You are on page 1of 13

International Journal of Fatigue 73 (2015) 119–131

Contents lists available at ScienceDirect

International Journal of Fatigue


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

Ratcheting-based microstructure-sensitive modeling of the cyclic


hardening response of case-hardened bearing steels subject to Rolling
Contact Fatigue
Anup S. Pandkar, Nagaraj Arakere ⇑, Ghatu Subhash
Mechanical & Aerospace Engineering, University of Florida, Gainesville, FL 32611, USA

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

Article history: Rolling Contact Fatigue (RCF) is a commonly observed phenomenon in bearings that imposes a highly
Received 17 August 2014 localized subsurface volume to a complex triaxial and non-proportional fatigue loading. Case-hardened
Received in revised form 28 November 2014 M50-NiL steel, a widely used aerospace bearing material, when subjected to RCF loading shows a definite
Accepted 3 December 2014
increase in hardness within the localized RCF-affected zone as a function of cycles. To simulate such cyclic
Available online 15 December 2014
hardening material response during RCF using finite element (FE) modeling, knowledge of localized cyclic
hardening properties of bearing steels is required. However, these properties are not available for
Keywords:
through-hardened and case hardened steels nor are they readily measurable from the conventional bulk
Ratcheting
Rolling Contact Fatigue
specimen testing methods used in high cycle fatigue. To overcome this challenge, an alternative method
Cyclic hardening properties to evaluate the material-specific localized cyclic hardening properties as a function of case-layer depth is
Micro-plasticity presented. This method uses the experimentally measured increase in micro-hardness within the
Carbide microstructure RCF-affected zone, in conjunction with elastic–plastic FE simulation of RCF cyclic hardening, to extract
localized cyclic constitutive properties, and investigate the underlying physics associated with the local
hardness increase. Case hardened bearing steels are complex composites having a heterogeneous graded
microstructure with carbide precipitates ranging from nanometer to micrometer length scales that are
largely responsible for the material hardening. Micro-plastic strain accumulation via ratcheting at the
scale of the carbide microstructure is proposed to be the underlying mechanism for cyclic hardening. This
manuscript presents for the first time a procedure to estimate microstructure-sensitive isotropic and
kinematic cyclic hardening parameters, from a judicious combination of numerical and experimental
methods. Methods presented are applicable to both through-hardened and case hardened bearing steels.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction 2.8–3.0 GPa [1,2]. The high strength properties of these steels are
largely derived from the heterogeneous carbide microstructure
Ball and roller element bearings used in aircraft engines endure with carbide sizes ranging from nanometer to micrometer length
hundreds of millions of Rolling Contact Fatigue (RCF) cycles. When scales. Extending the bearing life is a challenging task and requires
in service, the useful life span of a bearing is determined by an development of sophisticated numerical tools which can model
onset of a fatigue spall formation. Bearing and aircraft manufactur- material specific realistic carbide microstructures and predict the
ers have considerable interest in extending the useful RCF life of progression of bearing failure over a very large number of cycles.
bearings. It is not uncommon for aerospace bearings to be designed This calls for an in-depth understanding of the response of bearing
with an expected L10 life (number of cycles survived by 90% of the steels to RCF, including the effects of carbide microstructure that
bearing population under specified load and speed conditions) of are fundamental to accurate model the damage progression over
25,000 h or 1010–1011 RCF cycles. To achieve this, modern aero- billions of RCF cycles.
space bearings use high strength ultra-clean vacuum induction Failure of bearings is not an instantaneous process, but a result
melt vacuum arc remelt (VIMVAR) bearing steels with complex of gradual damage accumulation over billions of fatigue cycles
microstructures. Yield strength of such bearing steels can approach (1010–1011 cycles). Such damage initiates in a small subsurface
volume of contacting bodies which is subjected to triaxial and
⇑ Corresponding author. Tel.: +1 352 392 0856; fax: +1 352 392 1071. non-proportional cyclic state of stress [3–5]. Extremely localized
E-mail address: nagaraj@ufl.edu (N. Arakere). RCF-affected volume and complexity of the induced stress are

http://dx.doi.org/10.1016/j.ijfatigue.2014.12.002
0142-1123/Ó 2014 Elsevier Ltd. All rights reserved.
120 A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131

the key factors which make understanding of RCF failures challeng- in advanced aerospace bearing applications [20]. But even after
ing compared to the conventional high cycle fatigue. Many decades of research, practically no data exists for the cyclic harden-
researchers have shown that complex microstructural changes ing properties of even the through-hardened steels and none for
take place in the subsurface region during RCF [6–11]. Knowledge the case-hardened steels. Very limited information is available
of such subsurface changes can give crucial insights into the dam- for basic monotonic stress–strain response of high strength bearing
age accumulation process and prove helpful to better understand steels such as M50, M50-NiL and Pyrowear (P675) steel [1,21,22].
the RCF induced failures. In an attempt to understand and quantify Limited cyclic data from bulk specimens is available for AISI
the evolution of such subsurface changes, Bhattacharyya et al. [2] 52100 and AISI 1070 steels [23]. Such unavailability of the relevant
studied the cyclic micro-plastic deformation characteristics of cyclic material properties prevents modeling community from
M50-NiL bearing steel subjected to increasing number of RCF undertaking realistic studies on RCF induced damage evolution in
cycles in a ball-on-rod RCF tester. They performed micro-hardness bearings. Traditionally, the calibration of cyclic material parame-
measurements within the RCF affected region of the test specimen, ters requires data from the cyclic strain controlled experiments
which revealed a definite increase in micro-hardness compared to [16,24]. Through-hardened bearing steels have uniform material
the virgin (RCF unaffected) hardness of material. Further, it was properties throughout the volume; therefore, it is possible to pre-
reported that magnitude of this local hardness increase (DH) is pare specimens for conducting such cyclic tests. Case-hardened
proportional to the number of RCF cycles. Such localized mapping bearing steels (e.g. M50-NiL and Pyrowear 675), on the other hand,
of hardness evolution as a function of RCF cycles can be used as a have material properties which vary as a function of depth. As a
marker for the material damage evolution during RCF. result, extraction of specimens with homogeneous microstructure
The objective of this manuscript is to numerically investigate over the specimen dimension is not possible for the traditional fati-
the micromechanical behavior of bearing steels with carbide gue experiments. In addition, the depth of case-hardening layer in
microstructures subject to RCF, to understand the mechanism bearing steels is usually small (2.5 mm) [25,26]; therefore,
responsible for this experimentally observed local hardness extraction of multiple miniature specimens for cyclic testing
increase (DH). An elastic–plastic finite element (FE) analysis is becomes a challenging task. Currently, no experimental technique
used to simulate this cyclic hardening behavior. Simulation of such exists in the literature to conduct mechanistically sound cyclic tests
RCF phenomenon involves multiple loading cycles; therefore, it on such case-hardened steels with spatially varying material proper-
will also provide insights into the cyclically evolving stress–strain ties. Allison et al. [27] have recently devised a methodology to
fields within the RCF affected region. Such information about the perform uniaxial compression tests on the miniature specimens
stress–strain fields is useful for quantifying the local fatigue dam- (1–2 mm in size) extracted from inside of the RCF affected
age accumulation [12–14]. Maintaining an elastic material regions of through-hardened M50 steel to determine its local
response after shakedown for a longer period of time is critical monotonic constitutive response. But this method is limited to
for extending the bearing life [15]. For a given applied stress level the testing of homogeneous materials under monotonic loading con-
and operating temperature, the carbide morphology has been ditions, and cannot be readily extended to case-hardened materials
observed to alter the response of bearing steel to RCF; therefore, under cyclic loading. As a result, knowledge of cyclic hardening
it is suspected to influence the shakedown process as well. The parameters for case-hardened bearing steels continues to be an
microstructure-sensitive RCF model presented here will investi- unsolved problem. To overcome this challenge, we have developed
gate the role of carbide microstructure in achieving an elastic an integrated experimental and modeling method to estimate the
response after shakedown which may contribute toward better cyclic hardening parameters of case-hardened bearing steels. This
design of bearing steels. method uses the results obtained from micro-hardness measure-
Realistic FE simulation of material response to cyclic loading ments within the RCF affected region of M50-NiL bearing steel
requires knowledge of cyclic plasticity and availability of reliable reported by Bhattacharyya et al. [2].
material models. Over the past few decades, development of The manuscript is structured as follows. Section 2 briefly
accurate material models has received substantial attention from discusses ball-on-rod RCF test along with the experimental
researchers and as a result several constitutive material models results obtained from the local micro-hardness measurements
are available to model cyclic plasticity phenomenon including [2]. Based on these results, an alternative method to estimate
RCF. With availability of numerous material models, the choice the depth-dependent cyclic hardening parameters of case-
of appropriate model primarily depends on two factors; (i) capa- hardened M50-NiL steel is presented in Section 3. The cyclic
bility of model to capture the required material response and (ii) hardening behavior observed in experiments is thought to be a
ability and ease to calibrate the unknown material parameters result of micro-plastic strain accumulation via ratcheting, which
from experiments. With these considerations, we have chosen is commonly observed in RCF applications [5,17,28]. Therefore,
the Non-linear Isotropic and Kinematic Hardening (NIKH) model in Section 4, ratcheting is first explained as a cyclic hardening
developed by Chaboche and Lamaitre [16] to study the cyclic mechanism (rather than a mechanism of strain accumulation) from
hardening response of bearing steels under RCF. The advantage a one-dimensional perspective. This one dimensional analogy is
of this model lies in the fact that it has only four unknown mate- then extended to a multiaxial ball-on-rod RCF test in Section 5.
rial parameters. Nonetheless, this versatile model is capable of This section also describes the methodology adopted to simulate
simulating complex material responses such as Bauschinger the experimentally observed hardness increase (DH). Section 6
effect, cyclic hardening with plastic shakedown, ratcheting, and discusses the results of these FE simulations followed by conclu-
relaxation of mean stress, all of which are known to be prevalent sions in Section 7.
during RCF. As a result, it is popularly used by researchers for
studying the response of various materials under different spec-
trums of cyclic loading [5,12,13,17–19]. The four material param- 2. RCF Testing and experimental results
eters associated with this model govern the response of material
under cyclic loading; therefore their availability is essential for M50-NiL is one of the most widely used case-hardened bearing
accurate modeling of any cyclic hardening phenomenon including steel in the aerospace industry. The ball-on-rod RCF tester is com-
RCF. monly employed to test bearing steels under RCF loading [2,15].
Through-hardened steels (M50 and AISI 52100) and case- During this test, cylindrical M50-NiL rod specimen is radially
hardened steels (M50-NiL and Pyrowear 675) are commonly used loaded by three silicon nitride (Si3N4) balls, and thus three Hertzian
A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131 121

Fig. 1. (a) Schematic of the ball-on-rod RCF test setup consisting of M50-NiL rod and three Si3N4 balls. (b) Test tracks formed on the surface M50-NiL rod. Each track
corresponds to one RCF test conducted for specific number of cycles. (c) Sectioned, polished and etched test rod revealing the RCF affected regions. (d) micro-Vickers hardness
measurements along the centerline of the RCF region. (e) Plot of increase in the local hardness (DH) (primary axis) and corresponding increase in the yield strength (secondary
axis) within the RCF region (compared to virgin material) as a function of centerline depth after 246 million cycles [2].

Table 1
contacts are thus established at the ball–rod interfaces as shown in RCF test conditions for M50-NiL rod testing [2,30].
Fig. 1a1. The maximum Hertzian stress induced at the ball–rod
interface was chosen to be approximately 5.5 GPa in the tests con- Maximum Hertzian Semi-major contact Speed (rpm)
contact (MPa) width (lm)
ducted by Bhattacharyya et al. [2]. The high Hertzian stress is used
to accelerate fatigue damage since nominal stresses can take a very 5500 475 3600

large number of cycles to nucleate a fatigue spall. The central test rod
is rotated by an external motor at a speed of 3600 rpm such that it
experiences approximately 8600 RCF cycles per minute [29]. The test
conditions are summarized in Table 1 [2,30]. Additional details about depths, the hardness increase (DH) decreases, and eventually
this test can be found in [2,29]. reduces to zero at a depth of about 450 lm. This depth essentially
In each test, the rod is subjected to a specified number of RCF marks the end of the RCF affected region or the boundary
cycles, and upon completion it is advanced axially to expose the separating elastic and plastic regions. The decreasing trend of the
new virgin material. The test is then repeated for increasing num- cyclic hardening (DH) as a function of depth (Fig. 1e) can be attrib-
ber of cycles. This procedure facilitates multiple RCF tests to be uted to the diminishing values of contact stresses at increasing
conducted on a single rod specimen. The RCF test tracks formed depths.
on the surface of a cylindrical rod are shown schematically in The experimentally measured hardness values can be related to
Fig. 1b. The distance between two adjacent test tracks is approxi- the corresponding yield strengths using the Tabor’s rule [31] with a
mately 4000–6000 lm (4–6 mm). Each track corresponds to one constraint factor of 2.5 for the M50-NiL steel [1]. The resulting
RCF test conducted for specific number of cycles (N1, N2, N3 etc.). increase in yield strength (Dry ) is shown on the secondary axis
The tested rods were longitudinally sectioned, polished and etched of Fig. 1e. In addition to the magnitude of yield strength increase,
with 3% Nital to reveal the RCF affected region as shown schemat- the rate of cyclic hardening can also be studied by plotting the evo-
ically in Fig. 1c. To investigate and to quantify the subsurface lution of yield strength (ry ) as a function of number of cycles. Fig. 2
changes in this small RCF affected region, micro-Vickers hardness shows the plot of rate of cyclic hardening at a depth of 75 lm (from
measurements were performed along its centerline as shown in the surface) within the RCF affected region. The total extent of cyc-
Fig. 1d. It was observed that RCF affected region experiences a lic hardening i.e. increase in yield strength (Dry ) at 75 lm depth is
noticeable increase in local hardness (DH) after millions of fatigue about 412 MPa (=1030/2.5) as shown in Fig. 1e. Out of this, signif-
cycles [2]. Fig. 1e shows the plot of this increase in hardness (DH) icant hardening of about Dry  286 MPa (=3174–2888 MPa) takes
within the RCF affected region (compared to the virgin hardness place during initial 13.5 million RCF cycles (see Fig. 2). Thereafter,
before RCF testing) as a function of centerline depth from the sur- over next 232 million cycles, the rate of hardening gradually
face, after 246 million RCF cycles [2]. The maximum increase in decreases, during which the yield strength increases by only about
hardness (DH) of about 1030 MPa was observed at a depth of Dry  126 MPa (=3300–3174 MPa). The decreasing rate of yield
75 lm i.e. location of the indent closest to the surface. At higher strength increase (Dry ) indicates that cyclic hardening within the
RCF affected region will eventually saturate as indicated by the
dotted line in Fig. 2. The experimental results discussed here are
1
For interpretation of color in all the figures, the reader is referred to the web used in the next section to estimate the cyclic hardening parame-
version of this article. ters of the case-hardened M50-NiL steel.
122 A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131

strain (PEEQ). The governing equation for the non-linear kinematic


hardening under a uniaxial loading situation is given by,

2
q_ ¼ C e_ pl  cqe_ pl ð2Þ
3
where C and c are called as the kinematic hardening parameters. C
represents the initial kinematic hardening modulus of a material
which is a constant. However, as material accumulates plastic
strain, the instantaneous kinematic hardening modulus decreases,
and the rate of its decrease is controlled by the parameter c. The
backstress tensor and its increment is represented by q and q_
respectively, while e_ pl denotes the increment in accumulated effec-
tive plastic strain. The FE model uses the multiaxial generalization
of the non-linear kinematic hardening rule (Eq. (2)). The response
of a material using NIKH model is thus governed by four cyclic hard-
ening parameters Q, b, C and c, which are required to be estimated.
The virgin (RCF unaffected) and the RCF affected hardness val-
Fig. 2. Rate of cyclic hardening at a depth of 75 lm within the RCF affected region
ues of M50-NiL as a function of depth are measured by Bhattachar-
[2].
yya et al. [2]. These are converted to the corresponding yield
strength values using a constraint factor of 2.5 and are summarized
3. Estimation of cyclic hardening parameters in Table 3.
The virgin (RCF unaffected) hardness values when converted to
M50-NiL is a case-hardened bearing steel which consists of hard yield strength give values of the initial yield strength (ry0 ) at vari-
vanadium carbide particles in its microstructure. The carbide vol- ous depths within the case layer of M50-NiL. On the other hand,
ume fraction decreases from approximately 12.7% at the surface RCF affected yield strength values (after 246 million cycles) corre-
to 1.8% over a depth over 1.8 mm as shown in Table 2 [1]. This spond to the saturation of cyclic hardening discussed earlier (see
gradation results in variation of material properties such as elastic Fig. 2). The difference between initial and saturation values essen-
modulus, hardness and yield stress as a function of depth [1,2,26]. tially gives the total increase in yield strength ððDry Þmax Þ as plotted
It is expected that even the cyclic hardening properties will vary in Fig. 1e (secondary axis). This total increase ððDry Þmax Þ can be split
over this depth of case layer. Therefore, different sets of cyclic into two combined hardening components, namely, the isotropic
hardening parameters of NIKH model corresponding to various hardening ðDry Þiso and the kinematic hardening ðDry Þkin given by,
case-layer depths are required to be estimated. This parameter ðDry Þmax ¼ ðDry Þiso þ ðDry Þkin ð3Þ
estimation procedure is explained in the following paragraphs.
The Non-linear Isotropic/Kinematic Hardening (NIKH) material This idea of splitting the total hardening into isotropic and kine-
model used in ABAQUS 6.11 [32] is based on the von-Mises yield matic hardening has been previously used by researchers [17,33].
criterion with an associative flow rule and an additive decomposi- Under the multi-axial loading condition, Eq. (3) can be expressed
tion of the total strain tensor. The evolution equation for the non- in terms of cyclic hardening parameters as [32,34],
linear isotropic hardening is given by, rffiffiffi
2C
bepl ðDry Þmax ¼ Q þ ð4Þ
ry ¼ ry0 þ Qð1  e Þ ð1Þ 3c
where ry0 is the initial (or virgin) yield strength of material that It should be noted that Eq. (4) represents the maximum possible
represents the initial size of yield surface. Q and b are called as increase in the yield strength of material corresponding to the
the isotropic hardening parameters, in which Q represents the saturation of cyclic hardening as observed in experiments (Fig. 2).
amount of isotropic hardening i.e. maximum expansion of yield sur- We define the ratio of contribution of kinematic component
face, and b represents the rate of isotropic hardening i.e. the rate at ðDry Þkin to the total hardening ðDry Þmax as the mixity ratio (R).
which yield surface expands. epl is the accumulated effective plastic The value of R = 0 corresponds to 0% kinematic hardening, while

Table 2
Variation of carbide volume fraction as a function of depth for M50-NiL [1].

Depth (lm) 25 100 300 500 750 1000 1300 1800


Carbide volume fraction (%) 12.7 12.4 12.0 11.1 10.0 8.6 6.4 1.8

Table 3
Virgin and RCF affected values of hardness and yield strength of M50-NiL as a function of depth. Constraint factor of 2.5 is used to convert hardness to yield strength [1,2].

Depth (lm) Hardness (Hv) Yield Strength (MPa)


Virgin RCF affected Virgin RCF affected Increase ððDry Þmax Þ

75 736 841 2888 3300 412


150 731 827 2868 3245 377
225 718 792 2817 3108 291
300 705 758 2766 2974 208
375 695 730 2727 2865 137
450 683 694 2680 2723 43
A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131 123

R = 1 represents 100% kinematic hardening. In the absence of any of 75 lm from the surface. As shown in Table 3, the initial (virgin)
cyclic data, it is difficult to precisely determine the individual con- yield strength (ry0 ) at this depth is 2888 MPa and the total
tributions of isotropic and kinematic hardening components for increase in yield strength (until saturation) i.e. ðDry Þmax is about
M50-NiL steel. Therefore, as an initial guess, it is assumed that 412 MPa (i.e. =3300–2888 MPa). This measured (yield strength)
50% of the total hardening is due to the kinematic component, increase essentially represents an average increase in the hard-
while the remaining 50% is due to the isotropic component. This ness over a projected area of Vickers indent which is about
assumption corresponds to the value of R = 0.5. Although this 20  20 lm [2]. The microstructure of M50-NiL bearing steel con-
assumption may not be accurate, it facilitates the parameter esti- sists of hard carbide particles (about 1 lm in size) which are
mation procedure as discussed in the following paragraphs. In Sec- formed during the carburizing heat treatment [1]. Being smaller
tion 6.1, this assumption of R = 0.5 is generalized to study the than the size of an indent, some carbide particles lie inside the
influence of various mixity ratios (e.g. R = 0.1 and R = 0.9 etc.) on projected area of an indent. These carbide particles introduce
the cyclic hardening behavior of M50-NiL. stress concentration in the neighboring steel matrix, and assist
In the absence of relevant cyclic strain controlled experimental the cyclic hardening process due to increased plastic strain accu-
data for M50-NiL, we estimate the value of kinematic hardening mulation in their vicinity. The severity of stress concentration
parameter C (see Eq. (2)) based on the following argument. Within depends upon the location of a material point with respect to car-
the elastic limit, material follows the stress–strain curve with a bide. Therefore, it is possible that different material points (inside
slope equal to its elastic modulus (E). For M50-NiL this value of elas- an indent) will undergo different extents of cyclic hardening.
tic modulus is approximately 225 GPa over the depth of 500 lm Moreover, in the close proximity of a hard carbide particle, the
from the surface as shown in Fig. 3 [1]. Once the stress in material local increase could be significantly higher than the average
exceeds the yield strength, it starts to accumulate plastic strain and increase measured by an indent. To account for this higher
experiences transition from elastic to plastic state. During this tran- increase in local hardness (yield strength) near a carbide particle,
sition, the elastic modulus (E) changes to the hardening modulus. a multiplication factor of 1.5 is used. This factor is chosen such
As material continues to accumulate plastic strain, the hardening that the local hardening in the vicinity of carbide does not exceed
modulus keeps on decreasing (see Fig. 3). But, within the elastic– 3500 MPa. Allison et al. [27] performed compression tests on
plastic transition region when accumulated plastic strain is small, the miniature specimens extracted from within the RCF affected
the initial hardening modulus is not significantly different from region of M50 steel (whose ultimate strength properties are com-
the elastic modulus (E). Therefore, parameter C which is the initial parable to that of M50-NiL case layer). These tests showed that
kinematic hardening modulus just after the yield point, can be M50 steel can harden to stress levels of 3500 MPa. In the
considered as the transition modulus from elastic to plastic regime, absence of similar data for M50-NiL, it is assumed that M50-NiL
well before the steady state hardening sets in. During this can also locally harden to 3500 MPa. Based on this argument,
transition, C decreases from a value close to elastic modulus the multiplication factor of 1.5 is applied to the increase in yield
(E  225 GPa) [1] to a slightly lower value. Thus, we choose the strength of material during RCF i.e. amount of hardening. At the
value of C = 210 GPa, which is close to the elastic modulus but slightly depth of 75 lm, the total increase in the yield strength
lower. Since the elastic modulus does not vary significantly over a ððDry Þmax Þ is 412 MPa (=3300–2888 MPa). With the factor of 1.5,
depth of 500 lm [1], it is reasonable to assume that C does not vary ððDry Þmax Þ can now be updated to 618 MPa, which is 1.5 times
as a function of depth. Therefore a constant value of C = 210 GPa is 418 MPa. Addition of 618 MPa to the virgin yield strength of
used for depths up to 500 lm which contain the RCF affected 2888 MPa results in 3506 MPa (i.e. limited to 3500 MPa). A bet-
region. As material accumulates further plastic strain, C decreases ter estimate of this (stress) multiplication factor can be obtained
such that C2 < C1 < C etc., as shown in Fig. 3. This decrease in C is by measuring the hardness in the very close proximity of carbide
governed by parameter c Once the value of C = 210 GPa is esti- using nanoindentation technique and dividing it by a hardness
mated, other parameters can be obtained as follows. value away from carbide. However, a factor of 1.5 is considered
reasonable in the present study. Using the updated value of total
3.1. Sample estimation of cyclic hardening parameters increase in yield strength (ððDry Þmax ¼ 618 MPaÞ), values of other
parameters are now estimated.
Based on above discussion, sample parameter estimation pro- With the mixity ratio of R = 0.5, total increase in yield strength
cedure is presented for the properties of M50-NiL steel at a depth ððDry Þmax Þ of 618 MPa can be split into two equal components of
309 MPa corresponding to each of the isotropic and kinematic
hardening respectively. From Eqs. (3) and (4), the isotropic compo-
nent of 309 readily gives Q = 309 MPa. On the other hand,
C = 210 GPa and kinematic component of 309 MPa can be used to
determine the value of parameter c ¼ 555 using Eq. (4). The only
remaining parameter is b which describes the rate at which size
of yield surface (i.e. Q) increases. Being a rate parameter, b does
not affect the magnitude of isotropic hardening Q. In other words,
changing the value of b will only alter the number of cycles
required by a material to undergo isotropic hardening of Q. It is
observed that higher the value of b, faster is the rate of isotropic
hardening and vice versa. As the amount of cyclic hardening is
not affected by the choice of parameter b, we choose b = 40 for
all the simulations, in absence of any available data. In this way,
all the four cyclic hardening parameters of the case-hardened
M50-NiL steel are estimated at the depth of 75 lm. The above
described method of parameter identification is summarized in
Fig. 4. The procedure described in Fig. 4 is then repeated to esti-
Fig. 3. Schematic of stress–strain curve illustrating the initial kinematic hardening mate the cyclic hardening parameters at remaining depths within
modulus C. the case layer of M50-NiL. The parameters thus obtained are listed
124 A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131

Table 5
Cyclic hardening parameters and other test conditions used for uniaxial ratcheting
study [19].

ry0 rmax rmin Q C b c


MPa
120 250 50 120 218,500 13.20 1956.60

every loading cycle in the presence of non-zero mean stress. Being


a strain accumulation mechanism, researchers are primarily inter-
ested in studying the ratcheting strain response under various
combinations of mean and amplitude stresses [35–38]. Such stud-
ies generally overlook the cyclic hardening behavior (i.e. increase in
the yield strength) which simultaneously takes place with increas-
ing number of loading cycles. In the present context, such increase
in yield strength as a result of ratcheting can provide a mechanistic
explanation to the experimentally observed cyclic hardening
within the RCF affected region (Fig. 1).
To illustrate the concept of increase in yield strength via ratchet-
ing, uniaxial cyclic stress controlled simulations were performed
using commercial software MATLAB. The cyclic plasticity evolution
equations used in these simulations were based on the combined
non-linear isotropic/kinematic hardening (NIKH) model. For the
demonstration of concept, we selected 316 stainless steel for
Fig. 4. Flowchart for estimation of four cyclic hardening parameters of the NIKH which cyclic hardening parameters are already available in the lit-
model.
erature, and it is also known to manifest ratcheting under uniaxial
loading conditions [19]. In addition, this material possesses uni-
Table 4
form microstructure (without gradation of material properties as
Depth-dependent cyclic hardening parameters within the case-hardened layer of in case-hardened steels) which makes the interpretation of results
M50-NiL for a mixity ratio of R = 0.5. easier.
Depth ry0 Q C b c Table 5 lists the cyclic hardening parameters along with the
uniaxial test conditions used for simulations [19]. A uniaxial spec-
lm MPa
imen is subjected to a stress controlled cyclic loading (Fig. 5a.)
75 2888 309 210,000 40 555 between rmax = 250 MPa and rmin = 50 MPa with a tensile mean
150 2868 283 607
stress of 100 MPa for 100 cycles. The plot of von-Mises stress vs.
225 2817 218 787
300 2766 156 1099 equivalent plastic strain (PEEQ) is shown in Fig. 5b. In the first
375 2727 103 1665 cycle, material yields at 120 MPa and then strain hardens until it
450 2680 32 5297 reaches the stress limit of rmax = 250 MPa. It is followed by load
reversal to 50 MPa and then reloading to 250 MPa. This process
continues for 100 cycles and in every cycle accumulation of rat-
in Table 4. It should be noted that these are the very first estimates of cheting strain takes place under stress controlled conditions. It
the cyclic hardening parameters for case-hardened bearing steel. can be observed from Fig. 5c that with increasing number of cycles,
Parameters listed in Table 4 serve as the input to FE model of ratcheting strain (blue cross) continuously accumulates and it is
ball-on-rod RCF test (Section 5) used to simulate the cyclic harden- accompanied by a proportional increase in the yield strength of
ing behavior observed in experiments (Fig. 2). But, before discuss- material (red dots) which is nothing but the cyclic hardening. Also
ing FE modeling details, it is beneficial to investigate the physics note that the incremental ratcheting strain (Der ) keeps on decreas-
(i.e. the cyclic hardening mechanism) associated with the local ing in every subsequent cycle, and after few cycles, it approaches
increase in yield strength (Dry ) within the RCF affected region. Rat- zero. Proportionally, the incremental increase in yield strength
cheting is a mechanism of incremental plastic strain accumulation (Dry ) also decreases in every cycle and eventually leads to satura-
in every loading cycle, and it is commonly observed in applications tion when yield strength equals 250 MPa. Such decreasing rate of
involving RCF [5,28]. As continuous plastic strain accumulation can cyclic hardening followed by the saturation is also observed in
be related to the hardening of a material over number of cycles, the experiments (Fig. 2). From this correlation, the ratcheting phe-
ratcheting seems to be the primary driver for cyclic hardening nomenon can be reinterpreted to explain the underlying physics of
observed in experiments. For better understanding of this phenom- cyclic hardening observed in the RCF affected region.
enon, Section 4 discusses ratcheting (which is primarily a strain The red dots in Fig. 5b and c indicate instantaneous yield
accumulation mechanism) as a cyclic hardening mechanism respon- strength (ryi ) of material which continuously increases as material
sible for increase in yield strength of a material with increasing cycles. cyclically hardens. Based on this continuously increasing instanta-
To gain confidence in our approach, we have first illustrated the neous yield strength (ryi ), and the constant upper stress limit
concept for a uniaxial loading situation (Section 4), which is then (rmax ), an important conclusion can be drawn about the extent of
generalized to a multiaxial RCF problem in Section 5. cyclic hardening a material can experience. In the first cycle, instan-
taneous yield strength is same as the initial yield strength of
4. Ratcheting: A cyclic hardening mechanism (1D analogy) 120 MPa; therefore, the stress window (defined as rw ¼ rmax  ryi )
available for the accumulation of plastic strain is maximum and
Ratcheting is a mechanism of plastic strain accumulation during equal to 130 MPa (=250–120 MPa). Because of this maximum stress
which a material experiences an incremental cyclic plasticity in window (rw ), material accumulates highest increment of ratcheting
A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131 125

5. Finite element modeling of the ball-on-rod test

Once the mechanism of cyclic hardening is understood, experi-


mentally observed yield strength increase can be simulated using
an elastic–plastic FE model of the ball-on-rod RCF test. Hard car-
bide particles present in the microstructure of M50-NiL are not
only responsible for stress concentration, but they also influence
the local hardness measurements within the RCF affected region.
Moreover, these carbide particles introduce a non-zero mean shear
stress in the neighboring steel matrix during RCF. It was shown in
the previous study by the authors that such non-zero mean shear
stress facilitates ratcheting strain accumulation in the vicinity of
carbide particles [17]. This local ratcheting is responsible for cyclic
hardening of M50-NiL. Therefore, incorporating the effect of
carbide microstructure in the FE model becomes crucial for
realistic simulation of the experimentally observed increase in
yield strength. To achieve this, a two-step ‘global model-submodel’
approach is employed as shown in Fig. 6. The global model simu-
lates the overall ball-on-rod RCF test (Fig. 6a) as described in Sec-
tion 2. Macroscopic stress–strain fields induced in the vicinity of
ball–rod contact are obtained from this simulation. On the other
hand, submodel focusses on a small region of interest within the
global model and takes into account the influence of hard carbide
particles. Stress fields obtained from the global model are applied
as stress boundary conditions on the edges of submodel to deter-
mine the local (micro-scale) stress/strain fields in the vicinity of
carbide particles. Such two-step modeling approach for simulation
of the ball-on-rod RCF test was used in the previous work by the
authors [17]; therefore, only important modeling details are dis-
cussed here.
Fig. 6a shows the 2D plane strain FE model of ball-on-rod test
with three Si3N4 balls which apply radial load on the central
M50-NiL rod. Three Hertzian contacts are thus established with
the maximum Hertzian stress of 5.5 GPa. As the steel rod rotates,
every material point within the contact region experiences three
RCF cycles at each ball–rod interface per revolution. Simulation
of one complete revolution of M50-NiL rod takes approximately
40 h on a multiprocessor computer. As a result, to complete the
Fig. 5. (a) Stress controlled loading used for uniaxial ratcheting simulation. (b) analysis in a timely manner, only two revolutions of rod are simu-
Illustration of cyclic hardening phenomenon (increase in yield strength) as a
lated, which correspond to six RCF cycles. The M50-NiL test rod is
function of equivalent plastic strain. (c) Correlation between the ratcheting strain
accumulation and cyclic hardening behavior as a function of cycles.
modeled using 4 node plane strain, coupled temperature-displace-
ment elements (CPE4T) which allow users to specify temperature
dependent material properties. Such temperature dependence
strain (and also the equivalent plastic strain) in the first cycle. Due facilitates the specification of depth-dependent material proper-
to cyclic hardening, the instantaneous yield strength (ryi ) increases ties. The semi-minor contact width between ball and rod is about
continuously in subsequent cycles, while the upper stress limit 250 lm. Fig. 6b shows the von-Mises stress distribution in the sub-
(rmax ) remains constant resulting in continuous decrease of the surface of contact region. The subsurface region in the vicinity of
stress window (rw ). Decreasing rw results in decreasing ratcheting contact which deforms plastically and experiences cyclic harden-
strain increments (Der ) in subsequent cycles, which implies ing is called the RCF affected region. It typically extends up to a
decreasing rate of ratcheting (blue curve in Fig. 5c). This is accom- depth of about 200–400 lm [2].
panied by decreasing rate of cyclic hardening (indicated by red To capture the local increase in hardness (yield strength)
curve Fig. 5c).When the stress window (rw ) becomes zero (i.e. measured by a Vickers indent via FE simulations, a small region
ryi ¼ rmax ), the material is said to have cyclically hardened to the comparable to the size of an indent is considered within the RCF
stress level of rmax and thus, saturated. Further cyclic loading does affected zone. The small blue square (or diamond) shown in
not cause any yielding of material; therefore, accumulation of rat- Fig. 6b represents one of such indents. The experimentally mea-
cheting strain does not take place. This results in a phenomenon sured yield strength increase (shown in Table 3) is the average
called arrest of ratcheting after which the material response is an increase in hardness (converted to yield strength) measured within
elastic shakedown (i.e. stresses remain within the elastic limit). the projected area of Vickers indent which is about 20  20 lm [2].
As discussed earlier, such elastic response after arrest/shakedown This projected area encompasses hard carbide particles (see
of ratcheting is crucial for extending the bearing life. From above Fig. 6c), which influence the local hardness measurements. The
discussion, it is clear that cyclic hardening (i.e. increase in yield small volume fraction (12.7%) and uniform distribution of spher-
strength with number of cycles) is possible even under the condi- ical carbides within the microstructure of M50-NiL allows us to
tions of stress control via ratcheting. Furthermore, the upper stress assume that they do not interact with each other [1]. Therefore,
limit (rmax ) controls the extent of cyclic hardening a material can for modeling purpose, we consider a submodel with only a single
experience via ratcheting. carbide particle and assume that it represents the influence of all
126 A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131

Fig. 6. (a) Global FE model of the ball-on-rod RCF test with central M50-NiL rod radially loaded by three Si3N4 balls. (b) von-Mises stress distribution in the RCF affected zone
formed near contacting surfaces. The blue square (diamond) in this zone represents a Vickers indent used to measure the local hardness. (c) Schematic of Vickers indent
encompassing uniformly distributed spherical carbide particles. (d) Small region within this indent with single carbide is chosen as the submodel for simulation of yield
strength increase. This submodel is driven by stress fields obtained from the global model. (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

particles present inside the projected area of an indent. The sche- deformation. It should be noted that presence of carbides affects
matic of such submodel consisting of single carbide particle at its the yield strength (hardness) measurements of both the virgin
center is shown in Fig. 6d. In FE simulations, yield strength increase and RCF affected material. But, we assume that their influence on
obtained within the submodel is thus assumed to be representative the hardness measurement remains same even after the material
of the average yield strength increase over entire area of an indent. has undergone severe RCF loading. Based on the above discussion,
As shown in Fig. 1e and Table 3, the increase in yield strength var- methodology adopted to simulate the experimentally observed
ies as a function of depth within the RCF affected zone. Therefore, yield strength increase within the submodel using FE modeling is
different submodels are required to simulate the varying yield explained in the following paragraphs.
strength increase as a function of depth (Fig. 1e). The size of each In a multiaxial loading situation with J2 plasticity, the
submodel is calculated based on the carbide volume fraction at von-Mises stress governs yielding of a material; therefore, it can
respective depths (Table 2) using interpolation and the knowledge be considered as the measure of material’s yield strength. Fig. 7a
of carbide size (1 lm) [1]. shows a typical von-Mises stress distribution in the vicinity of car-
In submodel, the carbide particle is modeled as an elastic bide particle inside a submodel during RCF. On the other hand,
material (E = 490 GPa) because it is relatively harder and stiffer Fig. 7b shows variation of von-Mises stress experienced by an arbi-
compared to the surrounding steel matrix [1]. As a result, the trary finite element inside the submodel during six RCF cycles.
increase in yield strength (hardness) takes place only in the steel Following important observations can be made from Fig. 7.
matrix which deforms plastically and experiences cyclic hardening
via ratcheting. Carbides on the other hand do not contribute to this (i) In the absence of a carbide particle, the von-Mises stress dis-
local increase in yield strength as they do not undergo any plastic tribution is uniform as shown in Fig. 6b. But incorporation of

Fig. 7. (a) Typical von-Mises stress distribution in the vicinity of carbide particle inside a submodel. (b) Variation in the von-Mises stress at an arbitrary finite element (Black
Square) during six RCF cycles.
A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131 127

carbide alters this stress field in the surrounding steel matrix value of yield strength (ry ¼ ðrVM Þmax ) at all finite elements (in the
and also causes stress concentration which results in ampli- steel matrix inside the submodel) is used. This volume averaged
fied von-Mises stress values (Fig. 7a). Severity of this stress yield strength ðry ÞFEA is calculated using,
concentration depends on the location of a particular finite PN
element relative to the carbide particle. Depending on this n¼1 V n ððrVM Þmax Þn
ðry ÞFEA ¼ PN ð5Þ
location, different finite elements within the submodel n¼1 V n
experience different magnitudes of peak von-Mises stress.
(ii) Although carbide particle alters the von-Mises stress distri- where Vn and ððrVM Þmax Þn respectively represent volume and peak
bution in the neighboring matrix, nature of the RCF loading von-Mises stress in a given finite element n. The total number of
cycle still remains that of a stress controlled test in which finite elements in the steel matrix of a submodel is denoted by N.
magnitude of the peak stress remains constant from one The volume-averaged value i.e. ðry ÞFEA thus obtained represents
cycle to the next as shown in Fig. 7b. the average RCF affected yield strength inside the submodel corre-
sponding to the saturation of material. This value can be compared
By definition, ratcheting phenomenon can only be observed in a with the experimentally measured yield strength (Table 3) at the
cyclic stress controlled experiments in the presence of mean stress. respective depth within the RCF affected region, to test the accu-
In the present FE model of RCF test, hundreds and thousands of racy of FE simulation. The above described yield strength matching
such stress controlled tests are simultaneously being performed at methodology is summarized in the flowchart shown in Fig. 8.
every finite element in the vicinity of carbide particle. In addition
to the stress concentration, carbide particles also introduce a 6. Results and discussion
non-zero mean stress into the shear stress cycle, which promotes
ratcheting in the steel matrix surrounding carbides [17]. As dis- Results obtained from global and submodel FE simulations of
cussed in Section 4, such ratcheting can cause cyclic hardening of the ball-on-rod RCF test are presented here. They are first
the steel matrix which results in localized yield strength increase presented for a depth of 150 lm following the steps outlined in
within the RCF affected region. Therefore, the yield strength flowchart (Fig. 8). Later, they are extended to all other depths
increase reported in experiments can be captured inside the sub- within the RCF affected region where hardness measurements
model (which represents a Vickers indent) by considering ratchet- were conducted.
ing as the cyclic hardening mechanism. The experimental increase in the local yield strength and cyclic
As discussed in Section 4, the peak stress (rmax ) in a stress- hardening parameters used for the submodel at 150 lm depth are
controlled test decides the maximum cyclic hardening (increase given in Tables 3 and 4 (Step 1). The carbide volume fraction at this
in yield strength) a material can experience via ratcheting (see depth is 12.3% [1], which is used to calculate the area of submodel
Fig. 5). In the context of RCF simulations, various finite elements (Steps 2 and 3). To determine the average increase in yield
in the steel matrix (near carbide) will experience different extents strength, it is necessary to determine the extent of cyclic hardening
of cyclic hardening up to their respective values of peak von-Mises at all points inside the submodel which represents a Vickers indent
stress i.e. ry ¼ ðrVM Þmax . Therefore, to obtain a unique value of the (Step 4). This is achieved by examining the peak von-Mises stress
new yield strength resulting from simulation, the volume averaged induced in every finite element in the steel matrix of a submodel.

Fig. 8. Procedure for simulation of the cyclic hardening behavior within the RCF affected zone.
128 A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131

As an illustration, Fig. 9 shows sample von-Mises stress cycles depth values are also normalized with respect to the semi-minor
experienced by arbitrary finite elements within the submodel axis (250 lm) and are shown by a secondary X-axis. It should
during six RCF cycles. It can be observed that all finite elements be noted that these results are specific for the chosen mixity ratio
experience stress controlled von-Mises stress cycles, with varying of R = 0.5. The blue curve at bottom indicates variation of the virgin
peak values depending on their locations relative to the carbide yield strength of M50-NiL steel rod before RCF testing. Decrease in
particle. Recall that such stress controlled cyclic loading is the the virgin yield strength of the test rod as a function of depth is a
primary requirement for ratcheting. Moreover, as discussed in result of case hardening heat treatment. On the other hand, red and
Section 5, the peak value of von-Mises stress ððrVM Þmax Þ controls black curves show the variation of yield strength values of the RCF
the level of cyclic hardening a particular element can experience affected specimen, obtained from experiments and FE simulations,
via ratcheting. For example, as shown in Fig. 9, element 1 has a respectively. A reasonably well agreement is obtained between the
peak von-Mises stress of 3071 MPa compared to 3238 MPa in ele- two values within the RCF affected region.
ment 2. These values indicate that elements 1 and 2 will cyclically The highest discrepancy (approx. 7%) between experimental
harden up to 3071 MPa and 3238 MPa, respectively, via ratcheting. and simulated yield strength values is observed at the depth of
Similarly, peak von-Mises stress values at all other finite elements 75 lm which is location of the first indent. This indent lies very
in the steel matrix of submodel are determined. Using Eq. (5), close to the surface of the test rod (see Fig. 1d). At such proximity,
the volume averaged yield strength obtained from FE simulations indentation measurements are generally affected by the edge
for submodel at the depth of 150 lm is calculated, and is effects. Prior to indentation measurements, the sectioned test rods
ðry ÞFEA = 3224 MPa (Step 5). Comparison of this simulated value are polished [2], during which they are highly susceptible to the
with the experimental value (at 150 lm) of 3245 MPa (see Table 3) edge retention which often introduces error in the hardness mea-
shows a close agreement between the two (Step 6). surement. As a result, accuracy of the measured hardness (yield
To match final yield strengths at other depths within the RCF strength) cannot be guaranteed at such depth and could be the
affected region, additional submodels were developed at respec- cause of this discrepancy. In addition, during RCF testing, wear
tive depths where experimental measurements were conducted tracks of finite depth are formed on the surface of the test rod.
(Step 7). Results of these simulations are summarized in Fig. 10 Bhattacharyya et al. [30] have carefully measured the depth of
which shows comparison between yield strength predictions from wear tracks to be approximately 7–8 lm. But, the FE model used
FE simulations and experimental measurements. The centerline in the present study does not account for the formation of wear

Fig. 9. von-Mises stress cycles experienced by arbitrary finite elements within a submodel during six RCF cycles.
A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131 129

kinematic hardening (R = 0.9) are considered. The choice of mixity


ratio affects the parameter estimation process (Section 3). There-
fore, new sets of cyclic hardening parameters corresponding to
R = 0.1 and R = 0.9 are calculated using the procedure explained
in Fig. 4. These parameters are listed in Table 6.
Using the parameters listed in Table 6, FE simulations are
repeated at all the depths within the RCF affected region and yield
strength values as a result of cyclic hardening are predicted.
Results thus obtained are shown in Fig. 11.
In Fig. 11, solid curves represent variation of the experimentally
measured virgin and RCF affected yield strengths of M50-NiL as a
function of depth. The centerline depth values are also normalized
with respect to the semi-minor axis (250 lm) and are shown by a
secondary X-axis. Three dotted curves represent yield strength pre-
dictions from FE simulations for different mixity ratios. It can be
observed that for R = 0.9 (blue curve), close agreement is observed
between simulations and experimental measurements. On the
other hand, yield strengths are significantly under-predicted when
R = 0.1 (red curve) is used for the simulations. These results imply
Fig. 10. Comparison of yield strength predictions from FE simulations and that mixity ratio has a substantial influence on modeling of the
experimental measurements within the RCF affected region for R = 0.5. The
cyclic hardening response during RCF. The choices, R = 0.5 and
secondary X-axis on the top shows the centerline depth values normalized with
respect of to the semi-minor axis (250 lm). The virgin and experimental yield R = 0.9, which correspond to 50% and 90% contribution of kine-
strength values are obtained by dividing hardness measurements (in MPa) by a matic hardening toward total cyclic hardening; result in reason-
constrain factor of 2.5. able yield strength predictions. This confirms the commonly held
belief that kinematic hardening plays a dominant role in the
track and its depth. This could also be one of the reasons for the response of a material when subjected to RCF loads [23,40,41].
discrepancy in the hardness predictions. Finally, Hertzian contact While the exact contributions of isotropic and kinematic hardening
stress theory can also provide another explanation for this discrep- components may be unknown, our simulations indicate that
ancy between simulated and experimental yield strength values.
According to Hertzian theory, when two bodies are in contact,
Table 6
the resulting Hertzian stress distribution is such that it reaches
Depth-dependent cyclic hardening parameters within the case layer of M50-NiL for
maximum value at a certain depth in the subsurface of a material mixity ratios of R = 0.1 and R = 0.9.
[39]. In the present FE simulations, the contact between Si3N4 balls
Depth ry0 Q (MPa) C (MPa) b c
and M50-NiL rod results in maximum contact stress at a depth of
(lm) (MPa)
approximately 150 lm in the subsurface of rod (see Fig. 6b). Since R = 0.1 R = 0.9 R = 0.1 R = 0.9
the peak value of von-Mises stress decides the extent of cyclic 75 2888 556 62 210,000 40 2775 308
hardening, highest peak stress at 150 lm results in highest yield 150 2868 509 57 3032 337
225 2817 393 44 3928 436
strength prediction as shown in Fig. 10. Compared to 150 lm,
300 2766 281 31 5496 611
von-Mises stress at the depth of 75 lm obtained in the FE simula- 375 2727 186 21 8283 920
tion is smaller; therefore, it results in lower cyclic hardening and 450 2680 58 6 26,584 2954
lower yield strength prediction. This difference in the location of
maximum subsurface stress can be attributed to the 2D plane
strain assumption used for FE modeling. In actual experiments, ball
on rod geometry results in an elliptical contact patch compared to
the rectangular contact patch obtained in a plane strain FE model.
This difference between contact conditions influences the distribu-
tion of subsurface stresses and also affects the location of maxi-
mum contact stress. A 3D FE model is therefore recommended to
accurately capture the subsurface stress fields to potentially mini-
mize the discrepancy between the experimental measurements
and FE predictions.
Other than 75 lm depth, the agreement between simulated and
experimental yield strengths is reasonably good with a maximum
error of 2% at the depth of 375 lm. Such agreement therefore pro-
vides first verification of the depth-dependent cyclic hardening
parameters, estimated for M50-NiL (Section 3).

6.1. Effect of isotropic and kinematic hardening mixity ratio (R) on


yield strength predictions

Results presented in the previous section are based on cyclic


hardening parameters that are estimated using the mixity ratio of
R = 0.5. In this Section, the influence of different mixity ratios on Fig. 11. Effect of different mixity ratios (R) on yield strength predictions within the
RCF affected region. The secondary X-axis on the top shows the centerline depth
the cyclic hardening response of material is investigated. For this values normalized with respect of to the semi-minor axis (250 lm). The virgin
purpose, two extreme scenarios of combined hardening model, and experimental yield strength values are obtained by dividing hardness
namely dominant isotropic hardening (R = 0.1), and dominant measurements (in MPa) by a constrain factor of 2.5.
130 A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131

material parameters estimated using any mixity ratio between in the yield strength of a material from one cycle to the next.
R = 0.5 and R = 0.9 should be suitable to capture the cyclic harden- It is shown that such cyclic hardening takes place under the
ing response under RCF conditions reasonably well. conditions of stress controlled loading, in which, the maxi-
In our work, virgin yield strength values used in simulations mum possible hardening a material can experience is con-
were based on the experimental values listed in Table 3. Later, trolled by the upper stress limit.
the RCF affected yield strength values from simulations and exper- (2) During RCF, cyclic hardening via ratcheting is promoted at
iments were compared only after 246 million RCF cycles which the scale of carbide microstructure. The hard carbide parti-
correspond to the saturation of material hardening. It implies that cles amplify stress field in the neighboring region and also
these simulations only captured the extent of hardening (i.e. introduce a non-zero mean stress in the shear stress cycle
amount of increase in yield strength) without considering the rate under RCF loading. Moreover, it is observed that nature of
of hardening as shown in Fig. 2. This fact essentially permitted us to the RCF stress cycle in the vicinity of carbides is that of a
match the end results (i.e. initial and final yield strength values) by stress-controlled test. All of these conditions are conducive
using different sets of cyclic hardening parameters, although, theo- for ratcheting of the steel matrix, which continuously accu-
retically every material should have a unique set of cyclic hardening mulates local micro-plastic strain with increasing RCF
parameters. To accurately simulate the rate of hardening, it is nec- cycles. As ratcheting is responsible for the cyclic hardening
essary to model large number of RCF cycles which was not possible of material, it is concluded that heterogeneous carbide
for present simulations due to computational limitations. A ‘Jump- microstructure is the primary driver for localized subsurface
in-cycles’ algorithm implemented by [40,42] could be a possible hardening.
approach to simulate millions of RCF cycles, but it is not considered (3) Modeling of the cyclic hardening behavior is greatly influ-
in the present manuscript. The six simulated cycles represent enced by the choice of isotropic/kinematic material model.
246 million experimental cycles in this case since we are matching It is shown that use of the dominant kinematic component
the strain accumulation via experimentally measured hardness is better to accurately capture the cyclic hardening behavior
matching, at the end of six cycles. Choosing a larger number of sim- of a material during RCF. However, it is necessary to super-
ulated cycles will refine the modeling process. However, in this pose a small isotropic hardening component that allows
study, we have clearly identified the micromechanical ratcheting modeling of realistic material behaviors in which cyclic
to be the primary driver for cyclic hardening observed within the hardening is followed by the saturation. The micro-hardness
RCF affected region of M50-NiL rod. In the same experiments con- changes manifested through cyclic plastic strain accumula-
ducted by Bhattacharyya et al. [2], regions of cyclic softening were tion at the microstructure level can be used as a fatigue indi-
also observed surrounding the region of cyclic hardening. But, cap- cation parameter (FIP) in RCF.
turing such intricate microstructural changes via simulations (4) The methodology presented in this study can be readily
requires further in-depth analysis which is the focus of future translated to estimate the localized cyclic hardening proper-
research. ties of other through and case-hardened bearing steels.
Moreover, it can also be extended to variety RCF applications
7. Summary such as rail-wheels, transmission gears and cam and tappet
mechanisms as well.
Highly localized subsurface damage caused by complex triaxial
and non-proportional stresses over large number of cycles is the
key feature of RCF induced failures. To analyze such failures, accu- Acknowledgements
rate understanding of material response at the scale of carbide
microstructure is essential and it requires development of sophisti- This research work has been partially supported by the National
cated experimental and numerical techniques. These techniques Science Foundation Award (CMMI-0927849). The authors would
will enable researchers to develop material-specific numerical like to thank William P. Ogden at Pratt & Whitney (GOALI sponsor),
tools to monitor the progression of material damage eventually East Hartford, CT, Dr. Bryan Allison at SKF USA Inc., and Drs. Nelson
leading to bearing failures. But, the absence of relevant cyclic hard- Forster and Lewis Rosado at the Air Force Research Laboratories
ening material properties of high strength bearing steels has (WPAFB, Ohio) for valuable discussions.
always been a roadblock for such quantitative modeling endeavors.
This inability to obtain localized cyclic hardening data fundamen-
References
tally stems from the highly localized nature of the RCF, at which
scale the conventional testing methods cannot be practically [1] Klecka MA, Subhash G, Arakere NK. Microstructure-property relationships in
implemented. Inspired from this challenge, we have developed M50-NiL and P675 case-hardened bearing steels. Tribol Trans
an alternative methodology to estimate the cyclic hardening 2013;56:1046–59.
[2] Bhattacharyya A, Subhash G, Arakere N. Evolution of subsurface plastic zone
parameters of the case-hardened bearing steels. Specifically, the due to Rolling Contact Fatigue of M-50 NiL case hardened bearing steel. Int J
depth-dependent cyclic hardening parameters have been identi- Fatigue 2014;59:102–13.
fied for case-hardened M50-NiL steel, based on the results of [3] Sadeghi F, Jalalahmadi B, Slack TS, Raje N, Arakere NK. A review of Rolling
Contact Fatigue. J Tribol 2009;131:041403.
micro-hardness measurements conducted on the RCF tested spec- [4] Ekberg A, Kabo E. Fatigue of railway wheels and rails under rolling contact and
imens. Based on logical assumptions, this method provides mean- thermal loading—an overview. Wear 2005;258:1288–300.
ingful estimates of cyclic hardening parameters for the very first [5] Ringsberg JW. Cyclic ratchetting and failure of a pearlitic rail steel. Fatigue
Fract Eng Mater Struct 2000;23:747–58.
time. Using these parameters, cyclic hardening behavior of the
[6] Becker PC. Microstructural changes around non-metallic inclusions caused by
M50-NiL steel is investigated under RCF loading conditions. In par- Rolling Contact Fatigue of ball-bearing steels. Met Technol 1981;8:234–43.
ticular, the localized cyclic hardening driven by the microstructural [7] Voskamp AP, Österlund R, Becker PC, Vingsbo O. Gradual changes in residual
stress and microstructure during contact fatigue in ball bearings. Met Technol
heterogeneities such as carbide particles is studied. Following
1980;7:14–21.
important conclusions are drawn from the present study. [8] Grabulov A, Petrov R, Zandbergen HW. EBSD investigation of the crack
initiation and TEM/FIB analyses of the microstructural changes around the
(1) Ratcheting, which is conventionally perceived as a strain cracks formed under Rolling Contact Fatigue (RCF). Int J Fatigue
2010;32:576–83.
accumulation mechanism, can also be viewed as a cyclic [9] Österlund R, Vingsbo O. Phase changes in fatigued ball bearings. Metall Trans A
hardening mechanism responsible for continuous increase 1980;11:701–7.
A.S. Pandkar et al. / International Journal of Fatigue 73 (2015) 119–131 131

[10] Voskamp AP. Fatigue and material response in rolling contact, bearing steels: [27] Allison BD, Subhash G, Arakere NK, Chin H, Haluck D, Yamaguchi H. Extraction
into the 20th century. ASTM STP 1328; 1998. p. 152–66. and testing of miniature compression specimens from bearing balls subjected
[11] Voskamp AP. Material response to rolling contact loading. J Tribol to Rolling Contact Fatigue. J Tribol 2014;136. 021103–021103.
1985;107:359–64. [28] Kumar A, Hahn G, Rubin C. A study of subsurface crack initiation produced by
[12] Kabo E, Ekberg A. Material defects in Rolling Contact Fatigue of railway Rolling Contact Fatigue. Metall Trans A 1993;24:351–9.
wheels—the influence of defect size. Wear 2005;258:1194–200. [29] Glover D. A ball–rod Rolling Contact Fatigue tester. In: Rolling Contact Fatigue
[13] Kabo E. Material defects in Rolling Contact Fatigue — influence of overloads testing of bearing steels. Baltimore, MD: ASTM STP 771, American Society for
and defect clusters. Int J Fatigue 2002;24:887–94. Testing and Materials; 1982. p. 107–24.
[14] Ringsberg JW. Life prediction of Rolling Contact Fatigue crack initiation. Int J [30] Bhattacharyya A, Pandkar A, Subhash G, Arakere N. Cyclic constitutive
Fatigue 2001;23:575–86. response and effective S–N diagram of M50 NiL case-hardened bearing steel
[15] Arakere NK, Subhash G. Work hardening response of M50-NiL case hardened subjected to Rolling Contact Fatigue. J Tribol, submitted for publication.
bearing steel during shakedown in Rolling Contact Fatigue. Mater Sci Technol [31] Tabor D. The hardness of solids. In: Review of physics in
2012;28:5. technology. Cambridge: Cavendish Laboratory; 1970. p. 145–79.
[16] Lemaitre J, Chaboche J. Mechanics of solid materials. Cambridge, UK: Press [32] ABAQUS Analysis User’s Manual, Version 6.11. In: Dassault Systèmes Simulia
Syndicate of University of Cambridge; 1990. Corportion. Providence, RI, USA; 2011.
[17] Pandkar AS, Arakere N, Subhash G. Microstructure-sensitive accumulation of [33] Abdel-Karim M, Khan A. Cyclic multiaxial and shear finite deformation
plastic strain due to ratcheting in bearing steels subject to Rolling Contact responses of OFHC Cu. Part II: an extension to the KHL model and
Fatigue. Int J Fatigue 2014;63:191–202. simulations. Int J Plast 2010;26:758–73.
[18] Kabo E, Ekberg A. Fatigue initiation in railway wheels—a numerical study of [34] Jirásek M, Bažant ZP. Inelastic analysis of structures. Wiley; 2002.
the influence of defects. Wear 2002;253:26–34. [35] Xia Z, Kujawski D, Ellyin F. Effect of mean stress and ratcheting strain on
[19] Portier L, Calloch S, Marquis D, Geyer P. Ratchetting under tension–torsion fatigue life of steel. Int J Fatigue 1996;18:335–41.
loadings: experiments and modelling. Int J Plast 2000;16:303–35. [36] Park SJ, Kim KS, Kim HS. Ratcheting behaviour and mean stress considerations
[20] Rosado L, Forster NH, Thompson KL, Cooke JW. Rolling Contact Fatigue life and in uniaxial low-cycle fatigue of Inconel 718 at 649 °C. Fatigue Fract Eng Mater
spall propagation of AISI M50, M50NiL, and AISI 52100, Part I: Experimental Struct 2007;30:1076–83.
results. Tribol Trans 2009;53:29–41. [37] Lim CB, Kim KS, Seong JB. Ratcheting and fatigue behavior of a copper alloy
[21] Branch NA, Arakere NK, Subhash G, Klecka MA. Determination of constitutive under uniaxial cyclic loading with mean stress. Int J Fatigue 2009;31:501–7.
response of plastically graded materials. Int J Plast 2011;27:728–38. [38] Kang GZ, Kan QH, Zhang J. Experimental study on the uniaxial cyclic
[22] Branch NA, Subhash G, Arakere NK, Klecka MA. A new reverse analysis to deformation of 25CDV4.11 steel. J Mater Sci Technol 2005;21:5–9.
determine the constitutive response of plastically graded case hardened [39] Johnson KL. I. Books24x, contact mechanics. Cambridge [Cambridgeshire];
bearing steels. Int J Solids Struct 2011;48:584–91. New York: Cambridge University Press; 1985.
[23] Hahn GT, Bhargava V, Chen Q. The cyclic stress–strain properties, hysteresis [40] Warhadpande A, Sadeghi F, Kotzalas MN, Doll G. Effects of plasticity on
loop shape, and kinematic hardening of two high-strength bearing steels. subsurface initiated spalling in Rolling Contact Fatigue. Int J Fatigue
Metall Trans A 1990;21:653–65. 2012;36:80–95.
[24] Hu W, Wang CH, Barter S. Analysis of cyclic mean stress relaxation and strain [41] Bhargava V, Hahn GT, Rubin CA. Rolling contact deformation and
ratchetting behaviour of aluminium 7050. In: DSTO aeronautical and maritime microstructural changes in high strength bearing steel. Wear 1989;133:65–71.
research laboratory. Australia; 1999, pp. 39. [42] Bomidi JAR, Sadeghi F. Three-dimensional finite element elastic–plastic model
[25] Klecka MA, Subhash G, Arakere NK. Experimental and numerical modeling of for subsurface initiated spalling in rolling contacts. J Tribol 2013;136. 011402–
surface indentation response of plastically graded materials. J Eng Mater 011402.
Technol 2013;135. 041004–041004.
[26] Klecka MA, Subhash G, Arakere NK. Determination of subsurface hardness
gradients in plastically graded materials via surface indentation. J Tribol
2011;133. 031403–031403.

You might also like