Sie sind auf Seite 1von 11

International Journal of Fatigue 24 (2002) 783793

www.elsevier.com/locate/ijfatigue
Estimation methods for fatigue properties of steels under axial and
torsional loading
K.S. Kim
a,*
, X. Chen
b
, C. Han
a
, H.W. Lee
c
a
Department of Mechanical Engineering, Pohang University of Science and Technology, South Korea 790784
b
School of Chemical Engineering, Tianjin University, Peoples Republic of China, 300072
c
Department of Mechanical Engineering, Pusan National University, Pusan, South Korea
Received 7 June 2001; received in revised form 6 August 2001; accepted 27 September 2001
Abstract
Uniaxial and torsional fatigue tests have been conducted on eight steels. The cyclic equivalent stress and strain amplitudes can
be tted by the Ramberg-Osgood relationship. Fatigue lives are found correlated with the equivalent strain amplitude. Seven methods
for estimating uniaxial fatigue properties from tensile properties or hardness have been evaluated. The modied universal slopes
method by Muralidharan and Manson, the uniform material law by Baumel and Seeger and the hardness method by Roessle and
Fatemi predicted over 93% of test cases within the factor of 3 compared with observed lives. These methods are also found
applicable to torsional fatigue with fatigue properties estimated from uniaxial fatigue properties based on the equivalent strain
criterion. 2002 Elsevier Science Ltd. All rights reserved.
Keywords: Uniaxial fatigue; Torsional fatigue; Fatigue properties
1. Introduction
The strain-based fatigue life analysis is routinely per-
formed to assess the fatigue resistance of structural
components [1]. The characterization of fatigue endur-
ance of engineering materials is usually made through
uniaxial fatigue tests, and fatigue properties under such
loading are available for a large body of materials. Often,
however, circumstances are encountered in service
where the fatigue resistance of a component needs to be
answered within a short timeframe but fatigue data for
the material are not available. The situation becomes
even more difcult if the loading condition is multiaxial.
Some multiaxial fatigue criteria [2,3] require torsional
fatigue data to determine all material constants. How-
ever, torsional fatigue data can be found only for a lim-
ited number of materials. Therefore, it would be desir-
able to have an estimation scheme for torsional fatigue
properties from readily available material properties.
* Corresponding author. Tel.: +82-54-279-2182; fax: +82-54-279-
5899.
E-mail address: illini@postech.ac.kr (K.S. Kim).
0142-1123/02/$ - see front matter 2002 Elsevier Science Ltd. All rights reserved.
PII: S0142- 1123( 01) 00190- 6
Over the last few decades, many researchers have
attempted to develop relations between monotonic ten-
sile properties and uniaxial fatigue properties of engin-
eering materials. If reliable relations with reasonable
accuracy can be established, they can serve to provide
fast solutions to fatigue problems without time and cost
involved in fatigue testing. Manson [4] rst proposed
two methods; the four-point correlation method and the
universal slopes method, to estimate the strain-life curve
using monotonic tensile data. Mitchell [5] proposed a
method suitable for steels. Baumel and Seeger [6] pro-
posed the uniform material law for metals. Muralidharan
and Manson [7] proposed a modied universal slopes
method, and Ong [8] suggested a modied four-point
correlation method. Roessle and Fatemi [9] recently pro-
posed a method requiring only hardness and the modulus
of elasticity to estimate uniaxial fatigue properties of
steels. In a study conducted by Park and Song [10], six
methods [48] were evaluated for a total of 138 ferrous
and nonferrous metallic materials. They reported that
those proposed by Baumel and Seeger [6], Muralidharan
and Manson [7], and Ong [8] yielded better predictions
over others. Another assessment of different methods
was carried by Ong [11], in which the ASM data for 49
784 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
Nomenclature
e Strain range in axial fatigue test
s Stress range in axial fatigue test
t Shear stress range in torsional fatigue test
g Shear strain range in torsional fatigue test
N
f
Cycles to failure
s
f
Axial fatigue strength coefcient
e
f
Axial fatigue ductility coefcient
b Axial fatigue strength exponent
c Axial fatigue ductility exponent
t
f
Shear fatigue strength coefcient
g
f
Shear fatigue ductility coefcient
b
o
Shear fatigue strength exponent
c
o
Shear fatigue ductility exponent
K Cyclic strain hardening coefcient
n Cyclic strain hardening exponent
K
0
Cyclic shear strain hardening coefcient
n
0
Cyclic shear strain hardening exponent
s Equivalent stress
e Equivalent strain
E Youngs modulus
G Shear modulus
n Poisson ratio
e
f
True fracture ductility
s
u
Ultimate tensile strength
s
y
Yield strength (0.2%)
RA Reduction in area (%)
EL Elongation
HB Brinell hardness
steels were used with the modied and original versions
of the four-point correlation method [4,8] and the univer-
sal slopes method [4,7] and Mitchells method [5]. It was
concluded in [11] that the modied four-point corre-
lation method and the modied universal slopes method
gave a satisfactory agreement between predicted and
experimental fatigue lives.
In this paper, the results of strain controlled axial
fatigue tests and torsional fatigue tests on eight steels
are presented. Seven aforementioned methods are evalu-
ated for estimating uniaxial fatigue properties from ten-
sile properties or hardness. Also, the approaches based
on the equivalent strain criterion and the maximum shear
strain criterion will be investigated for estimating tor-
sional fatigue properties from uniaxial fatigue properties.
2. Experiment
The test materials used in this investigation are eight
steels purchased in the form of wrought bars. The chemi-
cal compositions of the materials are given in Table 1.
The monotonic tensile properties obtained from solid
specimens with a diameter of 6 mm are listed in Table 2.
The geometry of the fatigue specimen is shown in
Fig. 1. The specimen was gun-drilled and honed through
the center. The gage section of the outside contour was
machined, ground and polished with alumina powder
(0.3m). The effect of residual stresses that could have
been introduced in the machining process has been ignored
in this study.
Fatigue tests were conducted on a servo-hydraulic
MTS axial-torsional materials testing system. All tests
were carried out under strain control. A tension-torsion
extensometer with an axial gage length of 20 mm and
a diameter of 12.5 mm was used to control the strain.
Triangular waveforms were used for both axial and tor-
sional fatigue. Tests were carried out under fully
reversed strain cycling with frequencies in the range of
0.51Hz. The lower frequency was applied to higher
amplitude tests. Failure was dened as a drop of 10%
in load for axial tests, and a drop of 10% in torque for
785 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
Table 1
Chemical composition of test materials (wt.%)
Material C Si Mn P S Cu Ni Cr Mo
SNCM630 0.32 0.25 0.45 0.013 0.017 0.18 2.52 2.54 0.51
SNCM439 0.39 0.24 0.72 0.09 0.018 0.13 1.65 0.67 0.16
SCM440 0.42 0.22 0.71 0.01 0.01 0.13 0.08 1.01 0.22
SCM435 0.38 0.21 0.7 0.015 0.018 0.11 0.09 0.83 0.16
SFNCM85S 0.2 0.25 0.8 0.017 0.008 0.49 0.55 0.19
SF60 0.43 0.18 0.69 0.023 0.007
S45C 0.43 0.18 0.69 0.023 0.007
S25C 0.27 0.24 0.53 0.019 0.002
Table 2
Mechanical properties of test materials.
Material E (GPa) G (GPa) n s
y
(MPa) t
y
(MPa) s
u
(MPa) EL (%) RA (%) HB
SNCM630 196 77 0.273 951 581 1100 19 49 327
SNCM439 208 80 0.296 950 560 1050 13 37 323
SCM440 204 80 0.283 846 440 1000 13 36 319
SCM435 210 81 0.3 795 460 951 18 66 300
SFNCM85S 201 80 0.26 565 340 825 21 66 241
SF60 208 79 0.311 580 274 820 19 53 167
S45C 206 79 0.298 590 341 798 17 39 234
S25C 209 80 0.29 280 182 508 19 52 153
Fig. 1. Specimen geometry (unit: mm).
torsional tests. The stress amplitude was measured in a
preset interval of cycles. The stress amplitude at approxi-
mately half-life, where the stress-strain response was
stable, was used for the cyclic stress-strain curve. At
least seven tests were conducted for each type of axial
and torsional loading at strain amplitudes ranging from
0.2% to 2%.
3. Methods for estimating uniaxial fatigue
properties
The relationship between the applied strain amplitude
and fatigue life under uniaxial loading and torsional
loading can be expressed by Basquin-Cofn-Manson
equations:
e
2

s
f
E
(2N
f
)
b
e
f
(2N
f
)
c
for uniaxial fatigue, (1)
g
2

t
f
G
(2N
f
)
b
0
g
f
(2N
f
)
c
0
for torsional fatigue, (2)
The material constants s
f
, e
f
, b and c are the uniaxial
fatigue properties, and t
f
, g
f
, b
o
, and c
o
are the torsional
fatigue properties. The strain amplitudes of Eq. (1) and
Eq. (2) can be split into elastic and plastic components,
and they can be individually related to life by equating
to the rst and second terms, respectively, on the right.
An outline is given in the following section for esti-
mating uniaxial fatigue properties that are examined in
this study. The equations in the original papers have
been rewritten in the nomenclature of the present paper.
The details how these equations were obtained can be
found not only in the original paper for each method but
also in review papers [10,11]. It is also noted that the
true fracture ductility e
f
in some of the equations may
be obtained from e
f
ln[100/(100RA)].
3.1. Four-point correlation method
Manson [4] proposed the four-point correlation
method to estimate the strain-life curve using monotonic
tensile properties. The four points here include two
points on the elastic strain-life curve and two points on
786 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
the plastic strain-life curve. The fatigue properties are
related to monotonic tensile properties as follows:
s
f

E
2
10
b log 2+ log
2.5s
u
(1+e
f
)
E

, (3a)
b
log

2.5(1+e
f
)
0.9

log[1/(410
5
)]
, (3b)
e
f

1
2
10
c log
1
20
+ log

1
4
e
3/4
f , (3c)
c
1
3
log

0.0132e

1.91

1
3
log

1
4
e
3/4
f
, (3d)
where e

is the elastic strain range at 10


4
cycles and
is estimated by
e

10
b log(410
4
)+ log
2.5s
u
(1+e
f
)
E

(3e)
3.2. Universal slopes method
An alternative approach was also proposed by Manson
[4], in which it was assumed that the slopes of the elastic
strain-life and plastic strain-life curves do not vary with
materials. The fatigue properties are estimated by
s
f
1.9018s
u
, (4a)
b0.12, (4b)
e
f
0.7579e
0.6
f
, (4c)
c0.6. (4d)
3.3. Mitchells method
Mitchell [5] suggested the following equations for
steels with hardness below 500 HB:
s
f
s
u
345(MPa), (5a)
b
1
6
log

2(s
u
+345)
s
u

, (5b)
e
f
e
f
, (5c)
c0.6. (5d)
3.4. Modied four-point correlation method
Ong [8] proposed a modied four-point correlation
method. The estimation equations are given as
s
f
s
u
(1e
f
), (6a)
b
1
6

log

0.16

s
u
E

0.81

log

s
u
E

, (6b)
e
f
e
f
, (6c)
c
1
4
log

0.00737e

e
/2
2.074

loge
f
, (6d)
where e

e
is the elastic strain range at 10
4
cycles and
is given by
e

e
2

s
f
E
[10
2
3
log[0.16(s
u
/E)
0.81
] log(s
f
/E)
], (6e)
where s
f
s
f
.
3.5. Modied universal slopes method
The universal slopes method [4] was modied by
Muralidharan and Manson [7], and the following equa-
tions were proposed:
s
f
E0.623

s
u
E

0.832
, (7a)
b0.09, (7b)
e
f
0.0196e
0.155
f
s
u
E

0.53
, (7c)
c0.56. (7d)
3.6. Uniform material law
This method, proposed by Baumel and Seeger [9],
may be considered as a universal slopes method. How-
ever, it assigns different slopes to the unalloyed and low-
alloy steels, and to aluminium and titanium alloys. Only
the elastic modulus and tensile strength of the material
are needed for estimation of fatigue properties. For unal-
loyed and low-alloy steels, the equations are given by
s
f
1.5s
u
, (8a)
b0.087, (8b)
e
f
0.59y, (8c)
c0.58.
where
y1 for
s
u
E
0.003, (8e)
y1.375125.0
s
u
E
for
s
u
E
0.003. (8f)
Notice that b and c are very close to those of the modi-
ed universal slopes method.
787 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
3.7. Hardness method
Roessle and Fatemi [9] proposed a simple method to
estimate fatigue properties from Brinnell hardness and
elastic modulus for steels with hardness between 150 HB
and 700 HB. This method will be referred to as the
hardness method in this paper. This method has an
advantage in that it only requires hardness and the modu-
lus of elasticity, whereas all other methods described use
tensile properties. Hardness can be obtained from nonde-
structive tests, and the modulus of elasticity is essentially
constant for each class of materials. This makes the hard-
ness method particularly attractive where tensile data are
either not available or destructive tests cannot be done.
The exponents b and c are the same as those of the
modied universal slopes method, i.e. 0.09 and 0.56,
respectively. The remaining properties are
s
f
4.25(HB)225 (MPa), (9a)
e
f

1
E
[0.32(HB)
2
487(HB)191000]. (9b)
4. Results and discussion
4.1. Results of fatigue tests
The fatigue properties of the materials determined
from test results are listed in Table 3. All materials exhi-
bited cyclic softening except S25C, which hardened cyc-
lically. The failure crack size was in the range of 5 mm
to 15 mm for most tests. The crack path, under visual
inspection, was normal to the specimen axis (mostly
stage II growth) in uniaxial fatigue, and in torsional
fatigue the crack grew in the axial direction (stage I
growth) until it branched off at the onset of stage II
growth before the test was stopped. The midlife hyster-
esis loops from fatigue tests were used to determine the
stable cyclic stress-strain properties. The cyclic stress-
strain data can be tted by the Ramberg-Osgood
relationship:
Table 3
Fatigue properties of test materials obtained under uniaxial loading and torsional loading.
Material s
f
(MPa) e
f
b c t
f
(MPa) g
f
b
0
c
0
SNCM630 1270 1.54 0.0732 0.823 858 1.51 0.0606 0.706
SNCM439 1380 1.89 0.0722 0.801 969 3.68 0.0855 0.765
SCM440 1400 0.675 0.0879 0.650 754 0.315 0.0814 0.54
SCM435 1100 0.996 0.067 0.708 512 0.360 0.0454 0.519
SFNCM85S 1040 0.316 0.0924 0.522 533 0.251 0.071 0.406
SF60 978 0.187 0.082 0.439 504 0.286 0.0668 0.417
S45C 1400 0.449 0.107 0.564 630 1.22 0.0802 0.564
S25C 821 0.216 0.0961 0.458 426 0.249 0.0741 0.376
e
2

s
2E

s
2K

1/n
for uniaxial loading, (10)
g
2

t
2G

t
2K
0

1/n
0
for torsional loading, (11)
The cyclic strain hardening coefcients (K,K
0
) and
exponents (n,n
0
) obtained by tting the stress-strain
data are given in Table 4. The cyclic hardening coef-
cients and exponents can be estimated from their
relations to fatigue properties; n=b/c, K=s
f
/e
n
f
,
n
0
=b
0
/c
0
, K
0
=t
f
/g
n
0
f
. The results are also given in Table
4. The agreement with direct regression is found to be
good only for four materials (SCM435, SFNCM85S,
SF60, S25C). Substantial differences are found for other
materials (SNCM630, SNCM439, SCM440, S45C). The
stress amplitudes computed from two sets of K and n
for the latter four materials showed small differences at
low strain amplitudes, but up to 20% differences were
observed at high strain amplitudes. This discrepancy
would be attributed to the scatter in fatigue data, which
will inuence the values of fatigue coefcients and
exponents, and subsequently cyclic hardening coef-
cients and exponents. Thus, one might say that estimat-
ing cyclic stress-strain curves based on fatigue properties
could lead to considerable error in certain situations.
The equivalent stress (s ) and strain (e) in this study
are dened by
s (3/2s
ij
s
ij
)
1/2
, (12)
e(2/3e
e
ij
e
e
ij
)
1/2
(2/3e
p
ij
e
p
ij
)
1/2
, (13)
where s
ij
, e
e
ij
, and e
p
ij
are the deviatoric stress (s
ij
=s
ij

s
kk
d
ij
/3), elastic deviatoric strain (e
e
ij
=e
e
ij
e
e
kk
d
ij
/3), plas-
tic strain, respectively, and repeated indices imply sum-
mation over 1-3. It is obtained that
s =s, e=2(1+n)e
e
/3+e
p
for uniaxial loading (e
e
: elastic
strain, e
p
: plastic strain), and s =t3, e=g/3 for torsional
loading. Fig. 2 gives the equivalent stress-strain data for
all the materials. The solid lines in these gures rep-
resent the stress-strain curves tted by the Ramberg-
Osgood relationship.
It should be mentioned that the shear stress in torsion
788 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
Table 4
Cyclic stable stress-strain properties of test materials
Regression Estimation from fatigue properties
Material K n K
0
n
0
K n K
0
n
0
SNCM630 1,060 0.054 592 0.050 1,220 0.089 828 0.086
SNCM439 1,000 0.066 601 0.072 1,300 0.090 837 0.112
SCM440 1,040 0.094 643 0.108 1,480 0.135 898 0.151
SCM435 1,070 0.089 553 0.085 1,100 0.095 560 0.088
SFNCM85S 1,320 0.180 676 0.173 1,280 0.177 679 0.175
SF60 1,350 0.186 609 0.156 1,340 0.187 616 0.160
S45C 1,150 0.152 552 0.119 1,640 0.190 612 0.142
S25C 1,140 0.210 565 0.199 1,140 0.210 560 0.197
Fig. 2. The equivalent cyclic stress-strain curves of test materials.
was computed with the assumption of uniform distri-
bution through the wall thickness of the specimen. This
assumption may lead to inaccuracy in the stress (9.3%
underestimation if purely elastic). However, in the pres-
ence of substantial plasticity, which is the case for most
test conditions, the error becomes considerably smaller.
It is also noted that the fatigue criteria used in this study
are all based on the measured strain, and the error in
stress has no effect on the predicted lives.
The uniaxial and torsional fatigue lives were corre-
lated with equivalent strain amplitudes in Fig. 3. Good
correlations were found for all materials.
4.2. Uniaxial fatigue properties and life prediction
Life prediction was performed with the fatigue proper-
ties estimated by seven afore-mentioned methods on the
Fig. 3. The equivalent strain amplitudelife relationship of test
materials.
789 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
Fig. 4. Comparison of fatigue lives for uniaxial fatigueexperi-
mental and predicted by the four-point correlation method.
eight materials tested under uniaxial loading. Fig. 4
through 10 give the results for all the methods. The four-
point correlation method by Manson [4] overestimated
life up to the factor of 10 (Fig. 4), and not many points
fell within the bound of factor 3. The universal slopes
method of Manson [4] provided good statistics with 91%
of predicted lives within the factor of 3 (Fig. 5). Mitch-
ells method [5] yielded non-conservative lives for the
Fig. 5. Comparison of fatigue lives for uniaxial fatigueexperi-
mental and predicted by the universal slopes method.
Fig. 6. Comparison of fatigue lives for uniaxial fatigueexperi-
mental and predicted by Mitchells method.
majority of test cases. Only 66% of predicted lives based
on this method were within the factor of 3 (Fig. 6). The
modied four-point correlation method by Ong [8] gave
acceptable predictions for short lives, but overly non-
conservative predictions were obtained for lives greater
than 10
4
reversals (Fig. 7). As much as 24% of predicted
lives were outside of the factor 10 bound for this method.
The modied universal slopes method by Muralidharan
Fig. 7. Comparison of fatigue lives for uniaxial fatigueexperi-
mental and predicted by the modied four-point method.
790 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
Fig. 8. Comparison of fatigue lives for uniaxial fatigueexperi-
mental and predicted by the modied universal slopes method.
and Manson [7] (Fig. 8), the uniform material law by
Baumel and Seeger [6] (Fig. 9) and the hardness method
by Roessle and Fatemi [9] (Fig. 10) gave better predic-
tions than other methods with at least 93% of predicted
lives within the factor 3 bound. Notice that two of these
methods [6,7] are the ones selected for best statistics in
the study by Park and Song [10]. The hardness method
appears to have somewhat larger scatter. The scatter in
the three methods mainly comes from SNCM630 steel.
Fig. 9. Comparison of fatigue lives for uniaxial fatigueexperi-
mental and predicted by the uniform material law.
Fig. 10. Comparison of fatigue lives for uniaxial fatigueexperi-
mental and predicted by the hardness method.
This material had higher contents of nickel and chro-
mium, and highest strength of all test materials. Without
SNCM630 data points, the statistics can be improved
remarkably. A notable feature is that the fatigue ductility
exponent, c, and the fatigue strength exponent, b, are
almost the same in the three methods. Thus, the esti-
mates of b=0.09 and c=0.56 are deemed reasonable
for steels. However, s
f
and e
f
may vary somewhat
depending on the crack size used in the failure criterion
[1214]. This implies that the performance of individual
methods could vary, perhaps to a minor extent, with the
failure criterion used in the baseline fatigue data. With
the failure criterion of 10% drop in load or torque used
in this study, which produced 5 mm to 15 mm cracks,
the modied universal slopes method appears to provide
the best correlation.
4.3. Torsional fatigue properties and life prediction
It has not received much attention how to estimate
torsional fatigue properties from readily available
material properties. Almost all investigations in this area
are concerned with estimating only uniaxial fatigue
properties. This is attributable to the fact that the avail-
ability of torsional fatigue data is limited to a relatively
small number of materials. From the perspective of
multiaxial fatigue analysis, however, the development of
a reliable estimation scheme for torsional fatigue proper-
ties is highly desirable. This is because the fatigue cri-
teria suitable for shear fracture often require torsional
fatigue test data [2,3], while test facilities for such load-
ing are not available as widely as the uniaxial testing
791 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
systems. The utility of such an estimation scheme would
be wide considering the fact that the majority of steels
for structural applications fail in shear mode.
The observation of crack paths under torsional loading
indicated that the materials under study are all shear
fracture materials. The torsional fatigue lives were corre-
lated well with shear strain amplitudes, as given by Eq.
(2). However, when the uniaxial and torsional fatigue
data were plotted together, better consolidation of the
data points was found with the equivalent strain ampli-
tude (Fig. 3), than with the maximum shear strain ampli-
tude for all materials except SCM440. Fig. 3 shows that
the equivalent strain amplitude provides satisfactory cor-
relation for SCM440, but the maximum shear strain
amplitude gave somewhat better results. It may be also
worth noting that the effect on fatigue life of normal
strain (or stress) on the maximum shear strain plane that
is emphasized in widely accepted shear strain criteria
[2,3] was negligibly small for SCM440 [15].
Since the modied universal slopes method [7], the
uniform material law [6] and the hardness method [9]
excelled other methods in uniaxial fatigue, they are used
to evaluate torsional fatigue data in this section. The
fatigue ductility and strength exponents do not change
signicantly between axial and torsional fatigue as Leese
and Morrow [12] observed for 1045 steel. Fatemi and
Kurath [16]s data on In718 and 1045HR show essen-
tially similar trends, although some deviation is found
in fatigue ductility exponents for In718. The regressed
data for the exponents obtained in this study for eight
materials show a trend that both b and c are almost con-
sistently bigger in magnitudes for uniaxial fatigue than
for torsional fatigue (Table 3). These properties are usu-
ally sensitive to the scatter of data points, and the
addition of data points could alter the trend. Our experi-
ence is that the application of the exponents for uniaxial
fatigue to torsion with a proper adjustment of fatigue
coefcients does not alter predicted lives to a signicant
extent. Consequently, the torsional exponents were set
identical to the uniaxial counterparts. Since the equival-
ent strain criterion appears more suitable for the com-
bined axial and torsional life data than the maximum
shear strain criterion, the following estimates were
adopted (see Appendix A):
t
f
s
f
/3, (14a)
g
f
e
f
3, (14b)
b
0
b, (14c)
c
0
c. (14d)
Using these equations, the shear fatigue properties
were obtained from estimated uniaxial fatigue properties,
and the torsional fatigue lives of test materials were pre-
dicted. It was found that all three methods provide
Fig. 11. Comparison of fatigue lives for torsional fatigueexperi-
mental and predicted by the modied universal slopes method.
reasonably good results. The uniform material law [6]
(Fig. 12) and the hardness method [9] (Fig. 13) yielded
better statistics than the modied universal slopes
method [7], which gave slightly larger scatter (Fig. 11).
All three methods did not perform as superbly as in axial
fatigue, but at least 85% of the predicted lives were still
within the factor 3 bound.
Leese and Morrow [12] tried approximations
based on the maximum shear strain criterion
Fig. 12. Comparison of fatigue lives for torsional fatigueexperi-
mental and predicted by the uniform material law.
792 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
Fig. 13. Comparison of fatigue lives for torsional fatigueexperi-
mental and predicted by the hardness method.
(t
f
=s
f
/2, g
f
=1.5e
f
, b
0
=b, c
0
=c; see Appendix A) as
well as the equivalent strain criterion, Eqs. (14a, 14b,
14c and 14d). They used experimental data for uniaxial
fatigue properties. Their results show that both
approaches predict the torsional fatigue curve of 1045
steel closely. The maximum shear strain approach was
also tried in this study. It furnished, on the whole, more
conservative lives compared with the equivalent strain
appraoch. The correlation was found better for SCM440,
almost all data points of which fell within the bound of
factor 2 for all three methods of estimation. The overall
statistics were, however, not as good as the equivalent
strain approach: the percentage of data points within the
factor 3 bound ranged 53% to 71% for the three
methods, and the bound of factor 5 included 84%-86%
of data points.
5. Conclusions
The available estimation methods for uniaxial fatigue
properties from tensile properties or hardness were
evaluated in this study. The best methods were then
extended to estimating torsional fatigue properties. All
test materials used in this study can be regarded as duc-
tile steels, as judged from their ductility. Therefore, the
results obtained in this study may be valid only for duc-
tile steels. The conclusions of this study are as follows:
1. Fatigue tests were conducted on eight steels under
axial and torsional loading. The life data correlated
well with equivalent strain amplitudes. The cyclic
equivalent stress and strain data can be tted by the
Ramberg-Osgood relationship.
2. Seven methods for estimating uniaxial fatigue proper-
ties from readily available mechanical properties were
evaluated. Three methodsthe modied universal
slopes method by Muralidharan and Manson, the uni-
form materials law by Baumel and Seeger, and the
hardness method by Roessle and Fatemihave been
found to provide good life predictions with at least
93% within the factor 3 bound compared with experi-
mental lives.
3. Shear fatigue properties estimated from uniaxial
fatigue properties based on the equivalent strain cri-
terion provided satisfactory life estimates for the
majority of specimens tested under torsion, when the
uniaxial fatigue properties were obtained by the three
methods described in 2. At least 85% of predictions
fell within the factor 3 bound compared with experi-
mental lives.
Acknowledgements
The fatigue data reported in this study were obtained
in a research program of Pohang University of Science
and Technology funded by Pohang Steel Company.
Appendix A
The elastic and plastic parts of the Basquin-Cofn-
Manson equation, Eq. (1), for uniaxial fatigue can be
written as
e
e
2

s
f
E
(2N
f
)
b
, (A1)
e
p
2
e
f
(2N
f
)
c
. (A2)
In terms of the equivalent strain dened by Eq. (13), Eq.
(A1) and Eq. (A2) can be written as
e
e
2

s
f
E

(2N
f
)
b
, (A3)
e
p
2
e
f
(2N
f
)
c
, (A4)
where E

=
3E
2(1+n)
.
The addition of Eq. (A3) and Eq. (A4) provides the
equivalent strain criterion.
For torsional fatigue, the maximum shear strain cri-
terion given by Eq. (2) can be split into
g
e
2

t
f
G
(2N
f
)
b
0
, (A5)
793 K.S. Kim et al. / International Journal of Fatigue 24 (2002) 783793
g
p
2
g
f
(2N
f
)
c
0
. (A6)
In terms of the effective strain amplitude Eq. (A5) and
Eq. (A6) can be written as
e
e
2

t
f
3G
(2N
f
)
b
0
, (A7)
e
p
2

g
f
3
(2N
f
)
c
0
. (A8)
In the equivalent strain approach for estimating shear
fatigue properties, Eq. (14a) through Eq. (14d) are
obtained by setting equal the right sides of Eq. (A3) and
Eq. (A7), and the right sides of Eq. (A4) to Eq. (A8).
For the maximum shear strain approach, Eq. (A5) and
Eq. (A6) for uniaxial fatigue can be written as
e
e
2

t
f
(1+n)G
(2N
f
)
b
0
, (A9)
e
p
2

g
f
1.5
(2N
f
)
c
0
. (A10)
By comparing Eq. (A9) with Eq. (A1), and Eq. (A10)
with Eq. (A2), the relation of shear fatigue properties to
uniaxial fatigue properties can be obtained
(t
f
=s
f
/2, g
f
=1.5e
f
, b
0
=b, c
0
=c).
References
[1] Morrow J, Socie DF. The evolution of fatigue crack initiation
life prediction methods. In: Sherratt F, Sturgeon JB, editors.
Materials, experimentation and design in fatigue. Westbury
House (England): Warwick, 1981:3.
[2] Kandil FA, Brown MW, Miller KJ. Biaxial low-cycle fatigue
fracture of 316 stainless steel at elevated temperatures. Book 280,
London: The Metals Society, 1982:203-210.
[3] Fatemi A, Socie DF. A critical plane approach to multiaxial
fatigue damage including out-of-phase loading. Fatigue Fract Eng
Mater Struct 1988;11:14965.
[4] Manson SS. A complex subject-some simple approximations.
Experimental Mechanics 1965;5:193226.
[5] Mitchel MR. Fatigue and microstructure. American Society for
Metals, Metals Pack (OH), 1979:385
[6] Baumel A Jr., Seeger T. Materials data for cyclic loading, Sup-
plement I. Amsterdam: Elsevier Science Publishers, 1990.
[7] Muralidharan U, Manson SS. Modied universal slopes equation
for estimation of fatigue characteristics. ASME Trans J Engng
Mater and Tech 1988;110:558.
[8] Ong JH. An improved technique for the prediction of axial fatigue
life from tensile data. International Journal of Fatigue
1993;15:2139.
[9] Roessle ML, Fatemi A. Strain-controlled fatigue properties of
steels and some simple approaximations. International Journal of
Fatigue 2000;22:495511.
[10] Park J, Song J. Detailed evaluation of methods for estimation of
fatigue properties. International Journal of Fatigue
1995;17:36573.
[11] Ong JH. An evaluation of existing methods for the prediction of
axial fatigue life from tensile data. International Journal of
Fatigue 1993;15:1329.
[12] Leese GE, Morrow J. Low cycle fatigue properties of a 1045
steel in torsion. In: Miller KJ, Brown MW, editors. Multiaxial
fatigue. ASTM STP853, Philadephia: American Society for Test-
ing and Materials, 1985:482-496.
[13] Dowling NE. Cyclic stress-strain and plastic deformation aspects
of fatigue crack growth. ASTM STP637, Philadephia: American
Society for Testing and Materials, 1977:97-121
[14] Socie DF, Waill LA, Dittmer D. Biaxial fatigue of Inconel 718
including mean stress effect. Multiaxial fatigue, ASTM STP 853,
Philadephia: American Society for Testing and Materials,
1985:463-481.
[15] Han C, Nam KM, Kim KS, Cho IL. Multiaxial fatigue life predic-
tion under irregular loading. Proceedings of the Asian Pacic
Conference on Fracture and Strength 01, Oct. 20-22, 2001
Sendai (Japan);1:400405.
[16] Fatemi A, Kurath P. Multiaxial fatigue life predictions under the
inuence of mean-stresses. Journal of Engineering Materials and
Technology. Transactions of ASME 1988;110:3808.

Das könnte Ihnen auch gefallen