Sie sind auf Seite 1von 11

International Journal of Fatigue 67 (2014) 6272

Contents lists available at ScienceDirect

International Journal of Fatigue


journal homepage: www.elsevier.com/locate/ijfatigue

Multiaxial life predictions in absence of any fatigue properties


Nima Shamsaei a,, Sean A. McKelvey b,1
a
b

Mississippi State University, Department of Mechanical Engineering, 210 Carpenter Hall, Mail Box 9552, Mississippi State, MS 39762, United States
Chrysler Group LLC, 800 Chrysler Dr., Auburn Hills, MI 48326, United States

a r t i c l e

i n f o

Article history:
Received 25 September 2013
Received in revised form 10 February 2014
Accepted 23 February 2014
Available online 6 March 2014
Keywords:
Multiaxial loading
Fatigue life predictions
Classical theories
Critical plane approaches
Non-proportional loading

a b s t r a c t
The aim of this study is to estimate fatigue life of steels and super alloys under multiaxial loading based
on commonly available tensile properties. The state of loading for most components and structures is
multiaxial resulting from multidirectional loading or stress concentrations. Multiaxial fatigue models
have been developed to predict fatigue behavior under multiaxial loading. These models relate multiaxial
stress/strain components to uniaxial fatigue properties in order to predict fatigue life. In this study, MuralidharanManson, BumelSeeger, and RoessleFatemi prediction methods are employed to predict uniaxial fatigue properties based on simple tensile properties in the absence of any fatigue data. Appropriate
multiaxial fatigue models representing the damage mechanism are then used along with the estimated
uniaxial fatigue properties to predict fatigue lives under in-phase and out-of-phase multiaxial loading.
Predictions are compared with experimental multiaxial data for sixteen different steels and super alloys
from literature. Some approximation techniques to predict stress response for in-phase and out-of-phase
loading based on simple tensile properties are also reviewed. Stress estimated based on these approximation techniques are then used in multiaxial fatigue life predictions and results are compared with experimental observations. It is concluded that fatigue life of steels and super alloys under multiaxial loading
may be predicted reasonably well using appropriate damage models only requiring monotonic
properties.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
Multiaxial states of loading are very typical in many industrial
applications. The multiaxial stresses/strains in critical elements
of components and structures can result from multidirectional
loading, stress concentrations due to geometrical complexity, and
residual stresses generated from manufacturing processes. Multiaxial loading can be categorized as in-phase (IP) and out-of-phase
(OP) loading. For in-phase loading, the ratio of torsion to axial loading and principal directions remain xed. However, under out-ofphase loading, principal directions and consequently maximum
shear directions rotate in time.
Fatigue lives under out-of-phase loading are usually shorter
than in-phase loading at the same equivalent strain level. Kanazawa
et al. [1] related the shorter fatigue lives under out-of-phase
(non-proportional) loading to the non-proportional cyclic
hardening phenomenon. They [1] explained this additional nonproportional cyclic hardening phenomenon with the change in slip
Corresponding author. Tel.: +1 662 325 2364; fax: +1 662 325 7223.
E-mail addresses: shamsaei@me.msstate.edu (N. Shamsaei), SM1453@Chrysler.
com (S.A. McKelvey).
1
Tel.: +1 248 944 5083.
http://dx.doi.org/10.1016/j.ijfatigue.2014.02.020
0142-1123/ 2014 Elsevier Ltd. All rights reserved.

plane from one crystallographic slip system to another one resulting from the rotation of maximum shear plane under nonproportional loading. The interaction of active slip systems then
may cause an additional hardening under non-proportional cyclic
loading.
Multiaxial fatigue models can be used to relate multiaxial state
of loading to uniaxial fatigue properties. Classical models, such as
Maximum Principal Strain and von Mises, were rst proposed in
the early twentieth century as failure theories under static or
monotonic loading. These hypotheses were later extended to cyclic
loading and fatigue strength. For tensile failure mode materials, the
Maximum Principal Strain model has been commonly used to predict fatigue life. The Maximum Principal Strain is related to fatigue
properties and life as shown below:

e1;max

r0f
E

2Nf b e0f 2Nf c

where E is modulus of elasticity, 2Nf is the number of reversals to


failure, and r0f , e0f , b, c are the fatigue strength coefcient, fatigue
ductility coefcient, fatigue strength exponent, and fatigue ductility
exponent, respectively.
The von Mises equivalent strain is used for shear failure mode
materials. The equivalent von Mises strain is calculated as:

63

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

Nomenclature
 OP
Dr
Dr/2
Ds/2
e
ef
e1,max
ea
ee
ep

b
c
E
FS
HB
IP
k
K
K0
n
n0
2Nf
OP
SWT

axial fatigue strength exponent


axial fatigue ductility exponent
modulus of elasticity
FatemiSocie
Brinell Hardness
in-phase
material constant in the FS parameter
strength coefcient
cyclic strength coefcient
strain hardening exponent
cyclic strain hardening exponent
reversals to failure
out-of-phase
SmithWatsonTopper
a
non-proportional cyclic hardening coefcient
c
shear strain
Dcmax
maximum shear strain range
Dcmax/2 maximum shear strain amplitude
Dc/2
shear strain amplitude
De
equivalent strain range
Dee
equivalent elastic strain range
Dep
equivalent plastic strain range
De/2
axial strain amplitude
De=2
equivalent strain amplitude
De1,max/2 maximum principal strain amplitude
 IP
Dr
equivalent in-phase stress range

1
ea p

21 m

s
 2
 2
De
3 Dc
 2
1 m
2
2 2
2

e0f
w
k

me
mp
m
r

rmax
1
rmax
n
ru
ry

r IP
r OP
r0f
s

where De/2 is the axial strain amplitude, Dc/2 is the shear strain
 is the equivalent Poissons ratio and can be calcuamplitude, and m
lated from Eq. (3):

m

me Dee mp Dep
De

where ee , ep , and e are equivalent elastic, plastic, and total strains
and me and mp are elastic and plastic Poissons ratios. The von Mises
equivalent strain and fatigue life are related through the following
CofnManson equation (i.e. Eq. (4)), and therefore, this equation
can be used to calculate fatigue life based on the von Mises criterion, when equivalent strain is calculated from Eq. (2):

ea

r0f
E

2Nf b e0f 2Nf c

However, these classical models may only work for proportional or in-phase loading. For the case of non-proportional or
out-of-phase loadings, using classical models often leads to significant errors as these models do not consider the effects of load nonproportionality. Critical plane models which reect the damage
mechanism and predict the failure on the specic critical plane(s)
within the material have been developed over the last few decades
[2]. These models may be used for fatigue life estimations under
both IP and OP loading and also for predicting the direction of crack
initiation. Among all types of critical plane approaches, strain
stress-based models have the advantage of reecting the constitutive behavior of material such as non-proportional cyclic hardening. These models include both a strain component as the driving
parameter and a secondary stress component taking into account
the cyclic hardening due to non-proportionality of loading as well
as mean and residual stresses. SmithWatsonTopper (SWT) [3]

equivalent out-of-phase stress range


axial stress amplitude
shear stress amplitude
axial strain
true fracture strain
maximum principal strain
equivalent strain amplitude
equivalent elastic strain
equivalent plastic strain
axial fatigue ductility coefcient
material constant in BumelSeeger method
shear to axial strain ratio
elastic Poissons ratio
plastic Poissons ratio
equivalent Poissons ratio
normal stress
maximum normal stress on the maximum principal
strain plane
maximum normal stress on the maximum shear strain
plane
ultimate strength
yield strength
in-phase equivalent stress
90 out-of-phase equivalent stress
axial fatigue strength coefcient
shear stress

and FatemiSocie (FS) [4] damage parameters are two examples


of strainstress-based critical plane approaches for tensile and
shear failure mode materials, respectively.
The SmithWatsonTopper (SWT) critical plane model for tensile failure mode materials considers the maximum principal strain
amplitude, De1,max/2, as the primary parameter driving the crack
and the maximum normal stress on the principal plane, rmax
, as
1
the secondary parameter opening the crack and expediting the failure process if tensile, as presented below:

rmax
1

De1;max r02
f

2Nf 2b r0f e0f 2Nf bc


2
E

The FatemiSocie (FS) critical plane model for shear failure


mode materials is expressed as a function of maximum shear strain
amplitude, Dcmax/2, as the primary parameter driving the crack
and maximum normal stress acting on the maximum shear strain
plane, rmax
, as the secondary parameter, as presented by Eq. (6).
n
The maximum normal stress on the maximum shear plane opens
the crack and expedites the failure process if tensile or closes the
crack and retards the failure process if compressive. The uniaxial
form of the FS equation is given as:


 max 
Dcmax
r
1k n
2
ry



r0f
r0f
2Nf b
1 me 2Nf b 1 mp e0f 2Nf c 1 k
E
2r y

where ry is the material monotonic yield strength, and k is a material constant found by tting fatigue data from uniaxial tests to fatigue data from torsion tests.
Fatigue data are not always available and generating fatigue
properties is an expensive process. Furthermore, a slight change
in material chemical composition or any surface enhancements
such as shot peening or hardening may greatly affect the fatigue
behavior. Therefore, developing predictive techniques for fatigue

64

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

properties based on simple and commonly available material


properties has always been of great interest. Kim et al. [5] evaluated seven methods for estimation of fatigue properties for steels
and found the MuralidharanManson modied universal slopes
method [6], the BumelSeeger uniform material law [7], and the
RoessleFatemi hardness method [8] to yield better fatigue life
predictions under uniaxial or torsion loading as compared to the
other estimation techniques for fatigue properties.
Testing facilities for multiaxial loading are not as widely available as for uniaxial loading and generating multiaxial data is even
more time and cost consuming. Therefore, having a comprehensive
and simple procedure for fatigue analysis under multiaxial loading
will be helpful in many applications. This fact was the motivation
for a recent study by Shamsaei and Fatemi [9] to extend the RoessleFatemi [8] hardness method to multiaxial loading by combining it with FS parameter for shear failure mode materials. They
examined this method for ve different materials with Brinell
hardness as the only material property used and reported satisfactory predictions [9]. They also used this method for fatigue life
predictions under variable amplitude multiaxial loading [10] and
reported reasonable results.
In this study, MuralidharanManson, BumelSeeger, and RoessleFatemi approximation methods based on tensile properties
and hardness are used to predict uniaxial fatigue properties of
steels and super alloys in the absence of any fatigue data. Classical
criteria as well as critical plane approaches are then used to predict
fatigue lives for shear and tensile failure mode materials under IP
and OP multiaxial loading based on the approximated uniaxial
fatigue properties. In this paper, rst the experimental data from
literature are reviewed. Next, multiaxial fatigue models are
combined with MuralidharanManson, BumelSeeger, and RoessleFatemi approximation methods to predict fatigue lives for
shear and tensile failure mode materials under multiaxial loading.
The predicted lives are then compared with experimentally observed fatigue lives for a wide range of steels as well as some super
alloys. This analysis allows for two comparisons; rst a comparison
of the classical multiaxial fatigue models to the critical plane approaches, second a comparison of three uniaxial fatigue property
approximation methods implemented into the multiaxial fatigue
models. Then, in order to show how much accuracy is lost by estimating the fatigue properties used in the multiaxial fatigue models, life predictions are also made using the experimentally
determined fatigue properties and results are compared.
This is followed by an attempt to predict stress response under
multiaxial loading from only tensile properties in absence of any
cyclic data. Two recently proposed models from literature are used
to predict stress response under IP and OP loading. Finally, predicted fatigue lives based on these estimated stress responses are
compared with experimentally observed fatigue lives for several
shear and tensile failure mode materials from literature. This type
of approximation could prove to be valuable to the engineering
community, as the only way to determine the stress response under multiaxial loading is through multiaxial fatigue testing, which
is extremely costly and time consuming. While strain can be measured directly, stress has to be calculated from measured load and
knowing the geometry of the part. In most cases, geometries are
complicated and one needs to perform a nite element analysis.
Being able to predict the stress response under multiaxial loading
would allow an engineer to predict multiaxial fatigue life using
only measured strain and tensile properties.

investigated. Some data originates from testing performed by the


coauthor in the past [9,11], but most of the data sets were found
in literature [1224]. The eleven shear failure mode materials include 1050 normalized (N) [9], 1050 quenched and tempered
(QT) [9], and 1050 induction hardened (IH) steels [9], 304L stainless steel (SS) [11], S45C hot rolled steel (HR) [12], 1045 normalized (N) steel [13,14], SNCM630 quenched and tempered steel
(QT) [15], AL6XN stainless steel (SS) [16], 1Cr-18Ni-9Ti stainless
steel (SS) [17], Hastelloy-X super alloy (at room temperature)
[18], and Inconel 718 super alloy [19]. The tensile failure mode
materials include SA333Gr6 carbon steel [20], 304 stainless steel
(SS) [21,22], 310 stainless steel (SS) [23], Haynes 188 super alloy
(at 760 C) [24], and Hastelloy-X super alloy (at 649 C) [18]. Figs. 1
and 2, respectively, presenting cyclic stressstrain curves for shear
and tensile failure mode materials, illustrate the range of materials
used in this study. In addition, Tables 1 and 2 list the monotonic
and cyclic fatigue properties for these shear failure and tensile failure mode materials, respectively. As can be seen in Figs. 1 and 2, for
both shear failure mode materials and tensile failure mode materials, there is a wide range of deformation behavior from stainless
steels to medium carbon steels to super alloys considered in this
study.
Although the uniform material law is applicable to materials
other than steel such as aluminum and titanium, these materials
were not included in this study. At the time when the hardness
method was developed, only steel materials were examined [8].
However, the super alloys used in this study have elastic modulus
and hardness similar to those of steel. As a result, an attempt to apply the hardness method to super alloys was conducted and resulted in reasonable results. On the other hand, aluminums and
titanium alloys have a much lower elastic modulus compared to

2. Experimental data
In this study, multiaxial fatigue data from eleven shear failure
mode materials and ve tensile failure mode materials are

Fig. 1. Cyclic stressstrain curves for (a) shear failure mode materials and (b) close
up of lower strength shear failure mode materials considered in this study.

65

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

than the experimentally observed lives. Similarly, a factor of 5


means that the predicted lives are 5 times longer or 5 times shorter
than the experimentally observed lives. However, only 65% and 35%
of predicted fatigue lives for out-of-phase data fall within scatter
bands of 5 and 3, respectively. It should be noted that the modulus
of elasticity is approximated as 200 GPa for steels and super alloys
in this study.
To take into consideration the damage mechanism as well as
non-proportional cyclic hardening under out-of-phase loadings,
FS critical plane model is used and fatigue properties are estimated
from MuralidharanManson method based on tensile properties as
presented by Eq. (7):

Fig. 2. Cyclic stressstrain curves for tensile failure mode materials considered in
this study.


 max  h
ih
i
Dcmax
r
A2N f 0:09 B2Nf 0:56 1 kC2Nf 0:09
1k n
2
ry

A
steel. In addition aluminum materials have low hardness levels
which are outside the range of applicability of the hardness method. For these reasons, only steels and super alloys are included in
the following analysis.

r0:832
u
32; 000

B 19e0:155
r0:53
f
u
C

3. Mutiaxial life predictions based on tensile properties

2:42r0:832
u
1:17ru  426
0:075ru  32

3.1. MuralidharanManson modied universal slopes method

In order to estimate strain-life fatigue properties based on tensile properties, Muralidharan and Manson [6] modied the universal slopes method, previously proposed by Manson [25] as follows:

The parameter k here is derived from substituting from Eq. (7)


for r0f and b in the following equation for the material constant,
k, in FS model [9]:

r0f 0:623E

r 0:832
u

E
b 0:09; c 0:56

e0f 0:0196e0:155
f

r 0:53
u

;
7

r0:832
2Nf 0:09
u

or k 1

0:31ry
r0f 2Nf b

In Eqs. (8) and (9), yield strength can be estimated from


where ru is ultimate strength and ef is true fracture ductility. Using
von Mises criterion, Eqs. (2) and (4), and estimated fatigue properties based on MuralidharanManson method (i.e. Eq. (7)), 81% of
predicted fatigue lives for shear failure mode materials considered
in this study are within factor of 5 scatter bands and 59% are within
factor of 3 scatter bands, as presented in Fig. 3(a). Note, a factor of 3
means that the predicted lives are 3 times longer or 3 times shorter

ry = 1.17ru  426 [26]. Predicted lives employing Eq. (8) versus


experimental ones for shear failure mode materials are presented
in Fig. 3(b). About 92% of data are within scatter bands of 5 and
82% are within scatter bands of 3. Life predictions for out-of-phase
data resulted in 88% data within scatter bands of 5 and 82% within
scatter bands of 3. Thus, OP life predictions are signicantly improved using the FS critical plane multiaxial fatigue model as

Table 1
Material properties for shear failure mode materials.

Monotonic properties
Modulus of Elasticity, E (GPa)
Shear modulus, G (GPa)
Yield strength, ry (MPa)
Ultimate strength, ru (MPa)
Brinell hardness, HB
Strength coefcient, K (MPa)
Strain hardening exponent, n

1050 N
[9,11]

1050 QT
[9,11]

1050 IH S45C
[9]
[12]

1045 N
[13,14]

SNMC630 304L SS
[15]
[11]

AL6XN SS 1CR-18Ni-9Ti Hastelloy-X Inconel 718


[16]
[17]
[18]
[19]

206
81
421
709
198
1455
0.253

203
81
1009
1164
360
1461
0.060

198
79
2200
2248
565
NA
NA

186
NA
496
770
182
NA
NA

204
80
382
620
189
1185
0.230

196
NA
951
1103
348
NA
NA

195
NA
390
760
185
NA
NA

1027

195
77
208
585
130
680
0.214

193
NA
310
605
160
NA
NA

193
NA
379
767
172
NA
NA

209
78
1160
1850
370
1910
0.080

Cyclic properties
Fatigue strength coefcient, r0f (MPa) 1109

1346

4974

932

1008

1287

1157

1124

838

3950

Fatigue strength exponent, b


Fatigue ductility coefcient, e0f

0.100
0.292

0.062
2.010

0.152
0.529

0.099 0.107
0.359 0.322

0.042
1.168

0.145
0.122

0.115
1.165

0.091
0.807

0.058
0.178

0.151
1.500

Fatigue ductility exponent, c


In-phase cyclic strength coefcient,
K0 (MPa)
In-phase cyclic strain hardening
exponent, n0
Out-of-phase cyclic strength
coefcient, K0 (MPa)
Out-of-phase cyclic strain hardening
exponent, n0

0.456
1480

0.725
1558

0.910
3328

0.519 0.487
1246
1258

0.792
NA

0.394
2939

0.568
1036

0.665
1115

0.446
1047

0.761
1564

0.223

0.123

0.083

0.199

0.208

NA

0.374

0.189

0.130

0.129

0.068

1249

1420

7980

728

1358

NA

5056

6122

1680

NA

NA

0.177

0.113

0.221

0.093

0.208

NA

0.373

0.354

0.158

NA

NA

66

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

Table 2
Material properties for tensile failure mode materials.

Monotonic properties
Modulus of elasticity, E (GPa)
Shear modulus, G (GPa)
Yield strength, ry (MPa)
Ultimate strength, ru (MPa)
Brinell hardness, HB
Strength coefcient (MPa)
Strain hardening exponent
Cyclic properties
Fatigue strength coefcient,

r0f (MPa)

SA333Gr6 [20]

304 SS [21,22]

310 SS [23]

Haynes 188 760 C [24]

Hastelloy-X 649 C [18]

203
NA
302
450
137
NA
NA

183
83
325
650
160
1210
0.193

NA
NA
NA
641
145
NA
NA

170
NA
268
490
130
512
0.093

156
61
244
581
130
NA
NA

921

1000

NA

823

899

Fatigue strength exponent, b


Fatigue ductility coefcient, e0f

0.124
0.392

0.114
0.171

NA
NA

0.082
0.489

0.089
0.084

Fatigue ductility exponent, c


In-phase cyclic strength coefcient, K0 (MPa)
In-phase cyclic strain hardening exponent, n0
Out-of-phase cyclic strength coefcient, K0 (MPa)
Out-of-phase cyclic strain hardening exponent, n0

0.532
1015
0.211
683
0.078

0.402
1660
0.287
4668
0.344

NA
NA
NA
4450
0.338

0.730
891
0.113
851
0.088

0.446
1474
0.199
2562
0.237

Fig. 3. Predicted fatigue lives versus observed fatigue lives with factor of 3 and 5 scatter bands for predictions made using modied universal slopes with (a) von Mises
equivalent strain theory and (b) FatemiSocie critical plane approach, using uniform material law with (c) von Mises equivalent strain theory and (d) FatemiSocie critical
plane approach, and using the hardness method with (e) von Mises equivalent strain theory and (f) FatemiSocie critical plane approach for shear failure mode materials.

compared to predictions based on the von Mises criterion. It should


be noted that all life predictions presented in this paper using FS
model were done using k = 1.
For the ve tensile failure mode materials in this study, Maximum Principal Strain criterion, Eq. (1), with strain-life fatigue
properties estimated from MuralidharanManson modied universal slopes (i.e. Eq. (7)) is used to predict fatigue lives. This

results in 89% of total data and 57% of OP data within scatter bands
of 5, as presented in Fig 4(a). In addition, 70% of all data and 30% of
OP data are within scatter bands of 3.
The SWT critical plane multiaxial model modied based on
MuralidharanManson method as presented by Eq. (10) is also
used here to predict fatigue lives for tensile failure mode materials
and results are presented in Fig. 4(b). Approximately 79% of total

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

data and 43% of OP data are within scatter bands of 5 and the only
material properties required are ultimate strength and true fracture ductility. About 57% of all data and 26% of OP data are within
scatter bands of 3. Comparing these results with predictions from
Maximum Principal Strain theory, the SWT critical plane model
did not improve the predictions.

De1;max
P2Nf 0:18 Q 2Nf 0:65 ;
2
P 0:00012ru1:664 ; Q 61e0:155
r0:302
f
u

rmax
1

10

3.2. BumelSeeger uniform material law method


BumelSeeger [7] proposed a uniform material law method to
estimate fatigue properties based on ultimate strength, ru, and
modulus of elasticity, E, as follows:

r0f 1:5ru ; e0f 0:59w; b 0:087;


ru

11

Fatigue lives were predicted using von Mises criterion, Eqs. (2)
and (4), and estimated fatigue properties using BumelSeeger
method for the shear failure mode materials, as presented in
Fig. 3(c). It should be again mentioned that the modulus of elasticity
is approximated as 200 GPa for steels and super alloys and this value
was used in all life predictions in this study. As can be seen from this
gure, approximately, 73% of total data and 58% of OP data are within scatter bands of 5. About 46% of total data and 25% of OP data are
within scatter bands of 3. To take into consideration the non-proportional cyclic hardening and damage mechanism effects to improve
the predictions, FS critical plane model, Eq. (6), is used and fatigue
properties are estimated from BumelSeeger uniform material
law method (i.e. Eq. (11)), as presented below:


 max  h
ih
i
Dcmax
r
A2Nf 0:087 B2Nf 0:58 1kC2Nf 0:087
1k n
2
ry

ru

ru

; B 0:885w; C
;
103; 000
1:56ru  570
0:248ru  88
or k 1
k
ru 2Nf 0:087

12

The parameter k here is calculated from Eq. (9). In Eq. (12), yield
strength can be estimated from ry = 1.17ru  426; therefore, ultimate strength is the only required material property in this model.
Using this equation, predicted and experimentally observed fatigue
lives for eleven different shear failure mode materials are compared
in Fig. 3(d). Approximately, 85% of total data and 83% of OP data are
within scatter bands of 5. About 79% of all data and 77% of OP data
are within scatter bands of 3; hence, life predictions for OP loading
are signicantly improved using FS critical plane approach.
For tensile failure mode materials, Maximum Principal Strain criterion, Eq. (1), with fatigue properties estimated from BumelSeeger method (i.e. Eq. (11)) are employed in this study and 89% of total
data and 57% of OP data fall within scatter bands of 5, as can be seen
from Fig. 4(c). Approximately 69% of total data and 35% of OP data
are within the factor of 3 scatter bands. To account for non-proportional cyclic hardening effects, SWT critical plane model is also used
and fatigue properties are estimated from Eq. (11) as follows:

rmax
1
P

De1;max
P2Nf 0:174 Q 2Nf 0:667 ;
2

r2u
89; 000

; Q 0:885ru w

Experimental fatigue lives for tensile failure mode materials are


compared with predictions from Eq. (13) and results are presented
in Fig. 4(d). As can be seen from this gure, 88% of total data and
74% of OP data are within scatter bands of 5, respectively. Therefore, fatigue life predictions for tensile failure mode materials under OP loading are improved by using SWT critical plane model
(i.e. Eq. (5)) as compared to the Maximum Principal Strain theory
(i.e. Eq. (1)), if the factor of 5 in life is considered.
3.3. RoessleFatemi hardness method
Roessle and Fatemi [8] proposed a method for estimation of
uniaxial strain-life fatigue properties as a function of Brinell hardness (HB) as follows:

r0f 4:25HB 225; e0f


b 0:09; c 0:56

6 0:003;
c 0:58; w 1 for
E
r 
ru
u
> 0:003
for
w 1:375  125
E
E

13

67

i
1h
0:32HB2  487HB 191; 000 ;
E
14

Using von Mises criterion, Eqs. (2) and (4), and estimated fatigue properties based on Brinell hardness (i.e. Eq. (14)), 82% of data
for shear failure mode materials are within scatter bands of 5, as
presented in Fig. 3(e). However, only 62% of out-of-phase data fall
within scatter bands of 5. Furthermore, about 62% of all data and
27% of OP data are within scatter bands of 3. Shamsaei and Fatemi
[9] combined FS critical plane multiaxial model and Roessle
Fatemi hardness method as presented with Eq. (15) to predict multiaxial fatigue lives for several steels and reported satisfactory results [9,27].


 max  h
ih
i
Dcmax
r
A2Nf 0:09 B2Nf 0:56 1 kC2Nf 0:09
1k n
2
ry
2:53HB 293
0:48HB2  731HB 286; 500
; B
;
200; 000
200; 000
1
C
0:0022HB 0:38

15

In
Eq.
(15),
ry = 0.0044(HB)2 + 1.33(HB) [26] and
k = [0.0003(HB) + 0.0585](2Nf)0.09 [9], or k = 1 can simply be used;
therefore, hardness is the only required material property. More
details regarding these equations can be found in [9,27]. In this
study, the modulus of elasticity is approximated as 200 GPa for
steels. Eq. (15) is also employed here to predict fatigue lives for
shear failure mode materials and results are presented in
Fig. 3(f). In this study, k = 1 was used since the effect of k is not signicant for the uniaxial form of the FS equation. However, a more
accurate k value may be necessary when the shear form of the FS
equation is used. As can be seen in Fig. 3(f), 95% of total data and
92% of OP data are within scatter bands of 5. In addition, 88% of total data and 82% of OP data are within scatter bands of 3. Better OP
fatigue life predictions using the FS critical plane approach can be
explained by the fact that this model represents the failure mechanism and takes into consideration the constitutive behavior of
materials including non-proportional cycle hardening effects.
For tensile failure mode materials, Maximum Principal Strain
criterion is used to predict fatigue lives under multiaxial loading.
Comparing the experimental fatigue lives with predicted ones
using Maximum Principal Strain criterion and fatigue properties
estimated from RoessleFatemi hardness method (Eq. (14)), 82%
of total data and 52% of OP data are within scatter bands of 5, as
presented in Fig. 4(e). Fig. 4(e) also shows that 60% of total data
and only 13% of OP data are within scatter bands of 3.
The SWT critical plane model (i.e. Eq. (5)) for tensile failure
mode materials takes into account material constitutive behavior
including mean and residual stresses as well as non-proportional

68

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

Fig. 4. Predicted fatigue lives versus observed fatigue lives with factor of 3 and 5 scatter bands for predictions made using modied universal slopes with (a) Maximum
Principal Strain theory and (b) SmithWatsonTopper critical plane approach, using uniform material law with (c) Maximum Principal Strain theory and (d) SmithWatson
Topper critical plane approach, and using the hardness method with (e) Maximum Principal Strain theory and (f) SmithWatsonTopper critical plane approach for tensile
failure mode materials.

cycle hardening. The SWT multiaxial model can be rewritten based


on only hardness using RoessleFatemi hardness method (i.e. Eq.
(14)) as follows:

De1;max
P2N f 0:18 Q2Nf 0:65 ;
2
P 0:0854HB  12:36; Q 0:0013HB2 0:22HB 567 16

rmax
1

Note, P in Eq. (16) has been simplied through curve tting


from a 2nd order polynomial to a linear equation and Q was reduced from a 3rd order polynomial to a 2nd order polynomial. Predicted fatigue lives based on Eq. (16) for ve tensile failure mode
materials are compared with experimental lives in Fig. 4(f). As
can be seen from this gure, 91% of data are predicted within scatter bands of 5 based only on the hardness level of material. Comparing the results, 52% and 61% of OP data are predicted within
scatter bands of 5 using Maximum Principal Strain criterion and
SWT critical plane model, receptively. Furthermore, 48% of OP data
were predicted within factor of 3 scatter bands using SWT, but only
13% of OP data was predicted within the factor of 3 scatter bands
using Maximum Principal Strain criterion. Therefore, better fatigue
life predictions are obtained using critical plane approaches combined with the RoessleFatemi hardness method [8].
The strainstress-based critical plane approaches such as FS and
SWT may work for both short and long life regimes due to their
strain term. They are also based on failure mechanism of material
and represent the physical nature of damage process as well as the
multiaxial load interaction effects [28]. The stress term in these

critical plane approaches takes into consideration the constitutive


behavior of materials including non-proportional cyclic hardening
phenomenon. Nevertheless, the application of these multiaxial approaches is not limited to materials with cyclic hardening behavior, as 1050 QT and 1050 IH steels are materials with cyclic
softening behavior [9,29]. Classical theories such as Maximum
Principal and von Mises criteria, however, are not sensitive to multiaxial load interaction and path. These criteria do not consider the
additional cyclic hardening under non-proportional loading and often result in signicant errors under OP loading [28].
Less accurate predictions obtained using SWT critical plane approach compared to Maximum Principal Strain theory while fatigue properties estimated from MuralidharanManson method,
may be explained by the damage mechanism for tensile failure
mode materials. In SWT model, maximum principal strain and normal stress on the maximum principal plane are parallel; therefore,
the only advantage of the normal stress in this equation is to reect
the constitutive behavior of material. The effect of normal stress on
the failure mechanism of the tensile failure mode materials may
also depend on the crack tip plasticity induced closure which is
not included in SWT parameter. This, along with the fact that the
fatigue parameters are estimated and are not measured from
experimental data, may result in poor predictions when SWT is
combined with MuralidharanManson method. In contrast, normal
stress in FS parameter not only represents the constitutive behavior of materials, but it also takes into consideration the load interaction effects, which has shown recently to be one of the main

69

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

reasons for the shorter fatigue life under non-proportional loading


[28,30]. The plasticity induced closer is also adjusted by the material constant k in the FS parameter, as k is larger for longer lives and
harder materials [9].
Based on the results presented in Figs. 3 and 4, using appropriate
critical plane multiaxial model in conjunction with fatigue properties estimated from RoessleFatemi hardness method, yield better
fatigue life predictions compared to MuralidharanManson and
BumelSeeger estimation techniques for fatigue properties. Furthermore, hardness may be nondestructively measured, even for
in-service components, while measuring ultimate strength requires
a sample specimen and is a destructive test. If the failure mechanism
(i.e. shear failure mode or tensile failure mode) is unknown, fatigue
lives can then be calculated based on both failure modes and the
lower fatigue life may be used for a more conservative design.
Fig. 5(a) shows the multiaxial fatigue life predictions for shear
failure mode materials made using the experimental values of fatigue properties with FS model. For comparison purposes, Fig. 5(b)
shows the multiaxial fatigue life predictions for shear failure mode
materials using the fatigue properties predicted from the hardness
method with the FS method. Note, Fig. 5(b) is identical to Fig. 3(f),
and is repeated here for comparison purposes. Using experimental
values for fatigue properties, 95% of total data was within a factor
of 5 scatter bands and 89% of total data was within scatter bands of
3, as can be seen in Fig. 5(a). In addition, 92% of OP data were within scatter bands of 5 and 83% of OP data were within scatter bands
of 3. Comparing Fig. 5(a) and (b), it can be seen that the accuracy of
the fatigue life predictions was not signicantly affected when
using estimated fatigue properties as opposed to experimental values for fatigue properties.
Fig. 6 shows a similar comparison for the tensile failure mode
materials. Fig. 6(a) shows the results of the multiaxial fatigue life
predictions for tensile failure mode materials made using experimental values of fatigue properties with SWT model. Fig. 6(b) shows
the multiaxial fatigue life predictions for tensile failure mode materials made using fatigue properties estimated from the hardness
method with SWT model. It should be noted that Fig. 6(b) is similar
to Fig. 4(f); however, it does not contain predictions for 310SS, due to
lack of material properties available in literature. Using experimental values for fatigue properties (presented in Fig. 6(a)), 95% of total
data was within the scatter bands of 5 compared to 91% of total data
in scatter bands of 5 using the hardness approximation method for
fatigue properties (presented in Fig. 6(b)). Similarly, using experimental values for fatigue properties, 83% of total data was within
factor of 3 scatter bands. No improvement was seen in life predictions for OP data using experimental fatigue properties compared
to using fatigue properties estimated from the hardness method. It
can be seen in Fig. 6 that multiaxial fatigue life predictions for three
of the materials, SA333Gr6, Haynes 188, and Hastelloy-X, were improved using the experimental values for fatigue properties. However, 304SS showed little difference in multiaxial fatigue life
predictions when using experimental values for fatigue properties
compared to using fatigue properties estimated from the hardness
method. This could be due to the fact that the three materials,
SA333Gr6, Haynes 188, and Hastelloy-X, had hardness levels that
were slightly below the recommended range applicability of the
hardness method. The 304SS material hardness level was inside
the range of applicability of the hardness method.
It should be noted that the hardness method was developed by
Roessle and Fatemi using a minimum hardness of 150 HB. This
could explain the slightly less accurate life predictions for materials below this hardness level. For example, a shear failure mode
material 304L (SS), with a hardness of 130 HB, shows some predictions outside the factor of 5 scatter bands (this can be seen in
Fig. 3(f)). The same observation can be made also from Fig. 4(f),
for the SA 333 Gr6 tensile failure mode material with a hardness

of 137 HB. This suggests a need for developing another t for materials with hardness below 150 HB.
The life predictions for 1050 (IH) steel with a hardness of 565
HB, the highest strength material used in this study, had the worst
correlation with experimental data as seen in Fig. 3(f). However,
this material showed signicant scatter in experimental fatigue
test results [9]. This is due to the increased sensitivity of high hardness materials to impurities. As a result of the scatter in experimental data, the predictions were not expected to have great
correlation. Although the strainstress-based multiaxial fatigue
models such as FS and SWT may take into consideration the constitutive behaviors of material and load interaction effects [28], they
do not fully reect microstructural properties, and therefore, cannot explain the data scatters observed for different materials. To
account for such scatters in fatigue data, there is a need for developing microstructural sensitive multiaxial fatigue models. However, these microstructural sensitive fatigue models may require
signicantly more material constants [31,32] as compared to fatigue models which are only based on stress and/or strain.
4. Stress response predictions from tensile properties
One of the main advantages of using strainstress-based critical
plane approaches is their capabilities of taking into consideration
the effects of constitutive behavior of materials by the stress term.
However, one of the challenges of using strainstress-based critical
plane approaches such as FS and SWT is then requiring the stress
response of the material under multiaxial loading. Lopez and
Fatemi [33] recently developed a model to predict uniaxial cyclic
deformation of steels (i.e. RambergOsgood type equation)
employing common tensile properties and hardness. Since inphase multiaxial fatigue behavior is usually similar to uniaxial fatigue behavior, the LopezFatemi model can also be used to predict
cyclic deformation under in-phase multiaxial loading. Lopez and
Fatemi developed several equations for predicting the cyclic deformation behavior, but the following, based on ultimate and yield
strengths, provided the best correlation with the experimental
data, and therefore, was used in this study:

K 0 1:16ru 593 for


n0 0:37 log

ru =ry > 1:2

0:75ry 82
1:16ru 593

for

ru =ry > 1:2

K 0 3:0  104 ru 2 0:23ru 619 for


0

n 0:37log

ru =ry 6 1:2

3:0  104 ry 2  0:15ry 526

3:0  104 ru 2 0:23ru 619

for ru =ry 6 1:2


0

17
0

where K is the cyclic strength coefcient and n is the cyclic strain


hardening exponent.
Employing Eq. (17), the IP equivalent stress amplitude response,
 IP =2, can be related to IP equivalent strain amplitude, DeIP =2, by
Dr
the following RambergOsgood form equation:


1
 IP
 IP n0
DeIP Dr
Dr

0
2
2E
2K

18

It should be noted that a von Mises relationship is used to determine the in-phase and out-of-phase stress and strain amplitudes.
Details on deriving the in-phase strain amplitudes and stress
amplitudes used in Eq. (18) can be found in [9]. Derivation of the
out-of-phase stress and strain amplitudes are also discussed in [9].
In order to predict stress response under nonproportional multiaxial loading, Shamsaei and Fatemi [29] proposed a predictive

70

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

Fig. 5. Predicted fatigue lives using FatemiSocie method versus experimentally


observed fatigue lives for shear failure mode materials based on (a) experimental
fatigue properties and (b) fatigue properties estimated using hardness method.

model for the non-proportional cyclic hardening coefcient based


on the comparison of the monotonic and uniaxial cyclic deformation curves. This model is dened as follows:

 2  2nn0
  nn0
K
De
K
De

3:8
2:2
0
0
2
2
K
K

a 1:6

19

where De=2 is the equivalent strain amplitude, K is the strength


coefcient, and n is the strain hardening coefcient.
0
It should be noted that, if n and n for a specic material are not
very much different, then monotonic and cyclic stressstrain
curves are almost parallel. Therefore, the cyclic hardening coefcient and consequently the non-proportional cyclic hardening
coefcient are not a function of strain, but they are constant. The
0
strength coefcient, K, and cyclic strength coefcient, K , in Eq.
(19), which are intercepts on the loglog scale, are more dominant
variables in dening the level of non-proportional cyclic harden0
ing. In the case that n and n are very similar, a will be only a func0
tion of K and K .
The non-proportional cyclic hardening coefcient, a, is dened
as:


Dr
a OP  1
DrIP

20

 OP =2 is 90 out-of-phase equivalent stress amplitude and


where Dr
 IP =2 is in-phase equivalent stress amplitude at the same strain
Dr
amplitude level from a plot of equivalent stress amplitude versus

Fig. 6. Predicted fatigue lives using SWT method versus experimentally observed
fatigue lives for tensile failure mode materials based on (a) experimental fatigue
properties and (b) fatigue properties estimated using hardness method.

equivalent strain amplitude. Derivation of the IP and OP stress


amplitudes is described in [9].
In this study, the LopezFatemi method is utilized to generate
the in-phase cyclic stressstrain curves knowing only commonly
available or easily obtainable tensile properties, as presented by
Eq. (17). Using Eq. (19), with the cyclic deformation properties, K0
and n0 , calculated from Eq. (17) and experimentally determined
monotonic deformation properties, K and n, the non-proportional
cyclic hardening coefcient was estimated for a given equivalent
strain amplitude. The equivalent von Mises stress amplitude for
out-of-phase loading was then determined using this estimated
non-proportional cyclic hardening coefcient and Eq. (20). Since
an equivalent stress was calculated for out-of-phase loading, it
was necessary to separate the axial stress, r, and shear stress, s,
for the critical plane search. This was done using the following
relationship for the 90 out-of-phase multiaxial loading:

 
Ds Dr k

2
2 3

21

where k is the ratio of the shear strain to the axial strain.


Next, a critical plane search was performed using the calculated
axial and shear stress values along with known (i.e. measured)
axial and shear strain values, followed by the FS damage calcula-

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

71

Fig. 7. Predicted fatigue lives using FatemiSocie and hardness methods versus
experimentally observed fatigue lives for shear failure mode materials based on (a)
estimated stress response and (b) measured stress response.

Fig. 8. Predicted fatigue lives using SmithWatsonTopper and hardness methods


versus experimentally observed fatigue lives for tensile failure mode materials
based on (a) estimated stress response and (b) measured stress response.

tion. For life predictions, the RoessleFatemi hardness method was


used to determine the uniaxial fatigue properties. Therefore, the
only required information for these predictions were hardness
and tensile properties.
Fig. 7 shows life predictions based on FS critical plane approach
with the hardness method for ve of the eleven shear failure mode
materials, 1050 (N), 1050 (QT), 1045 (N), 304L (SS), and Inconel
718 for which the required monotonic properties were available.
Fig. 7(a) presents the life prediction based on the estimated stress
response. For comparison, Fig. 7(b) provides the life predictions
based on measured stress response. It can be seen in Fig. 7 that
the additional approximation of stress response did not reduce
the accuracy of the life prediction. It should be mentioned here that
the other six shear failure mode materials were omitted from this
comparison due to the lack of monotonic properties K and n.
Although these properties are easily determined through tensile
testing, they were not included in the literature.
Fig. 8 presents a similar comparison for two of the tensile failure
mode materials, 304 (SS) and Haynes 188 for which the monotonic
properties were reported in the literature. Again, the other tensile
failure mode materials were omitted from this part of analysis due
to the fact that the monotonic properties K and n, were not included in the literature. The life predictions in Fig. 8 were done
using SWT along with the RoessleFatemi hardness method.
Fig. 8(a) shows the life predictions based on the estimated stresses

and Fig. 8(b) shows the life predictions based on the measured
stress response. It can be seen in these gures that the predictions
made using the estimated stress response were as good as the predictions made with the measured stress response. These observations can be explained by the fact that maximum normal stress
in both FS and SWT critical plane multiaxial fatigue models is a secondary term. Strain terms in both models are the primary parameters driving the crack and the maximum normal stress is the
secondary parameter opening the crack and accelerating the damage if tensile. Therefore, the effects of normal stress in these multiaxial fatigue parameters are not as signicant as the effects of
strain terms.
5. Conclusions
The following conclusions can be made from the analyses performed in this study:
(1) Acceptable multiaxial fatigue life predictions were obtained
employing FatemiSocie strainstress-based critical plane
approach and fatigue properties estimated from Roessle
Fatemi, MuralidharanManson, or BumelSeeger method.
Therefore, using an appropriate multiaxial fatigue model
seems to be more essential than the selected approximation
technique for uniaxial properties.

72

N. Shamsaei, S.A. McKelvey / International Journal of Fatigue 67 (2014) 6272

(2) FatemiSocie critical plane model always provided signicantly better multiaxial fatigue life predictions for shear failure mode materials compared to von Mises criterion no
matter which method was used to predict uniaxial fatigue
properties.
(3) SmithWatsonTopper critical plane approach also
improved fatigue life predictions for tensile failure mode
materials compared to Maximum Principal Strain criterion
when it was combined with RoessleFatemi and Bumel
Seeger methods. However, when fatigue properties were
estimated from MuralidharanManson modied universal
slopes method, Maximum Principal Strain criterion provided
better fatigue life predictions than SmithWatsonTopper
critical plane model.
(4) Fatigue properties estimated from RoessleFatemi hardness
method, when used in the appropriate critical plane multiaxial model (i.e. FatemiSocie and SmithWatsonTopper)
yield better multiaxial fatigue life predictions compared to
other estimation methods for fatigue properties. In addition,
hardness can be measured nondestructively even for in-service components, whereas measuring ultimate strength is a
destructive test requiring a sample specimen.
(5) Stress response under in-phase and 90 out-of-phase loading
may also be approximated only knowing tensile properties.
Life predictions made using approximated stress response
from tensile properties were found to be as good as predictions made with measured stress response.

Acknowledgments
Chrysler Group LLC provided the support for this research
study. Authors would like to express their gratitude to Professor
Ali Fatemi from The University of Toledo, Professor Darrell Socie
from University of Illinois at Urbana-Champaign, and Dr. Yung-Li
Lee from Chrysler Group LLC for very useful discussions on this
subject.
References
[1] Kanazawa K, Miller KJ, Brown MW. Cyclic deformation of 1% Cr-Mo-V steel
under out-of-phase loads. Fatigue Fract Eng Mater Struct 1979;2:21728.
[2] Socie DF, Marquis GB. Multiaxial fatigue. SAE Inc.; 2000.
[3] Smith RN, Watson PP, Topper TH. A stress-strain parameter for fatigue of
metals. J Mater 1970;5:76778.
[4] 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.
[5] Kim KS, Chen X, Han C, Lee HW. Estimation methods for fatigue properties of
steel under axial and torsional loading. Int J Fatigue 2002;24:78393.
[6] Muralidharan U, Manson SS. Modied universal slopes equation for estimation
of fatigue characteristics. ASME Eng. Mater. Tech. 1988;110:558.

[7] Bumel Jr A, Seeger T. Materials data for cyclic loading: supplement


I. Amsterdam: Elsevier Science Publishers; 1990.
[8] Roessle ML, Fatemi A. Strain-controlled fatigue properties of steels and some
simple approximations. Int J Fatigue 2002;22:495511.
[9] Shamsaei N, Fatemi A. Effect of hardness on multiaxial fatigue behavior and
some simple approximations for steels. Fatigue Fract Eng Mater Struct
2009;32:63146.
[10] Shamsaei N, Fatemi A. Fatigue life predictions under general multiaxial loading
based on simple material properties. SAE Int J Mater Manuf 2011;4:6518.
[11] Shamsaei N, Fatemi A, Socie DF. Multiaxial fatigue evaluation using
discriminating strain paths. Int J Fatigue 2011;33:597609.
[12] Kim KS, Park JC, Lee JW. Multiaxial fatigue under variable amplitude loads.
ASME Eng Mater Technol 1999;121:28693.
[13] Fatemi A, Stephens R. Biaxial fatigue of 1045 steel under in-phase and 90
degree out-of-phase loading conditions. In: Multiaxial fatigue: analysis and
experiment. SAE AE-14; 1989. p. 12137.
[14] Kurath P, Jiang Y, Fatemi A. Strain-path inuence on multiaxial deformation
and fatigue damage. In: Multiaxial fatigue of an induction hardened shaft. SAE
AE-28; 1999. p. 11740.
[15] Lee B, Kim K, Nam K. Fatigue analysis under variable amplitude loading using
an energy parameter. Int J Fatigue 2002;25:62131.
[16] Kalanus S, Jiang Y. Fatigue of AL6XN stainless steel. ASME J Eng Mater Technol
2008;130:112.
[17] Chen X, An K, Kim K. Low-cycle fatigue of 1Cr-18Ni-9Ti stainless steel and
related weld metal under axial, torsional and 90 out-of-phase loading. Fatigue
Fract Eng Mater Struct 2004;27:43948.
[18] Jordan E. Elevated temperature biaxial fatigue. NASA Contractor Report:
175009; 1985.
[19] Fatemi A, Kurath P. Multiaxial fatigue life predictions under the inuence of
mean stresses. ASME J Eng Mater Technol 1988;110:3808.
[20] Arora P, Gupta S, Bhasin V, Vaze K, Sivaprasad S, Tarafdar S. Multiaxial fatigue
studies on carbon steel piping material of Indian PHWRS. In: SMiRT 21. Paper
No. 479; 2011.
[21] Socie D. Multiaxial fatigue damage models. ASME Eng Mater Technol
1987;109:2938.
[22] Itoh T, Sakane M, Ohnami M, Socie D. Nonproportional low cycle fatigue
criterion for type 304 stainless steel. ASME Eng Mater Technol
1995;117:28592.
[23] Doong S, Socie D, Robertson I. Dislocation substructures and nonproportional
hardening. ASME Eng Mater Technol 1990;112:45664.
[24] Kalluri S, Bonacuse P. In-phase and out-of-phase axial-torsional fatigue
behavior of Haynes 188 at 760C. NASA Technical Memorandum: 105765;
1991.
[25] Manson SS. A complex subject-some simple approximations. Exp Mech
1965;5:193226.
[26] Roessle ML. Correlations among microstructural parameters, tensile data, and
fatigue properties for steels. M.Sc. Thesis, The University of Toledo, Toledo,
Ohio; 1998.
[27] Shamsaei N. Multiaxial fatigue and deformation including non-proportional
hardening and variable amplitude loading effects. PhD Dissertation, The
University of Toledo, Ohio; 2010.
[28] Fatemi A, Shamsaei N. Multiaxial fatigue: an overview and some
approximation models for life estimation. Int J Fatigue 2011;33:94858.
[29] Shamsaei N, Fatemi A. Effect of microstructure and hardening on nonproportional cyclic hardening coefcient and predictions. Mater Sci Eng AStruct 2010;527:301524.
[30] Shamsaei N, Fatemi A. Small fatigue crack growth under multiaxial stresses.
Int J Fatigue 2014;58:12635.
[31] McDowell D, Gall K, Horstemeyer M, Fan J. Microstructure-based fatigue
modeling of cast A356T6 alloy. Eng Fract Mech 2003;70:4980.
[32] Xue Y, McDowell D, Horstemeyer M, Dale M, Jordan J. Microstructure-based
multistage fatigue modeling of aluminum alloy 7075T651. Eng Fract Mech
2007;74:281023.
[33] Lopez Z, Fatemi A. A method of predicting cyclic stress-strain curve from
tensile properties for steels. Mater Sci Eng A-Struct 2012;556:54050.

Das könnte Ihnen auch gefallen