, Engineers Digest
18 (1957) 109112, Translation from VDI Zeitschrift 98 (30)
(1956) 17151723.
[10] W. Koster, Eect of residual stress on fatigue of structural
alloys, Proceedings of the Third International Conference, ASM
International, Indianapolis, Indiana, USA, 1991, pp. 19.
[11] P.S. Maiya, D.E. Busch, Eect of surface roughness on low-cycle
fatigue behaviour of type 304 stainless steel, Metallurgical Trans-
actions 6A (1975) 17611774.
[12] J. Wareing, H.G. Vaughan, Inuence of surface nish on low-
cycle fatigue characteristics of type 316 stainless steel at 400
v
C,
Metal Science 13 (1979) 18.
[13] D. Taylor, O.M. Clancy, Fatigue performance of machined sur-
faces, Fatigue and Fracture of Engineering Materials and Struc-
tures 14 (2-3) (1991) 329336.
[14] R.W. Suhr, The eect of surface nish on high cycle fatigue of a
low alloy steel, in: K.J. Miller, E.R. de los Rios (Eds.), Mechan-
ical Engineering Publications, London, 1986, pp. 6986, ISBN 0-
85298-614-7.
[15] J.C. Newman Jr., E.P. Philips, M.H. Swain, Fatigue-life predic-
tion methodology using small-crack theory, International Jour-
nal of Fatigue 21 (1999) 109119.
[16] BS 1134-1. Assessment of surface texture. Part 1: Methods and
instrumentation, 1988.
[17] BS 1134-2. Assessment of surface texture. Part 2: Guidance and
general information, 1990.
[18] B. Griths, Manufacturing Surface TechnologySurface Integ-
rity and Functional Performance, Prenton Press, 2001, ISBN
1-8571-8029-1.
[19] K.J. Stout, P.J. Sullivan, W.P. Dong, E. Mainsah, N. Luo, T.
Zahouani, H. Zahouani, The Development of Methods for the
Characterisation of Roughness in Three Dimensions, Com-
mission of the European Communities, No. EUR 15178 EN,
University of Birmingham, 1993, ISBN 0-7044-1313-2.
[20] K.J. Stout, L. Blunt, Three Dimensional Surface Topography,
Penton Press, 2000, ISBN 1-85718-026-7.
[21] D. Whitehouse, Surfaces and Their Measurement, Hermes Pen-
ton Science, 2002, ISBN 1-9039-9601-5.
[22] Surfstand. Surfstand project Web site parameter description
pages, University of Hudderseld, http://scom.hud.ac.uk/
external/research/ pec/surfstand, 2001.
[23] L. De Chire, P. Leonardo, H. Trumpold, D.A. Lucca, G.
Brown, C.A. Brown, J. Raja, H.N. Hansen, Quantitative charac-
terisation of surface texture, Annals of the CIRP 49 (2) (2000)
635652.
[24] M.R. Bayomi, A.K. Abd El-Latif, Eect of surface nish on
fatigue strength, Engineering Fracture Mechanics 51 (5) (1995)
861870.
[25] G.R. Leverant, B.S. Langer, A. Yuen, S.W. Hopkins, Surface
residual stresses, surface topography and the fatigue behaviour
of Ti6/4, Metallurgical and Material Transactions A10 (1979)
251257.
[26] Surface Texture Analysis, The Handbook, Hommelwerke
GmbH, Muhlhausen, 1992.
[27] K.J. Stout, How smooth is smoothsurface measurements and
their relevance in manufacturing, The Production Engineer
(1980) 1722.
[28] ISO 12085:1996 Geometrical Product Specication (GPS), Sur-
face texture: Prole method, Motif parameters.
[29] L. De Chire, S. Christiansen, S. Skade, Advantages and indus-
trial applications of three-dimensional surface roughness analy-
sis, Annals of the CIRP 43 (1) (1994) 473478.
[30] D.J. Whitehouse, Function maps and the role of surfaces, Inter-
national Journal of Machine Tools and Manufacture 41 (2001)
18471861.
[31] S.O.A. El-Helieby, G.W. Rowe, Inuence of surface roughness
and residual stress on fatigue life of ground steel components,
Metals Technology (1980) 221225.
[32] J. Watanabe, J. Amano, H. Furuichi, Rise in fatigue strength of Ti
by extremely mild abrasive processing, Wear 174 (1994) 235237.
[33] J. Matsumoto, D. Magda, D.W. Hoeppner, T.Y. Kim, Eects of
machining process on the fatigue strength of hardened AISI
4340 steel, Journal of Engineering for Industry 113 (1991)
154159.
[34] J.F. Field, J.T. Kahles, J.T. Cammett, A review of measuring
methods for surface integrity, Annals of the CIRP 21 (1972)
219238.
[35] N. Zlatin, M. Field, Procedures and precautions in machining
titanium alloys, Titanium Science and Technology 1 (1973) 489
504.
[36] C. Sommer, Non-Traditional Machining Handbook, Advance
Publishing, Houston, 2000, ISBN 1-57537-325-4.
[37] S.J. Trail, P. Bowen, Eects of stress concentration on the fati-
gue life of a c-based titanium alloy, Material Science and Engin-
eering: Structural Materials Properties, Microstructure and
Processing A193 (1995) 427434.
[38] J.D. Fordham, R. Pilkington, C.C. Tang, The eect of dierent
proling techniques on the fatigue performance of metallic mem-
branes of AISI 301 and Inconel 718, International Journal of
Fatigue 19 (6) (1997) 487501.
[39] R.A. Hanlon, Investigation into the eect on fatigue life of
abrasive water-jet cutting, MEng thesis, Metallurgy and Materi-
als, The University of Birmingham, Birmingham, 2001.
[40] K. Hirano, K. Enomoto, E. Hayashi, K. Kurosawa, Eects of
water jet peening on corrosion resistance and fatigue strength of
type 304 stainless steel, International Journal of Fatigue 19 (10)
(1997) 734739.
[41] A. Abrao, D.K. Aspinwall, The surface integrity of turned and
ground hardened bearing steel, Wear 196 (1996) 279284.
[42] D.K. Aspinwall, R.C. Dewes, A.L. Mantle, E.-G. Ng, High
speed milling of advanced aerospace alloys and hardened
mould/die steels, MTTA, Warwick, 1998, ISBN 0-907-348-0401.
[43] A.L. Mantle, D.K. Aspinwall, Surface integrity and fatigue life
of turned gamma titanium aluminide, Journal of Materials Pro-
cessing Technology 72 (1997) 413420.
[44] A. Bentley, A.L. Mantle, D.K. Aspinwall, The eect of machin-
ing on the fatigue strength of a gamma titanium aluminide inter-
metallic alloy, Intermetallics 7 (1999) 967969.
D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134 133
[45] A.R.C. Sharman, D.K. Aspinwall, R.C. Dewes, D. Clifton, P.
Bowen, The eects of machined workpiece surface integrity on
the fatigue life of c-titanium aluminide, International Journal of
Machine Tools and Manufacture 41 (2001) 16811685.
[46] M.R. Salmani, Fatigue life of gamma titanium aluminides, PhD
thesis, Metallurgy and Materials, The University of Birming-
ham, Birmingham, 1999.
[47] R.C. Dewes, High speed machining of hardened ferrous alloys,
PhD thesis, Mechanical and Manufacturing Engineering, The
University of Birmingham, Birmingham, 1997.
[48] J.T. Berry, Y. Li, High speed machining and fatigue behaviour,
Proceedings of 4th International Conference on Industrial Tool-
ing, Southampton, 2001, pp. 8290.
[49] S. Gu ngo r, L. Edwards, Eect of surface texture on the
initiation and propagation of small fatigue cracks in a forged
6082 aluminium alloy, Material Science and Engineering A:
Structural Material Properties, Microstructure and Processing
A160 (1993) 1724.
[50] J.S. Eckersley, T.J. Meister, Intelligent design takes advantage of
residual stresses, Proceedings of the Third International Confer-
ence, ASM International, Indianapolis, Indiana, USA, 1991,
pp. 175181.
[51] L. Wagner, J.K. Gregory, Thermomechanical surface treatment
of titanium alloys, Materials Science Forum, Dortmund, Ger-
many, 1994, 163165, 159172 (ISBN 0-87849-678-5).
[52] L. Wagner, Mechanical surface treatments on titanium alloys:
fundamental mechanisms, Proceedings of a Symposium on Sur-
face Performance of Titanium, Cincinnati, Ohio, USA, 1997,
pp. 199216.
[53] L. Wagner, Mechanical surface treatments on titanium, alu-
minium and magnesium alloys, Materials Science and Engineer-
ing A263 (1999) 210216.
[54] R.W. Schutz, Surface treatments for expanding titanium alloy
application limits: an overview, Proceedings of a Symposium on
Surface Performance of Titanium, Cincinnati, Ohio, USA, 1997,
pp. 122, 1997, pp. 122.
[55] A. Eftekhari, J.E. Talia, P.K. Mazumdar, Inuence of surface
condition on the fatigue of an aluminium-lithium alloy (2090-
T3), Materials Science & Engineering A199 (1995) 36.
[56] B.R. Sridhar, K. Ramachandra, K.A. Padmanabhan, Eect of
shot peening on the fatigue and fracture behaviour of two
titanium alloys, Journal of Materials Science 31 (1996) 5953
5960.
[57] A. Drechsler, T. Dorr, L. Wagner, Mechanical surface treat-
ments on Ti-10V-2Fe-3Al for improved fatigue resistance, Mate-
rials Science and Technology A243 (1998) 217220.
[58] M.A.S. Torres, H.J.C. Voorwald, An evaluation of shot-peening,
residual stress and stress relaxation on the fatigue life of AISI
4340 steel, International Journal of Fatigue 24 (2002) 877886.
[59] W.Z. Zhuang, G.R. Halford, Investigation of residual stress
relaxation under cyclic load, International Journal of Fatigue 23
(2001) 3137.
[60] N. Hasegawa, Y. Watanabe, Y. Kato, Eect of shot peening on
fatigue strength of carbon steel at elevated temperature, Inter-
national Conference on Shot Peening, Oxford, UK, 1993,
pp. 157162.
[61] M. Schilling-Praetzel, F. Hegemann, P. Gomez, G. Gottstein,
Inuence of shot peening on fatigue life, International Confer-
ence on Shot Peening, Oxford, UK, 1993, pp. 227238.
[62] S.D. Thompson, D.W. See, C.D. Lykins, P.G. Sampson, Laser
shock peening versus shot peening. A damage tolerance investi-
gation, Proceedings of a Symposium on Surface Performance of
Titanium, Cincinnati, Ohio, USA, 1997, pp. 239251.
[63] A.M. Hassan, A.M.S. Momani, Further improvements in some
properties of shot peened components using the burnishing pro-
cess, International Journal of Machine Tools and Manufacture
40 (2000) 17751786.
134 D. Novovic et al. / International Journal of Machine Tools & Manufacture 44 (2004) 125134

You might also like