Você está na página 1de 11

International Journal of Fatigue 27 (2005) 10401050

www.elsevier.com/locate/ijfatigue

Application of bi-linear loglog SN model to strain-controlled fatigue


data of aluminum alloys and its effect on life predictions
A. Fatemia,*, A. Plaseieda, A.K. Khosrovanehb, D. Tannerb
a

Department of Mechanical, Industrial, and Manufacturing Engineering, The University of Toledo, Toledo, OH 43606, USA
b
General Motors Corporation, Warren, MI 48090, USA
Received 13 May 2004; received in revised form 23 January 2005; accepted 12 March 2005
Available online 23 May 2005

Abstract
Bi-linear loglog model is applied to stress amplitude versus fatigue life data of 14 aluminum alloys. It is shown that the bi-linear SN
model provides a much better representation of the data than the commonly used linear model for Al alloys. The effects of bi-linear model on
stressstrain, stress-life, and strain-life curves are discussed. Life predictions of aluminum alloys based on linear and bi-linear models are
also compared and discussed. Estimations of the bi-linear fit constants from the linear fit constants are then presented.
q 2005 Elsevier Ltd. All rights reserved.
Keywords: Fatigue of aluminum alloys; Fitting of Al alloys fatigue data; Fatigue properties of Al alloys; Life prediction of Al alloys

1. Introduction
The strain-based approach to fatigue is widely used for
different materials at present. Strain-life fatigue curves,
which are also often called low-cycle fatigue curves, are
plotted on loglog scales and total strain amplitude is
resolved into elastic and plastic strain components based on
data from the steady-state hysteresis loops [1]. Basquin [2]
observed that for steel and copper materials the stress-life
data could be linearized on loglog scale. The line can be
represented by
Ds
Z sa Z sf0 2Nf b
2

(1)

where Ds/2 is true stress amplitude, 2Nf is reversals to


failure, sf0 is fatigue strength coefficient, and b is fatigue
strength exponent. Coffin and Manson found the plastic
strain-life data could also be linearized on loglog scale.
This line can be expressed as
D3p
Z 3f0 2Nf c
2

(2)

* Corresponding author. Tel./fax: C1 419 530 8213.


E-mail address: afatemi@eng.utoledo.edu (A. Fatemi).

0142-1123/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijfatigue.2005.03.003

where D3p/2 is plastic strain amplitude, 3f0 is fatigue ductility


coefficient, and c is fatigue ductility exponent. The total
strain amplitude can then be considered as the summation of
elastic and plastic amplitudes and the resulting strain-life
curve can be expressed as:
D3p
D3
D3
s0
Z 3a Z e C
Z f 2Nf b C 3f0 2Nf c
2
2
2
E

(3)

The life at which elastic and plastic components of strain


are equal is called the transition fatigue life (2Nt). For lives
shorter than 2Nt the deformation is mainly plastic, whereas
for lives longer than 2Nt the deformation is mainly elastic.
Endo and Morrow [3] observed that for 2024-T4 and
7075-T6 aluminum alloys they investigated the usual linear
loglog relations between fatigue life and elastic and plastic
strains do not provide adequate correlations of the test
results. They recommended using actual fatigue data plots
for these materials rather than simple power functions.
Sanders et al. [4] showed that plots of plastic strain
amplitude versus cyclic life for aluminum alloys investigated reflect linearity of the CoffinManson relationship
down to a critical level of plastic strain. This plastic strain
level is alloy dependent and below this critical level there is
a departure from single slope behavior. Therefore, in the
lower plastic strain region, the CoffinManson relationship
does not obey the single slope behavior. They related this
deviation from single slope behavior of a CoffinManson

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050

1041

Nomenclature
b
b 1, b 2

fatigue strength exponent in the linear model


fatigue strength exponent in the bi-linear model
in region I, in region II
c
fatigue ductility exponent
E
modulus of elasticity
Kt
stress concentration factor
K0
cyclic strength coefficient
n0
cyclic strain hardening exponent
2Nf
reversals to failure
2Ns, 2Nt separation, transition fatigue life
Sa
nominal stress amplitude
Sf
fatigue limit

plot to the relative inability of the microstructure to develop


homogeneous slip during low plastic strain cycling. Wong
[5] investigated several aluminum alloys and reported that
the stress amplitude with the logarithm of total life did not
form a linear relationship. For both the plastic strain and
stress amplitudes, the data trends show a distinct downward
concavity. Later, Stephens and Koh [6] reported and
discussed bi-linearity of stress amplitude versus fatigue
life for three types of A356-T6 cast aluminum alloys. They
attributed the cause of this bi-linearity to the somewhat flat
nature of the cyclic stressstrain curve in the inelastic region
of these Al alloys. Bi-linear loglog behavior of elastic
strain versus fatigue life has also occurred in other
aluminum alloys [79]. Ignoring the non-linearity can result
in large data scatter leading to significant inaccuracies in life
predictions.
In this work, based on fatigue data obtained for 14
aluminum alloys, bi-linear loglog stress-life behavior of
these alloys is presented and the effects of this bi-linearity
on cyclic stressstrain, stress-life, and strain-life curves are
discussed. The effect of bi-linearity on life prediction of

Sy, Sy0 monotonic, cyclic yield strength


Su
ultimate tensile strength
3aZD3/2 total strain amplitude
3f0
fatigue ductility coefficient
D3e/2, D3p/2 elastic, plastic strain amplitude
saZDs/2 true stress amplitude
smax
maximum stress
sf0
fatigue strength coefficient in the linear model
0
0
sf1
, sf2 fatigue strength coefficient in the bi-linear model
in region I, in region II
Dss/2 stress amplitude at separation fatigue life in the
bi-linear model

aluminum alloys is then discussed, and estimations of bilinear fit constants from linear fit constants are presented.
Table 1 shows the summary of monotonic tensile and straincontrolled fatigue properties of the 14 aluminum alloys used
in this study. Data for the first six alloys were obtained as a
part of this study, whereas data for other alloys were taken
from [10].
Monotonic tension and constant amplitude fully reversed
fatigue tests for the alloys tested in this study were performed
using test methods specified by ASTM Standards E8 and
E606, respectively [11]. Flat plate specimens with square
cross section and uniform gage section length, as shown in
Fig. 1 were used. The relatively short gage section length was
chosen to prevent buckling during compression in fatigue
tests. A closed-loop servo-hydraulic 50 kN axial load frame
in conjunction with a digital servo-controller and hydraulicwedge grips was used to conduct the tests. Significant effort
was put forth to align the load train and minimize bending.
Total strain was controlled using an extensometer with a gage
length of 6 mm and rated as class B1, according to ASTM
classification. In order to protect the specimen surface from

Table 1
Summary of monotonic tensile and strain-controlled fatigue properties of aluminum alloys
Alloy

Process description

E (GPa)

Sy (MPa)

Su (MPa)

Sy0 (MPa)

sf0 (MPa)

3f0

Sfa (MPa)

6063
A356-T6
6260
6063
5754-NG
6082
AlMg4.5Mn
5456-H311
7475-T761
7075-T6
7075-T651
7075-T7351
2014-T6
2024-T3

Extruded profile
Casting
Extruded profile
Extruded tube
Sheet
Extruded bar
Sheet
Bar
Sheet
Sheet
Sheet
Sheet
Bar
Sheet

73.4
78.1
70.5
71.9
65.7
64.0
71.5
69.1
70.0
72.2
70.0
71.0
69.1
70.3

239
232
220
161
107
290
298
235
414
512

382
463
345

263
303
239
192
253

363
400
475
572

462
511
490

254
291
225
161
239
289
318
377
468
394
539
388
449
429

556
666
469
295
455
611
654
702
983
776
1231
989
776
835

0.74
0.09
27.2
0.91
9.19
1.08
0.45
0.20
4.25
2.57
0.26
6.81
0.27
0.17

K0.107
K0.117
K0.090
K0.069
K0.074
K0.099
K0.089
K0.102
K0.107
K0.095
K0.122
K0.140
K0.091
K0.096

K0.830
K0.610
K1.213
K0.706
K1.001
K0.857
K0.755
K0.655
K1.066
K0.987
K0.806
K1.198
K0.742
K0.644

60.1
59.3
73.0
70.9
97.6
79.0
103.4
85.7
107.0
108.1
98.2
53.9
117.5
114.9

Sf is fatigue limit calculated from sf0 (109)b.

1042

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050

a (log scale)

Linear

'12

Bi-linear

'f
'f1

s/2

b1

b
b2

II

2 Ns

2 Nf (log scale)

Fig. 2. Schematic relationship of sa versus 2Nf for linear and bi-linear


models.
Fig. 1. Specimen geometry and dimensions used for monotonic tension and
fatigue tests conducted in this study. All dimensions are in mm.

the knife-edges of the extensometer, epoxy coating was used


to cushion the attachment. Strain control was used in all tests,
except for some long-life and run-out tests (i.e. where little or
no plastic deformation exits), which were conducted in loadcontrol mode. For these tests, strain control was used initially
to determine the stabilized load. Then load control was used
for the remainder of the test. For the strain-controlled tests,
the applied frequencies ranged from 0.1 to 2.2 Hz, depending
on the applied strain amplitude. For the load-controlled
tests, the frequency was increased to up to 30 Hz in order to
shorten the overall test duration. All tests were conducted
using a triangular waveform. Test data were automatically
recorded at regular intervals throughout each test, and stable
stress and strain amplitude data at about midlife were used
to generate the fatigue properties.

2. Bi-linear SN model and its effects on strain-life


and stressstrain curves
2.1. Bi-linear SN model and its difference with the linear
SN model
To improve the stress-life (SN) as well as the strain-life
(3KN) fatigue models for aluminum alloys, a bi-linear
loglog sa versus 2Nf representation of the data for each
material is used. The data for a given material are divided
into two regions, I and II. Region I includes the fatigue data
for the plastic dominated life regime, whereas region II
includes the fatigue data for the elastic dominated life
regime. A loglog linear least square fit of the data is used
for each region. Fatigue properties for the bi-linear model
0
0
for the two regions are given as sf1
, b1, sf2
, and b2. Fig. 2
shows schematic representations of sa versus 2Nf for both
linear and bi-linear models, including the material properties obtained based on each model. In this figure, the life at

which the two lines for regions I and II of the bi-linear


model intersect is called the separation fatigue life (2Ns),
with the true stress amplitude at that life denoted by Dss/2.
Experimental sa versus 2Nf data as well as linear and
bi-linear fits for four typical Al alloys, out of the 14 alloys
considered, are shown in Fig. 3. As can be seen from this
figure, the commonly used linear fit represented by Eq. (1) is
not satisfactory for any of these alloys. The bi-linear model
provides much better fit for the experimental data for these
alloys. Eq. (1) can, therefore, be replaced by the following
equations:
Ds
0
Z sf1
2Nf b1
2

for region I; 2Nf % 2Ns

(4)

Ds
0
Z sf2
2Nf b2
2

for region II; 2Nf R 2Ns

(5)

Bi-linear loglog SN fatigue properties of the 14


aluminum alloys are given in Table 2. The differences
between fatigue lives based on bi-linear model versus
fatigue lives based on linear model for these alloys are
shown in Fig. 4. In this figure, stress amplitudes for the bilinear model are calculated from Eqs. (4) or (5) at fatigue
lives of 103, 104, 105, 106, and 107 reversals, while for the
linear model fatigue lives are calculated from Eq. (1) by
using the corresponding stress amplitudes calculated from
the bi-linear model. It can be seen that the differences
between the fatigue lives based on the two models can be
significant at both short and long lives for most Al alloys.
Plot of the separation life fatigue strength Dss/2 versus
cyclic yield strength, Sy0 , of the Al alloys considered is
shown in Fig. 5. The cyclic yield strength is obtained from
the cyclic stressstrain curve using an offset method at 0.002
plastic strain amplitude, analogous to the monotonic yield
strength obtained from a tensile stressstrain curve. As can
be seen from this figure, a good correlation exists,
represented by Dss/2Z0.92 Sy0 . Therefore, for stress
amplitudes larger than 92% of the cyclic yield strength of

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050

1043

Fig. 3. True stress amplitude versus reversals to failure for typical aluminum alloys considered. (a) A356-T6; (b) 5754-NG; (c) 7075-T6 and (d) 2014-T6.

a material region I behavior can be expected, while at lower


stress amplitudes region II behavior can be expected.
2.2. The effect of bi-linear loglog SN model
on strain-life curve
Eq. (3) relates true strain amplitude to fatigue life. True
strain amplitude versus reversals to failure plots for four
typical aluminum alloys are shown in Fig. 6, showing

the effect of bi-linearity on strain-life curves. To incorporate


the bi-linearity effect, similar to what Stephens and Koh did
for A356-T6 cast aluminum alloys [6], Eq. (3) is replaced
by:
3a Z

0
sf1
2Nf b1 C 3f0 2Nf c
E

for 2Nf % 2Ns

(6)

3a Z

0
sf2
2Nf b2 C 3f0 2Nf c
E

for 2Nf R 2Ns

(7)

Table 2
Summary of bi-linear loglog stress-life fatigue properties as defined in Fig. 2
Alloy

0
(MPa)
sf1

0
sf2
(MPa)

b1

b2

2Ns

Dss/2 (MPa)

b2/b

0
=sf0
sf1

0
sf2
=sf0

6063
A356-T6
6260
6063
5754-NG
6082
AlMg4.5Mn
5456-H311
7475-T761
7075-T6
7075-T651
7075-T7351
2014-T6
2024-T3

478
390
357
247
339
419
526
559
839
556
706
921
661
651

603
1114
654
388
1273
679
719
1233
2329
985
1276
1003
1860
983

K0.088
K0.046
K0.059
K0.050
K0.041
K0.049
K0.060
K0.056
K0.084
K0.048
K0.042
K0.128
K0.059
K0.057

K0.113
K0.156
K0.115
K0.089
K0.163
K0.107
K0.096
K0.149
K0.184
K0.113
K0.125
K0.143
K0.163
K0.112

8118
12997
50238
101682
50157
3862
5741
5179
25868
5806
1210
312
21345
1823

217
253
189
140
218
280
314
345
359
369
524
442
366
424

1.05
1.34
1.28
1.29
2.20
1.09
1.08
1.47
1.72
1.19
1.03
1.01
1.79
1.17

0.86
0.59
0.76
0.84
0.75
0.69
0.80
0.80
0.85
0.72
0.57
0.93
0.85
0.78

1.08
1.67
1.40
1.32
2.80
1.11
1.10
1.76
2.37
1.27
1.04
1.01
2.40
1.18

1044

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050

1.E+08

1.E+07

2Nf based on bi-linear model

stress effect, as well as to characterize notched (local)


fatigue behavior (i.e. Neubers parameter). Fig. 7 shows
SWT parameter versus reversals to failure for both linear
and bi-linear models for the same four aluminum alloys
shown in Figs. 3 and 6. For the linear model, at a given
life Nf, SWT parameter is calculated by first finding sa
and 3a from Eqs. (1) and (3), respectively, using the
material properties
listed
in Table 1 for each alloy. Then,
p
the expression E3a sa is calculated for Nf. For the bilinear model, sa and 3a for a given life Nf are calculated
from Eqs. (4) and (6) if Nf%Ns, and from Eqs. (5) and (7)
if NfRNs, using the material properties listed in Tables 1
and 2. Note that since mean stress tests were not
conducted, Fig. 7 shows the difference between linear
and bi-linear models when used in conjunction with the
SWT parameter, rather than showing mean stress effects.

6063
A356-T6
6260
6063
5754-NG
6082
ALMg4.5Mn
5456-H311
7475-T761
7075-T6
7075-T651
7075-T7351
2014-T6
2024-T3

1.E+06

1.E+05

1.E+04

1.E+03

1.E+02
1.E+02

1.E+03

1.E+04

1.E+05

1.E+06

1.E+07

1.E+08

2Nf based on linear model


Fig. 4. Fatigue life based on bi-linear SN model versus fatigue life based
on linear SN model for 14 Al alloys considered.

Stephens and Koh [6] used D3/2Z0.004 as the cut-off


between the two bi-linear regions. In this study, however,
according to Eqs. (6) and (7), the cut-off is the separation
life Ns. This life is obtained from the intersection of the two
bi-linear fits and can vary for different Al alloys.
As can be seen from Fig. 6, bi-linear stress-life behavior
has little effect on total strain-life curve in the low cycle
regime. The difference between linear and bi-linear models
can become significant in the high cycle regime, however,
due to the fact that this region is mainly governed by the
elastic strain-life line (i.e. SN line).
The difference between the two models is also
evaluated in the SmithWatsonTopper (SWT) parameter
p
( E3a smax , where smaxZsa for fully reversed loading)
plot. This parameter is often used to incorporate mean

2.3. The effect of bi-linear loglog SN model on cyclic


stressstrain curve and properties
The cyclic stressstrain curve reflects the resistance of a
material to cyclic deformation. Similar to the monotonic
deformation in a tension test, a plot of true stress amplitude
versus true plastic strain amplitude in loglog coordinates
for most metals including aluminum alloys results in a linear
curve represented by [1]

0
D3p n
sa Z K 0
(8)
2
where K 0 and n 0 are the cyclic strength coefficient and cyclic
strain hardening exponent, respectively. Substituting plastic
strain amplitude obtained from Eq. (8) into total true strain
range equation results in the cyclic stressstrain curve
represented by RambergOsgood equation:

 0
D3 D3e D3p
Ds
Ds 1=n
Z
C
C
Z
3a Z
2
2E
2K 0
2
2

600

sa  sa 1=n 0
C
E
K0

(9)

For the aluminum alloys used in this study, K 0 and n 0 are


given in Table 3. Note that the cyclic yield strength of each
alloy is obtained from Eq. (8) by substituting 0.002 for the
plastic strain amplitude and using K 0 and n 0 for the alloy, i.e.
0
Sy0 Z K 0 0:002n .
Alternatively, K 0 and n 0 can be calculated from straincontrolled fatigue properties using [1]:

s/2, MPa

400

200
s / 2= 0.92 S'y

sf0
3f0 b=c

(10)

b
c

(11)

K0 Z

R2 = 0.8816
0
0

200

400

600

S'y , MPa
Fig. 5. Correlation of separation life fatigue strength Dss/2 with cyclic yield
strength Sy0 .

n0 Z

Eqs. (10) and (11) are derived from compatibility


between Eqs. (1), (2) and (8). Values of K 0 and n 0 obtained

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050


(b) 1.00%
A356-T6 Aluminum

Linear model
Bi-linear model
0.10%
1E+2

1E+3

1E+4

1E+5

1E+6

5754-NG Aluminum

True Strain Amplitude, /2, %

True Strain Amplitude, /2, %

(a) 1.00%

1E+7

Linear model
Bi-linear model

0.10%
1E+2

1E+3

7075-T6 Aluminum

Linear model
Bi-linear model
1E+3

1E+4

1E+5

1E+5

1E+6

1E+7

1E+6

1.00%

True Strain Amplitude, /2, %

True Strain Amplitude, /2, %

(d)

1.00%

0.10%
1E+2

1E+4

Reversals to Failure, 2Nf

Reversals to Failure, 2Nf


(c)

1045

1E+7

Reversals to Failure, 2Nf

2014-T6 Aluminum

Linear model
Bi-linear model

0.10%
1E+2

1E+3

1E+4

1E+5

1E+6

1E+7

Reversals to Failure, 2Nf

Fig. 6. True strain amplitude versus reversals to failure for typical aluminum alloys considered. (a) A356-T6; (b) 5754-NG; (c) 7075-T6 and (d) 2014-T6.

from direct fitting of the experimental data and calculated


from the relations in Eqs. (10) and (11) can be very similar
or very different, depending on the goodness of linearized
fits represented by Eqs. (1), (2) and (8) [12]. A large
difference for aluminum alloys indicates that the elastic
strain-life behavior is not well represented by loglog
linearized fits. For the aluminum alloys investigated, large
differences between the two sets of K 0 and the two sets of n 0
values can be seen in Figs. 8(a) and 9(a), respectively.
If K 0 and n 0 values are calculated from Eq. (4) based on
0
bi-linear fatigue properties sf1
and b1 in region I, then by
substituting these properties into Eqs. (10) and (11) we
obtain:
K0 Z

n0 Z

0
sf1
0
3f b1 =c

b1
c

(12)

(13)

Region I of the bi-linear relation (Eq. (4)) is used since


this region represents the region with significant plastic
strain, and therefore, the plastic strain term in Eq. (9) with
K 0 and n 0 as material properties. Values of K 0 and n 0
obtained from Eqs. (12) and (13) as compared to the values
obtained from direct fit of the experimental data are shown
in Figs. 8 and 9. Comparisons of Fig. 8(a) with (b) for K 0 and
Fig. 9(a) with (b) for n 0 , show that correlation of K 0 and n 0
values from direct experimental fit of data with those from

compatibility equations are far better based on region I of


the bi-linear fit (i.e. Eqs. (12) and (13)), than based on the
linear fit (i.e. Eqs. (10) and (11)).
Values of K 0 and n 0 obtained from the compatibility
equations are also listed in Table 3. Note that the differences
between K 0 values as calculated from Eqs. (10) and (12), and
between n 0 values as calculated from Eqs. (11) and (13) are
large for many of the alloys listed in Table 3. For example,
for the A356-T6 alloy, K 0 values from Eqs. (10) and (12)
listed in Table 3 are 1046 and 465 MPa, respectively.
Values of n 0 for this alloy from Eqs. (11) and (13) are listed
as 0.191 and 0.075, respectively. These represent a
difference of 125% between the K 0 values and a difference
of 155% between the n 0 values. Implications of these large
differences between the compatibility equations values
based on linear versus bi-linear fits on life predictions are
presented in Section 3.
Cyclic true stress amplitude versus true strain amplitude
curves for the four typical aluminum alloys are shown in
Fig. 10. This figure shows that the cyclic true stress
amplitude versus true strain amplitude curves obtained
based on K 0 and n 0 from Eqs. (12) and (13) are the same or
very close to the curves based on direct fitting of
experimental data. On the other hand, the curves based on
K 0 and n 0 from Eqs. (10) and (11) can be much different
from the curves based on direct fitting of experimental data.
The difference between linear and bi-linear models in this
figure is more pronounced at high stress amplitudes where

1046

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050

(b) 1000

(a) 1000

100
1E+2

1E+3

1E+4

1E+5

1E+6

Linear
Region I
Region II
Fatigue Data

SWT parameter, MPa

SWT parameter, MPa

Linear
Region I
Region II
Fatigue Data

100
1E+2

1E+7

1E+3

1E+4

1E+5

1E+6

1E+7

Reversals to Failure, 2Nf

Reversals to Failure, 2Nf


(c) 1000

(d) 1000
Linear

Linear

Region I

Region I

Region II

Region II
Fatigue Data

100
1E+2

1E+3

1E+4

1E+5

1E+6

SWT parameter, MPa

SWT parameter, MPa

Fatigue Data

1E+7

100
1E+2

1E+3

Reversals to Failure, 2Nf

1E+4

1E+5

1E+6

1E+7

Reversals to Failure, 2Nf

Fig. 7. SWT parameter versus reversals to failure for typical aluminum alloys considered. (a) A356-T6; (b) 5754-NG; (c) 7075-T6 and (d) 2014-T6.

Table 3
Comparisons of K 0 and n 0 from direct least square fit and from compatibility equations based on linear and bi-linear fits
Alloy

6063
A356-T6
6260
6063
5754-NG
6082
AlMg4.5Mn
5456-H311
7475-T761
7075-T6
7075-T651
7075-T7351
2014-T6
2024-T3

From experimental data fit

From compatibility equations based on


linear model

From compatibility equations based on


bi-linear model

n0

K 0 (MPa)

n 0 Zb/c

K 0 Zsf0 =3f0 b=c


(MPa)

n 0 Zb1/c

0
K 0 Zsf1
=3f0 b1 =c
(MPa)

0.067
0.063
0.047
0.068
0.032
0.051
0.125
0.084
0.059
0.045
0.074
0.094
0.072
0.109

384
430
301
245
294
397
693
636
675
521
852
695
704
843

0.129
0.191
0.074
0.097
0.074
0.115
0.118
0.155
0.100
0.096
0.151
0.117
0.123
0.149

578
1046
367
298
386
605
719
901
850
709
1506
790
912
1082

0.106
0.075
0.048
0.070
0.041
0.057
0.079
0.086
0.078
0.048
0.052
0.107
0.080
0.088

494
465
304
249
310
417
561
641
749
532
756
750
734
759

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050

1400

1400

K' from least sq. fit, MPa

(b) 1600

K' from least sq. fit, MPa

(a) 1600

1200
1000
800
600
400

1047

1200
1000

200

800
600
400
200

0
0

200 400 600 800 1000 1200 1400 1600

K' from compatibility eq.(10), MPa


0

200 400 600 800 1000 1200 1400 1600

K' from compatibility eq. (12), MPa

Fig. 8. K from least square fit versus K from compatibility equation based on (a) linear method (b) bi-linear method.

plastic strains are significant. Therefore, the compatibility


equations based on the bi-linear model are more realistic for
Al alloys, as compared to those based on the linear model.

3. Effect of bi-linearity on life prediction of aluminum


alloys
Fatigue failures in components typically initiate from
notches or at stress concentrations. At such locations, local
plastic deformation usually exists and fatigue life is
controlled by both local stress and strain amplitudes. A
parameter incorporating both stress and strain amplitudes is
Neubers rule. This rule relates notch stress and strain to
nominal stress and strain, and is the most widely used notch
stressstrain analysis model used for life prediction of
notched members [1]. The use of Neubers rule in this work
allows investigation of the bi-linear SN model effects on
strain-life and stressstrain curves simultaneously, rather
than the effect on each curve separately.
To investigate the bi-linearity effects on life predictions,
a notched component with KtZ2.5 is considered. The
strain-life (i.e. local strain) approach based on Neubers
rule is used. Application of the strain-life approach for

3a sa Z

(b)

0.19

0.19

0.17

0.17

0.15

0.15

0.13
0.11
0.09
0.07

Kt Sa 2
E

0.13
0.11
0.09
0.07

0.05

0.05

0.03

0.03

0.01
0.01 0.03 0.05 0.07 0.09 0.11 0.13 0.15 0.17 0.19

0.01
0.01 0.03 0.05 0.07 0.09 0.11 0.13 0.15 0.17 0.19

n' from compatibility eq. (11)

(14)

The ratio of Sa with respect to S 0 was calculated for each


Al alloy to ensure nominal elastic behavior. For all the
materials considered, this ratio at the highest load level (i.e.
corresponding to a life of 103 reversals) was less than 0.85.
Using nominal stress amplitudes based on bi-linear model
and Eq. (14), local stress amplitudes based on the linear
model and Neubers rule were then obtained by combining

n' from least sq. fit

n' from least sq. fit

(a)

the component involves two steps. First, it requires


determination of local (notch) stresses and strains based
on Neubers rule. Life prediction can then be made using the
local stresses and strains, based on the strain-life equation.
For Al alloys considered, strain amplitudes were obtained at
five lives (103, 104, 105, 106, and 107 reversals) from either
Eqs. (6) or (7), as appropriate, based on bi-linear model.
Next, local stress amplitudes were obtained from Eq. (9)
based on bi-linear model, where K 0 and n 0 used in the
equation are obtained from Eqs. (12) and (13), respectively.
Nominal stress amplitudes, Sa, based on Neubers rule were
then obtained from:

n' from compatibility eq. (13)

Fig. 9. n 0 from least square fit versus n 0 from compatibility equation based on (a) linear method; (b) bi-linear method.

1048

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050

(b)

(a) 400

300
250
200
150
100
Data
Based on K' and n' from eq. (8)
Based on K' and n' from eqs. (10) and (11)
Based on K' and n' from eqs. (12) and (13)

50
0
0.0%

0.2%

0.4%

0.6%

0.8%

1.0%

True Stress Amplitude, /2 (MPa)

True Stress Amplitude, /2 (MPa)

350

300

250

200

150

100

0
0.0%

1.2%

450

450

400
350
300
250
200
150
Data
Based on K' and n' from eq. (8)
Based on K' and n' from eqs. (10) and (11)
Based on K' and n' from eqs. (12) and (13)

0
0.0%

0.2%

0.4%

0.6%

0.8%

1.0%

1.2%

True Stress Amplitude, /2 (MPa)

(d) 500

True Stress Amplitude, /2 (MPa)

(c) 500

50

0.2%

0.4%

0.6%

0.8%

1.0%

1.2%

True Strain Amplitude, /2 (%)

True Strain Amplitude, /2 (%)

100

Data
Based on K' and n' from eq. (8)
Based on K' and n' from eqs. (10) and (11)
Based on K' and n' from eqs. (12) and (13)

50

400
350
300
250
200
150
100

Data
Based on K' and n' from eq (8)
Based on K' and n' from eqs. (10) and (11)
Based on K' and n' from eqs. (12) and (13)

50
0
0.0%

0.2%

0.4%

0.6%

0.8%

1.0%

1.2%

True Strain Amplitude, /2 (%)

True Strain Amplitude, /2 (%)

Fig. 10. True stress amplitude versus true strain amplitude for typical aluminum alloys considered. (a) A356-T6; (b) 5754-NG; (c) 7075-T6 and (d) 2014-T6.

Eqs. (14) and (9) to obtain [1]:


 s 1=n0
s2a
K S 2
C sa a0
Z t a
E
K
E

(15)

Note that K 0 and n 0 used in Eq. (15) are obtained from


Eqs. (10) and (11), respectively. Once local stress
amplitude, sa, is obtained from Eq. (15), 3a is then
calculated from Eq. (9), and substituted into Eq. (3) to
find the corresponding life based on the linear model.
Fig. 11 shows fatigue life based on the bi-linear model
versus fatigue life based on the linear model. The
differences between fatigue lives based on the two models
range between factors of 2 to 3 at shorter lives, to more
than an order of magnitude at longer lives. At longer lives
the difference between predicted lives based on the two
models increases, mainly due to increased error in

extrapolation of the linear model strain-life curve to


long lives. This error results from the use of the linear
model at long lives and is on the non-conservative side.
Note that since many components and structures are
designed for long lives (i.e. lives longer than 106 cycles),
the use of the linear model can result in substantial overestimation of the fatigue life, which can lead to premature
fatigue failure. This is due to the fact that, as stated
earlier, the difference between the linear and bi-linear
model life predictions increases at longer lives, with the
linear model over-estimating fatigue lives. This is
particularly important in light of the fact that for
aluminum alloys, the so called fatigue limit or endurance
limit occurs at much longer lives (i.e. typically in the
range of 108109 cycles), as compared to steels (typically
around 106 cycles).

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050


1.E+10

1.E+08
1.E+07
1.E+06

2Nf based on experimental bi-linear model

2Nf based on bi-linear model

1.E+09

1.E+08
6063
A356-T6
6260
6063
5754-NG
6082
ALMg4.5Mn
5456-H311
7475-T761
7075-T6
7075-T651
7075-T7351
2014-T6
2024-T3

1.E+05
1.E+04
1.E+03
1.E+02
1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07 1.E+08 1.E+09 1.E+10

1.E+07

1.E+06

1.E+05

6063
A356-T6
6260
6063
5754-NG
6082
ALMg4.5Mn
5456-H311
7475-T761
7075-T6
7075-T651
7075-T7351
2014-T6
2024-T3

1.E+04
Data with in
Life factor scatter bands
1.E+03

1.E+02
1.E+02

1.E+03

2Nf based on linear model


Fig. 11. The effect of bi-linear model versus linear model on life prediction
of 14 aluminum alloys considered.

4. Prediction of bi-linear fit constants from linear


fit constants
Since strain-life fatigue properties reported in the
literature are generally based on linear fits of the data (i.e.
sf0 and b in Eq. (1)), it is desirable to convert these properties
0
0
to bi-linear properties (i.e. sf1
, b1, sf2
, and b2 in Eqs. (4) and
(5)). Of course if the raw fatigue data are available, these
properties can be directly obtained from bi-linear fits of the
data (i.e. as in Fig. 3). Such raw data, however, are not
generally reported. Therefore, in this section approximations of the bi-linear fatigue constants based on the
commonly reported data for Al alloys are discussed.
As can be seen from Table 3, the bi-linear slope b1 has a
narrow range between K0.041 and K0.06 for 11 of the 14
materials, with an average value of K0.05. Therefore, the bilinear slope b1 can be regarded as a universal constant for Al
alloys. The ratio of bi-linear slope b2 to linear slope b, b2/b,
also has a relatively narrow range between 1.01 and 1.47 for 11
of the 14 materials as shown in Table 3, with an average value
of about 1.2. Therefore, b2 can be estimated as b2Z1.2b.
For all the 14 aluminum alloys considered, the ratio of bi0
linear constant sf1
to the linear constant sf0 , has a relatively
narrow range between 0.57 and 0.93 as shown in Table 3,
0
with an average value of about 0.75. Therefore, sf1
can be
0
0
estimated from sf1 Z 0:75sf . Similarly, for 11 of the 14
0
0
aluminum alloys the ratio sf2
=sf1
has a range between 1.01
and 1.76 (see Table 3), with an average value of about 1.3.
0
0
Therefore, sf2
can be estimated from sf2
Z 1:3sf0 .
The separation fatigue life can then be estimated from:
 0 1=b1Kb2 

sf2
1:3sf0 1=K0:05K1:2b
Z
2NS Z
0
sf1
0:75sf0
p
Z 31=K0:05K1:2b
(16)

1049

1.E+04

80%

90%

94%

1.E+05

1.E+06

1.E+07

1.E+08

2Nf based on estimated bi-linear model


Fig. 12. Fatigue life based on experimental bi-linear model versus fatigue
life based on estimated bi-linear model for 14 Al alloys considered.

The first part of this relation is obtained from intersection


of the two lines represented by Eqs. (4) and (5) at the
separation fatigue life (2Ns). The last part of the relation is
0
0
found by substituting for sf1
Z 0:75sf0 and sf2
Z 1:3sf0 ;
b1 ZK0:05; and b2 Z 1:2b, as shown in Eq. (16).
Predictive capability of bi-linear fit constants from linear
fit constants as presented above, can be evaluated from
Fig. 12, for all the materials considered. In this figure,
fatigue lives based on experimentally obtained bi-linear
constants are compared with fatigue lives based on the
estimated bi-linear constants. To generate this figure, stress
amplitudes were obtained at five lives (103, 104, 105, 106
and 107 reversals, as shown in Fig. 12) from either Eq. (4) or
(5), as appropriate, based on experimentally obtained bi0
0
linear properties listed in Table 2 (i.e. sf1
; sf2
, b1, b2, and
2Ns). These stress amplitudes were then used again in
Eqs. (4) or (5), as appropriate, using the estimated bi-linear
0
0
properties (i.e. sf1
Z 0:75sf0 ; sf2
Z 1:3sf0 , b 1ZK0.05,
b2Z1.2b, and 2Ns from Eq. (16)) to find the life, 2Nf,
based on estimated bi-linear model. As can be seen, 80% of
the data are within a life factor of G2 and 90% of the data
are within a life factor of G3, indicating good agreement.

5. Conclusions
Based on the experimental data presented and the
analysis performed for 14 Al alloys considered, the
following conclusions can be drawn:
1. The bi-linear loglog stress amplitude-life (SN) model
provides a better representation of fatigue behavior for
aluminum alloys, as compared with the commonly used
linear loglog model. The differences between the SN

1050

2.

3.

4.

5.

6.

A. Fatemi et al. / International Journal of Fatigue 27 (2005) 10401050

fatigue lives based on linear versus bi-linear models can


be significant at both short and long lives.
Fatigue strength at the separation fatigue life, where the
transition between the two regions of bi-linear fits
occurs, was found to be a function of the material cyclic
yield strength. For stress amplitudes larger than 92% of
cyclic yield strength of an Al alloy region I behavior can
be expected, while at lower stress amplitudes region II
behavior can be expected.
Bi-linear stress-life behavior generally has little effect on
total strain-life curve in the low cycle regime. However,
the difference between linear and bi-linear models can
become significant in the high cycle regime, due to the
fact that this region is mainly governed by the elastic
strain-life line.
Values of K 0 and n 0 obtained from direct experimental
fits of cyclic stressplastic strain data agree far better
with those obtained from compatibility equations based
on region I of the bi-linear fit, than those obtained based
on the linear fit. Compatibility equations based on the bilinear model are more realistic for Al alloys, as compared
to those based on the linear model.
For notched members, the difference between fatigue
lives based on linear versus bi-linear models range
between factors of 2 to 3 at shorter lives, to more than an
order of magnitude at longer lives. At longer lives the
difference between predicted lives based on the two
models increases, due to increased error in extrapolation
of the linear model strain-life curve to long lives. The
error resulting from the use of the linear model is on the
non-conservative side.
Bi-linear fit constants can be estimated from commonly
available linear fit constants. Fatigue lives obtained
based on the proposed estimations agreed well with those

based on experimental bi-linear fits of the data for the 14


Al alloys considered.

References
[1] Stephens RI, Fatemi A, Stephens RR, Fuchs HO. Metal fatigue in
engineering. 2nd ed. New York: Wiley; 2000.
[2] Basquin OH. The exponential law of endurance tests. Am Soc Testing
Mater Proc 1910;10:62530.
[3] Endo T, Morrow J. Cyclic stressstrain and fatigue behavior of
representative aircraft metals. J Mater 1969;4(1):15975.
[4] Sanders Jr TH, Mauney DA, Staley JT. In: Jaffee RI, Wilcox BA,
editors. Strain control fatigue as a tool to interpret fatigue initiation of
aluminum alloys. Fundamental aspects of structural alloy design. NY,
USA: Plenum Publishing; 1977.
[5] Wong WA. Monotonic and cyclic fatigue properties of automotive
aluminum alloys. SAE technical paper no. 840120 1984.
[6] Stephens RI, Koh SK. In: Stephens RI, editor. Bi-linear loglog elastic
strain-life model for A356-T6 cast aluminum alloy round-robin low
cycle fatigue data. Fatigue and fracture toughness of A356-T6 cast
aluminum alloy. SAE SP-760 1998.
[7] Wigant CC, Stephens RI. Low cycle fatigue of A356-T6 cast
aluminum alloy. SAE technical paper no. 870096 1987.
[8] Stephens RI, Berns HD, Chernenkoff RA, Indig RL, Koh SK,
Lingenfelser DJ, et al. In: Stephens RI, editor. Low cycle fatigue of
A356-T6 cast aluminum alloya round-robin test program. Fatigue
and fracture toughness of A356T6 cast aluminum alloy. SAE SP760 1988.
[9] Wong WA, Bucci RJ, Stentz RH, Conway JB. Tensile and straincontrolled fatigue data for certain aluminum alloys for application in
the transportation industry. SAE technical paper no. 870094 1987.
[10] Boller CHR, Seeger T. Materials data for cyclic loading-part D:
aluminum and titanium alloys. Material science monographs. New
York: Elsevier; 1987.
[11] Annual Book of ASTM Standards. Metals test methods and analytical
procedures. vol. 03.01. West Conshohocken, PA: ASTM; 2003.
[12] Roessle ML, Fatemi A, Khosrovaneh AK. Variation in cyclic
deformation and strain-controlled fatigue properties using different
curve fitting and measurement techniques. SAE technical paper no.
1999-01-0364 1999.

Você também pode gostar