You are on page 1of 12

International Journal of Fatigue 31 (2009) 526–537

Contents lists available at ScienceDirect

International Journal of Fatigue


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

Experimental evaluation and modeling of sulfur content and anisotropy


of sulfide inclusions on fatigue behavior of steels
Nisha S. Cyril 1, Ali Fatemi *
Mechanical, Industrial and Manufacturing Engineering Department, The University of Toledo, 2801 West Bancroft Street, Toledo, OH 43606, USA

a r t i c l e i n f o a b s t r a c t

Article history: This paper presents and discusses experimental results on the effect of sulfur content and sulfide inclu-
Received 25 January 2008 sions on fatigue behavior of steels with different sulfur and hardness levels under different loading direc-
Received in revised form 31 March 2008 tions. Ductility and toughness of the transverse samples were found to reduce considerably by the
Accepted 1 April 2008
increase in sulfur content, while the differences in the yield and ultimate tensile strengths were not sig-
Available online 11 April 2008
nificant. High sulfur resulted in significant adverse effect of inclusions on fatigue behavior of the trans-
verse samples, particularly in the long life regime. The difference between the high and low sulfur
Keywords:
materials was larger at the higher hardness level. Both sub-surface and surface failure modes were
Anisotropy p
Life prediction
observed at long lives, where the sub-surface failures exhibited much longer life. The area parameter
Inclusions was used to estimate the fatigue limit of the materials used, resulting in reasonable estimations in most
Sulfur effect cases. Roessle–Fatemi equation, which is used to predict the strain-life curve of steels based on hardness,
was modified in order to incorporate the effects of sulfur under transverse loading. It is shown that this
modified equation results in relatively good predictions of fatigue lives.
Ó 2008 Elsevier Ltd. All rights reserved.

1. Introduction Furthermore, sulfides are often present in clusters. Poorly


spaced inclusions favor crack propagation and the coalescing of
Inclusions are major contributors to fatigue and mechanical small cracks originating at inclusions to form larger ones. Clusters
anisotropy in steels. It is not practically possible or commercially of inclusions reduce the toughness and ductility considerably [1].
feasible to always avoid these defects. Metal working processes Kage and Nisitani [2–4] reported significant decrease of ductility
such as rolling and forging result in highly anisotropic microstruc- in transversely oriented samples cut-out from a rolled steel, espe-
ture and/or elongated inclusions. cially at high hardness. Other studies have also shown that for
Fatigue design methods require the knowledge of material transverse samples with high sulfur content, among the tensile
properties to predict fatigue behavior. These properties are often properties the ductility was affected the most [5,6].
p
different in the longitudinal and transverse directions as a result The area parameter model can be used to estimate the fatigue
of anisotropy. A method to predict the transverse behavior from limit of a steel using hardness value, residual stress value and the
available longitudinal material properties or other easily available maximum effective inclusion area obtained from inclusion inspec-
parameters such as hardness would be a useful tool in design. One tion [7]. Inclusions near a free surface have a more deteriorating
of the objectives of this study was to develop such a method. effect on fatigue limit. The probability of extremes can be used to
Sulfur in steels is often in the form of sulfide inclusions like predict the maximum inclusion size for a certain volume of mate-
manganese sulfide (MnS). The presence of sulfur (S) improves the rial. However, this may not always be accurate unless a good num-
machinability of a material [1]. However, sulfide inclusions are ber of measurements are made for maximum inclusion size and
deformable and get elongated in the rolling direction, causing a there are enough points on the probability of extremes graph so
considerable difference in the effective inclusion areas as seen it appears linear [7]. Predicting the maximum inclusion size by
from the longitudinal and transverse directions. This in turn the statistics of extremes may not be very accurate if the inclusion
p
changes the stress concentrations caused in the two directions distribution in the material is not exponential [8]. The area
and affects the properties in each direction differently. parameter model for predicting the fatigue limit may be modified
for elliptical inclusions in cases where length to width ratio of the
inclusion is greater than four [9].
Temmel et al. [5] state that the fatigue limits in the transverse
* Corresponding author. Tel./fax: +1 419 530 8213.
E-mail address: afatemi@eng.utoledo.edu (A. Fatemi). direction were roughly 50% lower than in the longitudinal direc-
1
Present address: Cummins Inc., P.O. Box 3005, Columbus, IN 47201, USA. tion for both the 0.042% S and 0.004% S level 42CrMo4 steels they

0142-1123/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijfatigue.2008.04.001
N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537 527

Nomenclature

%EL percent elongation Su ultimate tensile strength


%RA percent reduction in area Sfl fatigue limit defined at run-out
b fatigue strength exponent S0y cyclic yield strength
c fatigue ductility exponent YS monotonic yield strength
E, E0 monotonic, midlife cycle modulus of elasticity De strain amplitude
HB, HRC, HV Brinell, Rockwell C-Scale, Vicker hardness number Dee, Dep elastic, plastic strain range
K, K0 monotonic, cyclic strength coefficient ef ; e0f true fracture ductility, fatigue ductility coefficient
n, n0 monotonic, cyclic strain hardening exponent Dr stress amplitude
2Nf reversals to failure r; rf ; r0f true stress, true fracture stress, fatigue strength coeffi-
R minimum to maximum stress ratio cient
S sulfur content (%) rw estimated fatigue limit based on inclusion size

investigated. Increasing the sulfur content by a factor of 10 of the fatigue limit and HCF fatigue also increase as the hardness
resulted in a decrease of the fatigue limit in the transverse direc- of material increases due to an increase in notch sensitivity, mak-
tion by up to 38%. In another study by Furuya et al. [10], at 108 cy- ing the higher hardness material more susceptible to inclusion ef-
cles the fatigue strength in the transverse direction of the spring fects [3]. In both bending and torsional fatigue, circular plate-
steels they investigated was about half of that in the rolling direc- shaped inclusions have a stronger negative effect on fatigue life,
tion and the fracture origins were always the large and slender than bar-shaped inclusions [15].
MnS inclusions. The two types of deformable inclusions (oxides This study investigated the effect of sulfur content and sulfide
and sulfides) exhibited different effects on fatigue strength. The inclusions on tensile properties and fatigue behavior of forged steel
deformable oxide-type inclusions never caused fish-eye fracture, under transverse loading conditions. Three different sulfur levels
while the sulfide inclusions did. were tested (high: 0.077% S; low: 0.012% S and ultra low: 0.004%
Decohesion of spherical carbide from the matrix around a non- S) each at two hardness levels. In this paper, the variation of fatigue
metallic inclusion [11] and the trapping of hydrogen by a non- properties and behavior with sulfur content are discussed and a
metallic inclusion cause structural changes around the inclusion, model is presented to predict strain-life curves of steels under
helping in the initiation of a fatigue crack [12]. It has also been sta- transverse loading conditions as a function of sulfur content and
p
ted that a material with low inclusion content would form a typical Brinell hardness. The area parameter was also used to estimate
striation pattern, while a material with high inclusion content the fatigue limit of the materials used, which were then compared
would produce a different surface. The debonding of the inclusion to the experimental values. Tensile and CVN impact properties and
would cause superimposing stress fields. This introduces second- their variation with sulfur content for the materials investigated
ary fatigue cracks which ultimately grow together with the pri- are presented and discussed in Cyril et al. [16].
mary fatigue crack [5]. The fracture surfaces of transverse
samples with fairly high inclusion content would be rough due
to several cracks originating from inclusions [5,13]. 2. Experimental program
For both push–pull and bending fatigue of a low carbon rolled
steel plate, Kage and Nisitani [14] reported that there is significant 2.1. Materials
anisotropy between the fatigue strength in the rolling and thick-
ness directions. This was attributed to the fatigue crack initiating The material used in this study was SAE 4140 steel and was
from slip bands and grain boundaries in the rolling direction spec- treated to three different sulfur levels (high: 0.077% S; low:
imens, while cracks initiated from inclusions in the thickness 0.012% S and ultra low: 0.004% S). The carbon content ranged be-
direction specimens. In both cases, the anisotropy is significant in tween 0.404% and 0.410%. The material was continuous cast as
both LCF and HCF [2,13,14]. 150 mm square cross-section. Pieces were reheated to 1180 °C,
Kage and Nisitani [2] reported that the bending fatigue limit of a and forged into 65 mm square cross-section bars. The reduction ra-
rolled low carbon steel plate reduced by 10.5% in the transverse tio was 5.76 to 1. After forging, the material was normalized at
direction. The anisotropy of fatigue limit was observed only in 900 °C and square cross-sectioned 16 mm  16 mm  65 mm
cases where the fracture originated at inclusions. The anisotropy transverse fatigue sample blanks were cut from 1/4th thickness

Fig. 1. Sample arrangement for testing of transverse specimens in 4140 steels (a), and configuration and dimensions of specimen used for both fatigue and monotonic tension
testing (all dimensions are in mm) (b).
528 N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537

position, as shown in Fig. 1a. The specimens were then machined


from the sample blanks. Samples for longitudinal fatigue tests at
40 HRC, results of which were also available and used for compar-
ison with transverse fatigue test results, had been obtained from
30 mm diameter hot rolled bars with a reduction ratio of 33 to 1.
Photomicrographs of inclusions at the three sulfur levels in the
plane perpendicular to the applied load for the transverse speci-
mens can be seen in Fig. 2 along with the inclusion rating numbers
obtained from these figures. The inclusions were predominantly
sulfides, although some small amounts of globular inclusions also
existed in all materials. From this figure, the maximum inclusion
size appears to be the same for the low sulfur and the ultra low sul-
fur materials, although the ultra low sulfur material has a sparser
distribution of sulfides.
Inclusions were also characterized based on area fraction [17]
and size distribution [18]. Fig. 3 shows the relation between the
area fraction of the inclusions in the transverse and longitudinal
directions and the sulfur content. For the transverse samples, the
percent of sulfide area fractions were very close to the percent sul-
fur weight, while the percent of MnS area fractions were close to
zero for both low and high sulfur levels in the longitudinal
direction.
Fig. 4 shows inclusion characterization by standard DIN K.
Fig. 4a indicates that each material has a higher number of the
smaller sized inclusions and the number of inclusions present de-
creases as the size increases. Fig. 4b indicates that the high sulfur
transverse material has a much higher number of MnS inclusions
of each size than the other materials. The ultra low sulfur material
did not have any of the largest sized (i.e. 1600 sq. lm) MnS
inclusions.
By correlating the fatigue properties of various steels with hard-
ness, Roessle and Fatemi [19] showed that the strain-life curve of a
steel can be approximated fairly accurately based on just the Bri-
nell hardness. The literature also shows that the effect of inclusions
on rolled/forged single phase steels depends on the maximum
inclusion size, inclusion density and distribution and the hardness
of the steel rather than on the steel chemistry or other character-
istics. Therefore, the results obtained in this study are expected
to extend to other steels.

2.2. Specimen preparation, heat-treatment, and test procedure

The cylindrical specimens were 65 mm long with 16 mm long


grip sections and 10 mm long test section, as shown in Fig. 1b.
The reduced diameter of the test section was 5 mm, while the grip
section was 12.6 mm in diameter.
Machined specimens were reheated to 900 °C, quenched, and
tempered to achieve the 43 HRC and 52 HRC hardness levels. Sam-
ples for longitudinal fatigue tests had 40 HRC hardness. The SAE
4140 quenched and tempered steel investigated had a martensitic
microstructure in all cases. After heat-treatment, the specimens Fig. 2. Photomicrographs of inclusions present in transverse samples with (a) high
sulfur, rated as A thin = 3.5 and A heavy = 1.0, D thin = 1.0 and D heavy = 0.5 (no B or
were ground and then polished to a 3 lm surface finish prior to
C), (b) low sulfur, rated as A thin = 0.5 and A heavy = 0, D thin = 1.0 and D heav-
testing. y = 1.0 (no B or C), and (c) ultra low sulfur, rated as A thin = 0.5 and A heavy = 0, D
Closed-loop servo-hydraulic test frames controlled by digital thin = 1.0 and D heavy = 0.5 (no B or C). Loading direction is perpendicular to the
servo-controllers were used to perform the tests. An extensometer page.
rated as ASTM Class B1 was used to control strain. An alignment
fixture was used to align the load train. to 25 Hz. A triangular waveform was used for all tests. All tests
Constant amplitude fatigue tests were conducted according to were fully-reversed (i.e. R = 1, where R is the ratio of minimum
ASTM Standard E606 [20]. At least 17 specimens at 6 or 7 different to maximum strain or stress).
strain amplitudes ranging from 0.225% to 2% were utilized for each
material and hardness. The tests were conducted in strain control 3. Experimental results and comparisons
with the exception of some long life tests and run-out tests where
the control mode was switched from strain to load after the load 3.1. Tensile properties and sulfide inclusions effects
had stabilized in strain control. In strain control test frequencies
used varied from 0.2 Hz to 2 Hz depending on the strain amplitude, The monotonic tensile properties of the different materials
while for run-out tests in load control the frequency was increased are listed in Table 1. The ductility and toughness were reduced
N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537 529

0.08
0.07

sulfide inclusions, %
SAE 4140 steel
0.06

Area fraction of
0.05 Transverse
0.04 Longitudinal
0.03
0.02
0.01
0
0 0.02 0.04 0.06 0.08 0.1
Sulfur content, %

Fig. 3. Area fraction of sulfide inclusions as a function of sulfur content.

80
Hi S Trans
% of total sulfide inclusions in that material

70
sulfide inclusions of a certain size as

Lo S Trans
60 U Lo S Trans
Hi S Long
50 Lo S Long
40

30

20

10

0
100 sq. 200 sq. 400 sq. 800 sq. 1600 sq.
microns microns microns microns microns
Size

450
certain size present in an area of 1000 mm2

Hi S Trans
400
number of sulfide inclusions of a

Lo S Trans
350 U Lo S Trans

300 Hi S Long
Lo S Long
250

200

150

100

50

0
100 sq. 200 sq. 400 sq. 800 sq. 1600 sq.
microns microns microns microns microns
Size

Fig. 4. Sulfide inclusion characterization with size and population.

considerably by the increase in sulfur content for the transverse nificant. Because of the low ductility of the high sulfur mate-
samples. The high sulfur transverse samples failed shortly rial causing it to fail before reaching the ultimate tensile
after yielding with no significant reduction of area or elongation strength, this material exhibited slightly lower ultimate tensile
of gage length before fracture. However, the differences in strength. More details of the tensile and CVN impact toughness
the yield strengths and ultimate tensile strengths for the different properties as a function of sulfur level can be found in Cyril
sulfur level materials at a given hardness level were not sig- et al. [16].
530 N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537

Table 1
Summary of monotonic tensile and cyclic properties for SAE 4140 steels

Sulfur level (%) 0.077 0.012 0.004 0.077 0.012 0.004 0.077 0.015
Direction Trans Trans Trans Trans Trans Trans Long Long
Hardness (HRC) 53 53 51 45 44 42 40 40
AISI test iteration 77 76 98 81 80 99 97 96
Monotonic tensile properties
YS (MPa) 1602 1669 1622 1263 1291 1261 1167 1158
Su (MPa) 1678 1924 1818 1299 1380 1333 1240 1248
EL (%) 3 9 19 1.5 23 31 25 25
RA (%) 1.3 16 43 1.3 24 44 47 48
K (MPa) 2600 2626 2466 1534 1629 1557 1460 1499
n 0.077 0.071 0.065 0.030 0.038 0.035 0.038 0.044
rf (MPa) 1702 1859 2017 1170 1385 1518 1537 1311
ef (%) 1.3 16 56 1.3 27 58 64 65
Cyclic properties
r0f (MPa) 3149 2785 2887 1812 1838 1708 1637 1601
b 0.160 0.107 0.104 0.099 0.083 0.075 0.072 0.067
e0f 0.119 0.221 7.258 0.061 0.553 0.625 2.221 1.267
c 1.011 0.712 1.078 0.744 0.677 0.645 0.794 0.721
S0y (MPa) 1520 1402 1314 1119 927 874 851 856
K0 (MPa) 3064 2722 2232 1753 1872 1664 1617 1696
n0 0.113 0.071 0.085 0.072 0.113 0.104 0.103 0.110
Sfl (MPa) 452 575 620 465 563 569 623 621

1800

1600
True Stress Amplitude, Δσ/2 (MPa)

1400

1200

1000

800

600
77 Trans Hi S 53 HRC
400
76 Trans Lo S 53 HRC
200
98 Trans U Lo S 51 HRC
0
0.0% 0.5% 1.0% 1.5% 2.0% 2.5%
True Strain Amplitude, Δε/2 (%)

1400

1200
True Stress Amplitude, Δσ/2 (MPa)

1000

800

600
81 Trans Hi S 45 HRC
400 80 Trans Lo S 44 HRC
99 Trans U Lo S 42 HRC
200
96 Long Lo S 40 HRC
97 Long Hi S 40 HRC
0
0.0% 0.5% 1.0% 1.5% 2.0% 2.5%
True Strain Amplitude, Δε/2 (%)

Fig. 5. Cyclic stress–strain curves with super-imposed data for SAE 4140 steel at 52 HRC (a), and 43 HRC (b).
N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537 531

3.2. Fatigue behavior and sulfide inclusion effects Fig. 6b, the two curves for the ultra low and low sulfur 43 HRC
transverse specimens are very close to the corresponding curves
Stress amplitude variation with applied cycles indicated a pro- for the longitudinal specimens. The increased hardness from 43
gressive cyclic softening trend for all materials in tests which in- HRC to 52 HRC did not result in improved life for the high sulfur
volved plastic deformation. Steady state hysteresis loops were materials in the transverse direction in the long life regime.
used to determine the cyclic properties for each material at each All three sulfur level materials at 52 HRC exhibited two differ-
hardness level, as listed in Table 1. ent failure modes at some of the lower stress amplitude levels in
The cyclic stress–strain curves are presented in Fig. 5. The cyclic the fully elastic region (at or lower than 0.385% strain amplitude),
curve of the high sulfur material is higher than the curves for the with some specimens demonstrating sub-surface failure and oth-
low sulfur materials at both hardness levels, indicating lower plas- ers, surface failure. There was considerable scatter of fatigue life
tic deformation at a given strain level, and consequently less cyclic at these stress amplitude levels due to the different failure modes.
softening. The two sulfur level materials in the longitudinal direc- The sub-surface failures exhibited much longer life. Surface com-
tion show similar curves. The ultra low sulfur transverse material pressive residual stresses from grinding can persist at the lower
had similar cyclic stress–strain curves to the longitudinal cyclic stress amplitudes where there is no cyclic plastic deformation
stress–strain curve, as seen from Fig. 5b. (i.e. HCF regime) and play a role in shifting the crack nucleation site
The stress–life (S–N) curves are presented in Fig. 6. The SS nota- from surface to sub-surface. This is discussed further in Cyril et al.
tion in this and other similar figures indicates sub-surface failure [16].
where failure originated at an internal inclusion. Relatively small For the same life, cyclic plastic strains of the high sulfur trans-
difference is observed between the low and the ultra low sulfur le- verse material were over an order of magnitude lower when com-
vel materials at both hardness levels. The high sulfur transverse pared to the other two lower sulfur level materials. This would be
material, however, is significantly inferior to the other materials, expected, as in low cycle fatigue the ability of a material to plasti-
particularly at long lives, indicating a much more adverse effect cally deform has a strong influence on its fatigue life. For the 52
of inclusions in the longer life regime. As can be seen from HRC transverse samples, in the short life regime the ultra low

Fig. 6. True stress amplitude versus reversals to failure for 52 HRC materials (a), and 43 HRC materials (b), where SS indicates sub-surface failure.
532 N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537

material showed significantly higher cyclic plasticity compared to At 43 HRC, there is about a factor of forty difference in life at
the low sulfur material, unlike for the 43 HRC transverse speci- high strain amplitudes (strain levels involving plastic deformation)
mens, where not much difference was noticed between the low and about one order of magnitude difference at low strain ampli-
and ultra low sulfur materials. The higher cyclic plasticity at 52 tudes (fully elastic regime) between the high sulfur and the ultra
HRC is a consequence of improvement in ductility and toughness low sulfur materials (see Fig. 7b). The difference between the high
resulting from the very low sulfur content. For the longitudinally and low S materials is, however, larger at the higher hardness level
loaded samples at 40 HRC, both high and low sulfur level materials of 52 HRC. At 43 HRC the strain-life curves for the ultra low S and
showed the same plastic strain amplitude for a given life, and sim- low S materials in the transverse direction are close to each other
ilar to the ultra low sulfur transverse material at that life. and to the curves for the longitudinally loaded samples.
Fig. 7 displays the strain-life curves, and superimposed fatigue
data. As can be seen from Fig. 7a for 52 HRC, the ultra low S mate- 3.3. Fractography
rial shows considerable improvement over the low S material at
very short life, which is a consequence of relatively high improve- Most of the failures in all materials were surface initiated. A few
ment in ductility between the two materials. In the long life re- sub-surface failures originating from large inclusions were ob-
gime, the curves of the low S and ultra low S materials are close served in the long life regime for the higher hardness material
to each other. At 52 HRC, there was about a factor of 8 difference loaded in the transverse direction. This is further discussed in Cyril
in fatigue life in the low cycle fatigue regime and over an order et al. [16]. A typical sub-surface failure initiated at an inclusion can
of magnitude difference in the long life regime between the high be observed in Fig. 8a. In the short life region, often no inclusion
sulfur and the low sulfur transverse materials. This difference was found at the initiation site for the ultra low and low sulfur
was about a factor of 30 in the low cycle regime and about two or- material, while several elongated MnS inclusions were found at
ders of magnitude in the high cycle regime between the high and and near the failure origin for the high sulfur materials.
the ultra low S materials.

10.00%
98 Trans U Lo S 51 HRC
76 Trans Lo S 53 HRC
77 Trans Hi S 53 HRC
True Strain Amplitude, Δε/2 (%)

1.00%

(SS)
(SS) (SS)

(SS)

0.10%
1E+1 1E+2 1E+3 1E+4 1E+5 1E+6 1E+7 1E+8
Reversals to Failure, 2Nf

10.00%
99 Trans U Lo S 42 HRC
80 Trans Lo S 44 HRC
True Strain Amplitude, Δε/2 (%)

81 Trans Hi S 45 HRC
96 Long Lo S 40 HRC
97 Long Hi S 40 HRC

1.00%

(SS ) (SS)

0.10%
1E+1 1E+2 1E+3 1E+4 1E+5 1E+6 1E+7 1E+8
Reversals to Failure, 2Nf

Fig. 7. Strain-life curves for the 52 HRC materials (a), and 43 HRC materials (b), where SS indicates sub-surface failure.
N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537 533

Fig. 8. (a) Fish-eye on fracture surface for low S material at 52 HRC at strain amplitude of 0.325%. (b) Elongated inclusions at and around the fracture origin for high S material
at 52 HRC at strain amplitude of 1%. (c) Flat and rough fracture surface for high S material at 43 HRC at strain amplitude of 0.275%. (d) Overview of the fracture surface for
surface initiated failure of low S material at 52 HRC at strain amplitude of 0.325%.

Fig. 8b shows a high magnification view of the failure initiation 3.4. Fatigue limit values and their estimations
site at the surface for a high sulfur 52 HRC specimen, tested at high
strain amplitude of 1%, where several elongated inclusions indi- Fatigue limit (Sfl) values, defined as the stress amplitude at
cated as MnS inclusions by EDS can be observed at and around which no failure occurred in 10 million reversals for at least one
the fracture origin. The fracture surfaces for both high and low of the three duplicate tests conducted are presented in Table 1.
strain amplitude tests of the high S material were very rough. This At 43 HRC, the ultra low S material loaded in the transverse direc-
type of surface would be caused by several cracks originating from tion exhibited almost the same fatigue limit as the low S material
inclusions, propagating and merging, leading to fracture. Fig. 8c (i.e. 569 MPa). This can be explained by the maximum size of the
shows such a fracture surface. inclusions being the same for the two lower S materials, as can
For both the ultra low and low sulfur materials the main crack be observed from Fig. 2, despite a more sparse inclusion distribu-
typically initiated at the surface (no inclusion was found at or tion for the ultra low S material.
near the fracture surface) and then grew across the fracture ori- Both the high sulfur and the low sulfur materials have practi-
gin. Unlike the high sulfur material, the fracture surfaces of these cally the same fatigue limit in the transverse direction at 43 HRC
materials indicated a clear initiation point, a crack growth region, and at 52 HRC (i.e. independent of hardness). This indicates that
and a region of final fracture. The fracture surface for the low S in the long life regime in the transverse direction, the sulfur level
materials was not flat, indicating ductile rupture. Overview of a effect dominates more than the difference in hardness effect. How-
typical fracture surface for the low S material where failure initi- ever, the ultra low sulfur material at 52 HRC has a higher fatigue
ated from surface is shown in Fig. 8d, where crack initiation site limit (620 MPa) than the one at 43 HRC (569 MPa). It appears that
is shown with an arrow and shear lips are observed around the when the sulfur level is reduced to that extent, the beneficial effect
circumference. Energy dispersive spectroscopy (EDS) of the failure of the increased hardness begins to show for the transverse
initiation site did not indicate existence of MnS inclusion at this samples.
location, although the inclusion may have fallen off during or At 52 HRC and in the transverse direction, the fatigue limit was
after failure. The fracture mechanism was microvoid coalescence 27% lower for the high sulfur material when compared to the ultra
in all cases. low sulfur material (452 MPa versus 620 MPa) and at 43 HRC this
534 N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537

Table 2
p
Fatigue limit estimations using the area parameter model
p
HV HRC S Surface Exp. fatigue area Estimate fatigue limit for Ratio of est. to exp. Estimate fatigue limit for Ratio of est. to exp.
(wt.%) residual stress limit, Sfl (lm) interior inclusion (MPa) fatigue limit for int. surface inclusion (MPa) fatigue limit for surf.
(MPa) (MPa) incl. incl.
562 53 0.077 200 452 40 575 1.27 585 1.29
562 53 0.012 200 575 28 610 1.06 616 1.07
530 51 0.004 70 620 20 615 0.99 577 0.93
440 45 0.077 186 465 40 472 1.01 488 1.05
440 44 0.012 186 563 28 501 0.89 513 0.91
413 42 0.004 70 569 20 505 0.89 477 0.84

difference was 18%. The fatigue limit for the transverse samples of into account as a mean stress through the R ratio in this equation,
low and ultra low sulfur at 43 HRC were about 9% lower than the where
fatigue limits in the longitudinal direction at 40 HRC. R ¼ ðrresidual  rw Þ=ðrresidual þ rw Þ ð2Þ
According to Murakami et al. [7], for high strength steels con-
taining inclusions that are larger than the critical size, individual An iterative process is used to find rw from Eqs. (1) and (2) since the
specimens have specific fatigue limits depending on the position value of rw in Eq. (1) is a function of R, while the value of R in Eq. (2)
of inclusions with respect to the surface. In addition, a large scatter is a function of rw.
is observed in fatigue limit of such steels. For such steels, for inte- Estimated fatigue limits based on Eq. (1) for all the materials are
rior inclusions the fatigue limit, rw, may be predicted based on listed in Table 2. The maximum inclusion size from the inclusion
Vickers hardness, HV, and the area of the largest inclusion size characterization in Fig. 4(a) is 1600 sq. lm, 800 sq. lm and
using [7]: 400 sq. lm for the high, low and ultra low sulfur transverse mate-
p
p rials, respectively. This implies area values of 40 lm, 28 lm and
rw ¼ 1:56½ðHV þ 120Þ=ð areaÞ1=6 ½ð1  RÞ=2a ð1Þ 20 lm, respectively. Surface residual stresses resulting from speci-
where R is the stress ratio, and a = 0.226 + HV  104. For inclusions men grinding were measured on some specimens from the differ-
just touching the surface, the multiplier constant in Eq. (1) is 1.41, ent test materials, as are listed in Table 2. As residual stresses from
rather than 1.56, which gives the lower bound for fatigue limit, grinding were only measured on the surface and the sub-surface
rw. For fully-reversed loading and no residual stress, R = 1, and values are expected to be small, the residual stress was not taken
the last term in Eq. (1) is equal to 1. Residual stress effect is taken into account for interior defects. The ratios of estimated fatigue

0
Fatigue Strength Coefficient, σf' (MPa)

3500 Trans 52 HRC


-0.02
Trans 43 HRC
Fatigue Strength Exponent, b

3000 -0.04
Long 40 HRC
2500 -0.06

2000 -0.08
-0.1
1500
Trans 52 HRC -0.12
1000
-0.14
Trans 43 HRC
500 -0.16
Long 40 HRC
0 -0.18
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
Sulfur Content, % Sulfur Content, %

8 0
Trans 52 HRC
Fatigue Ductility Coefficient, εf'

Trans 52 HRC
Fatigue Ductility Exponent, c .

7 -0.2 Trans 43 HRC


Trans 43 HRC
6
Long 40 HRC
Long 40 HRC -0.4
5
4 -0.6

3
-0.8
2
-1
1
0 -1.2
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
Sulfur Content, % Sulfur Content, %

Fig. 9. Variation of fatigue strength coefficient (a), fatigue strength exponent (b), fatigue ductility coefficient (c), and fatigue ductility exponent (d) with sulfur content.
N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537 535

limits to the experimental ones are listed in Table 2 and range be- steeper downward slope in the long life regime of the strain-life
tween 0.89 and 1.27 if the inclusions were considered interior curve. Since a 52 HRC material is very susceptible to stress concen-
inclusions, and between 0.84 and 1.29 if the inclusions were con- trations, especially in the fully elastic region, it is noted that the va-
sidered to be just touching the surface. As can be seen from Table lue of the slope of b at this hardness is large. In the longitudinal
2, except for the case of high S and high hardness material, the esti- direction, since the effective area of the inclusions in the plane per-
mated fatigue limit is within about 15% of the experimental value, pendicular to the applied load is not very different for the 0.012% S
which may be considered as reasonable. and the 0.077% S materials (see Fig. 2), the b value does not vary
much.
4. Variation of fatigue properties with sulfur content and Good correlations with S content were also observed for the fa-
estimation of strain-life curves under transverse loading tigue ductility coefficient (e0f ) and fatigue ductility exponent (c) at
43 HRC, although these correlations were poor at 52 HRC (see
It can be seen from Fig. 9a that r0f does not alter significantly Fig. 9c and d). However, e0f and c values for the ultra low S and high
with sulfur content either in transverse or longitudinal directions. S materials at 52 HRC were not considered to be very accurate. This
As was the case with the other strength related properties (YS, Su, K is because these values for the ultra low S material were obtained
and K0 ), the values of r0f in the longitudinal direction were slightly from a linear fit of data which did not exhibit a linear trend, and for
lower than in the transverse direction. The lower strength of the the high S material the values were obtained from a fit of very lim-
high and low sulfur materials loaded in the longitudinal direction ited plastic strain data due to very low cyclic ductility of this
could be due to the slightly lower hardness level (40 HRC), when material.
compared to the hardness of the transverse direction samples It should be noted that since the area fractions of sulfide inclu-
(43 HRC). sions are closely correlated with the sulfur content for the trans-
Fig. 9b indicates that b shows a decreasing trend with increas- verse direction (see Fig. 3), correlations between the sulfide
ing S at both hardness levels. This implies that the effect of increas- inclusion area fractions and the fatigue properties are the same
ing the sulfur content (and hence the sulfide inclusions) results in a as those between the sulfur content and the fatigue properties.

Fig. 10. Predicted strain-life curves in the transverse direction using Eq. (7) along with experimental transverse strain-life curves at 52 HRC (a), and 43 HRC (b).
536 N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537

1E+8

1E+7

Predicted Reversals to Failure


1E+6

1E+5

77 Trans Hi S 53 HRC
1E+4
76 Trans Lo S 53 HRC
98 Trans U Lo S 51 HRC
1E+3 81 Trans Hi S 45 HRC
80 Trans Hi S 44 HRC
99 Trans Hi S 42 HRC
1E+2
Life factor of 2
Life factor of 5
1E+1
1E+1 1E+2 1E+3 1E+4 1E+5 1E+6 1E+7 1E+8
Reversals to Failure from Experimental Data

Fig. 11. Comparison of predictions by Eq. (7) with actual experimental lives.

An equation to approximate the strain-life curve of a material cial steels have sulfur contents of about 0.05% and it is not common
using Brinell hardness (HB) as the variable was proposed by Roes- for any commercial steel to have such high sulfur content. Note that
sle and Fatemi [19]: when S  0 (i.e., for longitudinal loading direction), values of e0f and c
in Eqs. (5) and (6) revert back to the values in Eq. (3).
De 4:25ðHBÞ þ 225
¼ ð2Nf Þ0:09 The proposed equation is then given by
2 E
0:32ðHBÞ2  487ðHBÞ þ 191000 De 4:25ðHBÞ þ 225
þ ð2N f Þ0:56 ð3Þ ¼ ð2Nf Þ½0:3ðSÞ0:09
E 2 E
0:32ðHBÞ2  487ðHBÞ þ 191000
where E is in MPa. The equation is valid for a hardness range of 150 þ ½1  7:25ðSÞ
HB to 700 HB. The strain-life curve predicted by the above equation E
½1:22ðSÞ0:65
for the material at 40 HRC which is equivalent to 375 HB was com-  ð2N f Þ ð7Þ
pared with the experimental strain-life curves of the available lon- for 150 < HB < 700 and S not greater than 0.14%. Fig. 10 shows the
gitudinal data of the material investigated. In the life region predicted curves based on Eq. (7). For the different sulfur level
between 103 to 106 cycles, the predicted curve agrees fairly well materials and at both hardness levels, most of the predicted curves
with the experimental curve, within a factor of 2 on the conserva- are within a factor of about two in the life range between 102 and
tive side. 106 cycles.
As was mentioned previously, the fatigue strength coefficient Experimental lives versus predicted lives based on Eq. (7) are
(r0f ) did not vary significantly with sulfur content in the transverse presented in Fig. 11. It is noted that 73% of the data fall within a
direction. Therefore, the coefficient of the first term on the right scatter band of 2% and 92% of data fall within a scatter band of 5.
hand side of Eq. (3) was left as is, independent of sulfur level and An important observation from this figure is that much of the scat-
as a function of hardness alone. The exponent b, however, de- ter seen is due to the scatter in experimental fatigue life data from
creases as a function of sulfur level in the transverse direction duplicate tests (i.e. variations in the horizontal direction), rather
(see Fig. 9b). The following equation provides a reasonable repre- than from differences between the predicted and experimental
sentation for variation of b with S level at both hardness levels fatigue lives.
shown in Fig. 9b, for loading in the transverse direction:
b ¼ 0:3ðSÞ  0:09 ð4Þ 5. Conclusions
The value of 0.09 as the constant value in this equation ensures
Based on the observed experimental results and analysis pre-
that the modified equation reverts back to the value in Eq. (3) for
sented, the following conclusions can be made:
the case when S  0 (which is equivalent to the effective inclusion
area when loading is in the same direction as the sulfide inclusions,
1. The inclusions were predominantly sulfides and for the
i.e. longitudinal direction).
transverse samples the percent of sulfide area fractions were
Correlations between the fatigue ductility coefficient (e0f ) and
very close to the percent sulfur weight. The maximum inclu-
exponent (c) with S content for loading in the transverse direction
sion size was about the same for the low sulfur and the ultra
were good at 43 HRC (see Fig. 9c and d) and can be represented
low sulfur materials, although the ultra low sulfur material
reasonably well by
had a sparser distribution of sulfides. For each material the
0:32ðHBÞ2  487ðHBÞ þ 191000 number of inclusions decreased as the inclusion size
e0f ¼ ½1  7:25ðSÞ ð5Þ increased.
E
c ¼ 1:22ðSÞ  0:56 ð6Þ 2. The ductility and toughness were reduced considerably by
the increase in sulfur content for the transverse samples.
As stated previously, however, at 52 HRC these correlations were However, the differences in the yield strengths and ultimate
poor for the reasons discussed earlier. Also, Eq. (5) is not valid for tensile strengths for the different sulfur level materials at a
S > 0.14% as that would make e0f negative. However, many commer- given hardness level were not significant.
N.S. Cyril, A. Fatemi / International Journal of Fatigue 31 (2009) 526–537 537

3. The stress–life (S–N) curves showed small difference strength and ductility exponents (b and c) are functions of
between the low and the ultra low sulfur level materials at sulfur level. Comparisons between the predicted and exper-
both hardness levels. The high sulfur transverse material imental results indicate satisfactory predictions for most of
was significantly inferior to the other materials, particularly the data from the three sulfur level materials and at both
at long lives, indicating a much more adverse effect of inclu- hardness levels investigated.
sions. S–N curves for the ultra low and low sulfur transverse
specimens were close to the corresponding curves for the
longitudinal specimens. Acknowledgements
4. For the same life, cyclic plastic strains of the high sulfur
transverse material were over an order of magnitude lower Financial support for this project was provided by the Ameri-
when compared to the other two lower sulfur level materi- can Iron and Steel Institute (AISI). Technical assistance of the AI-
als, due to reduced ductility. For longitudinally loaded sam- SI’s Bar Fatigue Committee Chairman, Dr. Tom Oakwood, is
ples, the high and low sulfur level materials showed the appreciated. Mr. Peter Bauerle of Chrysler conducted the heat
same plastic strain amplitude for a given life, and similar treatments, Mr. Steve Ferdon of Cummins Engine provided resid-
to the ultra low sulfur transverse material at that life. ual stress measurements, and Mr. Robert Cryderman of
5. Strain-life curve at 52 HRC for the ultra low S material MACSTEEL provided the inclusion ratings for the materials
showed considerable improvement over the low S material investigated.
at very short life, as a consequence of better ductility, while
in the long life regime the two curves were close to each References
other. At 52 HRC, there was about a factor of 8 difference
in fatigue life in LCF regime and over an order of magnitude [1] Kiessling R. Nonmetallic inclusions and their effects on the properties of
difference in HCF regime between the high sulfur and the ferrous alloys. In: Ilshner Bernhard, editor. Encyclopedia of material science
and technology. Oxford, UK: Elsevier Science Ltd.; 2001. p. 6278–83.
low sulfur transverse materials. This difference was about [2] Kage M, Nisitani H. Some consideration on the anisotropy of fatigue limit in a
a factor of 30 in LCF regime and about two orders of magni- rolled steel plate, based on the observation of fatigue process in
tude in HCF regime between the high and the ultra low S electropolished specimens. Bull JSME 1972;15(83):565–74.
[3] Nisitani H, Kage M. Effect of heat treatment on anisotropy of fatigue strength of
materials. a rolled steel. Bull JSME 1980;23(180):799–806.
6. At 43 HRC, there was about a factor of 40 difference in life in [4] Kage M, Nisitani H. Fracture surface of anisotropic rolled steel in fatigue. J Soc
LCF and about one order of magnitude difference in HCF Mater Sci Japan/Zairyo 1980;29(321):574–9.
[5] Temmel C, Karlsson B, Ingesten N. Fatigue anisotropy in cross-rolled, hardened
between the high sulfur and the ultra low sulfur materials.
medium carbon steel resulting from MnS inclusions. Metall Mater Trans A
At this hardness the strain-life curves for the ultra low and 2006;37a:2995–3007.
low sulfur materials in the transverse direction are close to [6] Wilson AD. Comparing the effect of inclusions on ductility, toughness, and
fatigue properties. In: Glodowski RJ, editor. ASTM STP 794, vol.
each other and to the curves for the longitudinally loaded
75. Philadelphia, PA: American Society for Testing and Materials; 1983. p.
samples. 130–46.
7. The increased hardness from 43 HRC to 52 HRC did not [7] Murakami Y, Kodama S, Konuma S. Quantitative evaluation of effects of non-
result in improved HCF behavior of the high sulfur material metallic inclusions on fatigue strength of high strength steels – parts I and II.
Trans JSME 1988;54:291–8. p. 299–307.
in the transverse direction. This is due to a more pronounced [8] Juvonen P, Hanninen H. Fatigue behavior of calcium-treated and noncalcium
effect of inclusions at higher hardness and at longer life. treated AISI 8620 carburizing steel. In: Advances in mechanical behavior,
8. The three sulfur level materials at 52 HRC exhibited sub-sur- plasticity and damage. Euromat 2000, vol. 2. UK: Elsevier; 2000. p. 1059–63.
[9] Makino T. The effect of forging ratio and the metal flow direction on very high-
face as well as surface failure modes at long lives, resulting cycle fatigue properties of steel bars. In: McDowell DL, Newman Jr JC, Saxena A,
in considerable scatter of fatigue life. The fracture surfaces editors. Proceedings of the ninth international fatigue congress, Paper
of the high sulfur transverse material were very rough, Reference No. FT581 (CD ROM), Atlanta, GA. Oxford: Elsevier; 2006.
[10] Furuya Y, Matsuoka S, Abe T. Inclusion controlled fatigue properties of
caused by several cracks originating from inclusions, propa- 1800 MPa steels. Metall Mater Trans A 2004;35a(12):3737–44.
gating and merging. For the ultra low and low sulfur materi- [11] Shiozawa K, Morii Y, Nishino S. Subsurface crack initiation and propagation
als fracture surfaces were not flat, and indicated ductile mechanism in the super-long fatigue regime for high speed tool steel, JIS
SKH51, by fracture surface topographic analysis. Nippon Kikai Gakkai
rupture.
Ronbunshu, A Hen/Trans Jpn Soc Mech Eng A 2004;70(3):495–503.
9. At 43 HRC, the ultra low and the low sulfur materials loaded [12] Murakami Y, Yokoyama N, Takai K. The effect of hydrogen trapped by
in the transverse direction exhibited similar fatigue limits, inclusions on ultra-long life fatigue failure of bearing steel. Zairyo/J Soc Mater
Sci 2001;50(10):1068–73.
since the maximum size of inclusions was the same for the
[13] Kage M, Nisitani H. Anisotropy of fatigue limit and of the crack propagation
two lower sulfur materials, despite a more sparse inclusion property of a rolled steel plate. Bull JSME 1975;18(126):1385–94.
distribution for the ultra low sulfur material. Also, the high [14] Kage M, Nisitani H. Anisotropy of low cycle torsional and push-pull fatigue in a
and the low sulfur materials had similar fatigue limits in rolled steel. Bull JSME 1977;20(149):1353–8.
[15] Kage M, Nisitani H. Effect of inclusions on the torsional fatigue of anisotropic
the transverse direction at 43 HRC and at 52 HRC (i.e. inde- rolled steel. Bull JSME 1978;21(156):948–54.
pendent of hardness), indicating the sulfur level effect dom- [16] Cyril N, Fatemi A, Cryderman B. Effects of sulfur level and anisotropy of sulfide
inates more than the hardness effect. The ultra low sulfur inclusions on tensile, impact, and fatigue properties of SAE 4140 steel. SAE
Technical Paper 2008-01-0434, Society of Automotive Engineers, PA, USA;
material at 52 HRC, however, had a higher fatigue limit than 2008.
at 43 HRC. [17] IS G0555. Microscopic testing method for the non-metallic inclusions in steel.
p JIS handbook: ferrous metals and metallurgy I. Japanese Standards
10. Murakami’s area method based on hardness and the size of
Association; 2002.
inclusions resulted in reasonable estimates of the fatigue [18] DIN K 50602. Microscopic examination of special steels using standard
limit for most cases, where the estimated fatigue limits were diagrams to assess the content of non-metallic inclusions. Deutsche Norm;
within about 15% of the experimental values. 1985.
[19] Roessle ML, Fatemi A. Strain-controlled fatigue properties of steels and some
11. A proposed model to predict strain-life curves under trans-
simple approximations. Int J Fatigue 2000;22:495–511.
verse loading is presented based on Roessle–Fatemi strain- [20] ASTM Standard E606-92. Standard practice for strain-controlled fatigue
life estimation model. The original Roessle–Fatemi equation testing. Annual book of ASTM standards, vol. 03.01. West Conshohocken,
PA: American Society for Testing and Materials; 2004. p. 593–606.
is in terms of material hardness, whereas in the proposed
modified equation for transverse loading the fatigue

You might also like