Você está na página 1de 16

A Microstructure-Based Fatigue-Crack-Initiation Model

KWAI S. CHAN
This article presents results on the development of a microstructure-based fatigue-crack-initiation
model which includes explicit crack-size and microstructure-scale parameters. The current status of
microstructure-based fatigue-crack-initiation models is briefly reviewed first. Tanaka and Muras
models[1,2] for crack initiation at slipbands and inclusions are then extended to include crack size and
relevant microstructural parameters in the response equations. The microstructure-based model for
crack initiation at slipbands is applied to predicting the crack size at initiation, small-crack behavior,
and notch fatigue in structural alloys. The calculated results are compared against the experimental
data for steels and Al-, Ti-, and Ni-based alloys from the literature to assess the range of predictability
and accuracy of the fatigue-crack-initiation model. The applicability of the proposed model for treating
variability in fatigue-crack-initiation life due to variations in the microstructure is discussed.

I. INTRODUCTION

FATIGUE in engineering structures involves crack-initiation and growth processes. Current life-prediction
approaches for military gas turbines address both crack initiation and propagation lives. Typically, a gas-turbine component is designed to have (1) a minimum low-cycle fatigue
(LCF) crack-initiation life exceeding the total specified service life and (2) a crack-propagation life from an assumed
initial inspection that is twice the number of service cycles
between inspections.[3] According to Cowles,[3] the current
life-prediction systems are expensive to establish and substantiate, because large experimental databases are required.
The current life-prediction systems may be improved by
development of better crack-initiation-life models, probabilistic treatment of the variability of crack-initiation life, and
enhanced local notch analysis capability.[3] One of the limitations of current crack-initiation-life models is that they are
incapable of predicting the crack size at fatigue-crack initiation. Consequently, an initial crack size must be assumed in
the prediction of crack-growth life.
Recent advances in materials processing have facilitated
the design and manufacture of components with tailored
microstructures at desired locations. For example, a coarsegrained microstructure may be placed in regions where creep
resistance is desired, while a fine-grained microstructure
may be introduced in regions where fatigue-crack-initiation
resistance is needed. This design approach requires precise
control of the microstructure during processing as well as
detailed knowledge of the microstructure/property relationship, so that optimization of microstructures to achieve a
balance of material properties can be executed in an efficient
and effective manner. In particular, the time to material
development can be significantly reduced if computer-based
methods can be implemented to reduce the amount of iterations required to develop the desired microstructures and
properties for the intended applications. For computer-based
design of damage-tolerant materials and structures, microstructure-based models of fatigue-crack initiation and
growth are required. These models would also be essential
KWAI S. CHAN, Institute Scientist, is with the Southwest Research
Institute, San Antonio, TX 78238. Contact e-mail: kchan@swri.edu
Manuscript submitted November 9, 2001.
METALLURGICAL AND MATERIALS TRANSACTIONS A

for the development of probabilistic life-prediction methods,


since the tail ends of the defect distribution, which are often
related to the microstructural-size distribution, mostly influence fatigue life.
Prior reviews established the important roles of microstructure in fatigue-crack initiation and growth in metals.
The need to introduce a microstructural-size parameter in a
fatigue-crack-growth model was identified in a review article
by Bailon and Antolovich,[4] which compared fatigue-crackgrowth models against experimental data available at that
time. A microstructure-based fatigue-crack-growth model
for large cracks was developed by the author in an earlier
investigation.[5] Applications of the fatigue-crack-growth
model in a probabilistic life-prediction framework have also
been demonstrated for steels and a Ti alloy.[6] Microstructurebased fatigue-crack-initiation models have been developed
by Tanaka and Mura on the basis of the accumulation of
dislocation dipoles generated by irreversible slip on slipbands.[1,2] These models treat slipband crack initiation in
favorably oriented surface grains, often known as stage I
fatigue, as well as crack initiation at inclusions by slip
impingement and dislocation-dipole accumulation. With one
exception,[7] these microstructure-based crack-initiation
models have rarely been applied to structural alloys. Recent
generalization of the dislocation-dipole model by Venkataraman et al.[8,9] led to a stressinitiation life relation that
appears to predict a grain-size dependence that is contradictory to the models proposed by Tanaka and Mura.[1,2] Thus,
there exists a need to evaluate the various microstructure/
fatigue-life relationships against experimental data. In addition, there is also a need for a microstructure-based fatiguecrack-initiation model that explicitly incorporates crack size,
microstructural size, and other relevant material parameters.
This article presents results on an investigation whose aim
was to develop microstructure-based fatigue-crack-initiation
models that include explicit crack-size and microstructurescale parameters. Since fatigue is a wide topic, the present
article focuses solely on the microstructural-size effects and
the crack size at fatigue-crack initiation at ambient temperature under uniaxial cyclic loading. While they are important,
the roles of temperature, frequency, multiaxial loading, and
environment will not be addressed. In this article, the current
status of microstructure-based fatigue-crack-initiation models is briefly reviewed. Following Tanaka and Muras
VOLUME 34A, JANUARY 200343

approach,[1,2] fatigue-crack initiation is modeled by considering the accumulation of dislocation dipoles generated on
slipbands during cyclic loading. Tanaka and Muras models
for crack initiation at slipbands and inclusions[1,2] are
extended to include crack size and relevant microstructural
parameters in the response equations. The microstructurebased model is then applied to treating fatigue-crack initiation and the growth of small cracks in structural alloys such
as steels and Al-, Ti-, TiAl-, and Ni-based alloys. Stress
initiation life (S-Ni) and fatigue-crack-growth (dc/dN )
curves are computed as functions of crack length and microstructural-scale sizes. These results are compared against
experimental data from the literature to assess the range of
predictability and accuracy of the fatigue-crack-initiation
models, as well as the feasibility of their implementation
into the current damage-tolerance life-prediction framework.
Applicability of the models for treating the variability in
fatigue-crack-initiation life due to variations in the microstructure will also be discussed.
II. REVIEW OF EXISTING FATIGUE-CRACKINITIATION MODELS
Current fatigue-crack-initiation models are formulated
based on either stress or strain. For uniaxial cyclic loading,
fatigue-crack initiation in metals can be described in terms
of the stress-life (S-Nf) approach via the relation[10]

f N af

[1]

where a is the stress amplitude, f is the fatigue-strength


coefficient, Nf is the number of cycles to failure, and is
the fatigue-strength exponent. Failure is usually defined as
fatigue fracture of the specimens, but can also be defined
as initiation of a fatigue crack to a certain length or depth.
In the strain-life approach, the plastic strain range (p) is
related to the cycles to failure according to the Coffin
Manson relation, given by[11,12]
p
f N bf
2

[2]

where f is the fatigue-ductility coefficient, and b is the


fatigue-ductility exponent. Both phenomenological models
(stress-life and strain-life equations) are well known, but
neither contain parameters that describe the role of microstructure in the fatigue-crack-initiation process or the
fatigue life.
Fatigue-crack initiation is often associated with the formation of persistent slipbands (PSBs).[1317] These PSBs are
produced by random irreversible slip on surface grains that
leads to surface roughening, the formation of extrusions and
intrusions, and eventually to fatigue-crack initiation within
the PSBs. A crack-nucleation criterion based on the formation of a critical notch depth by random slip in PSBs was
proposed by Cheng and Laird.[18] This criterion, which is
essentially identical to the CoffinManson relation,[11,12] has
the form given by[18]
p
N i C
2

[3]

for metals subjected to a plastic-shear-strain range (p).


The number of fatigue cycles to crack initiation is Ni , and
C is a constant. The value of the exponent is 0.78, which
44VOLUME 34A, JANUARY 2003

was obtained by relating the slip offsets in the PSBs to


the applied-strain amplitude. Statistical considerations of
plastic-strain accumulation in a localized slipband led to a
value of 0.5 for .[9] In an earlier article,[19] Saxena and
Antolovich demonstrated that the parameter in Cu-Al
alloys varied from 0.5 to 1; this variation was related to the
stacking-fault energy and to the slip morphology. Thus, there
is experimental evidence and a general consensus that the
value of ranges from 0.5 to 1,[9] depending on the plasticstrain-localization process. Although it originated from consideration of PSB formation by irreversible random slip, Eq.
[3] by itself does not provide any insight into the role of
microstructure in the crack-initiation process.
More recent fatigue-crack-initiation models incorporate
microstructural parameters in a direct and explicit manner.
In a series of three articles, Tanaka and Mura proposed
dislocation models for treating fatigue-crack initiation at
slipbands, inclusions, and notches.[1,2,20]
Tanaka and Mura envisioned that fatigue-crack initiation
occurs by the accumulation of dislocation dipoles during
strain cycling. For crack initiation along a slipband, the
dislocation-dipole accumulation model is given by[1]
( 2k)N i1/2


8Ws
d

1/2

[4]

where is the shear-stress range, k is the friction stress of


dislocation, is the shear modulus, d is the grain size, and
Ws is the specific fracture energy per unit area along the
slipband. For crack initiation at a grain boundary, Eq. [4]
still applies, but the term Ws is replaced by the specific
fracture energy per unit area of grain boundary (WG). Similarly, the criterion for crack initiation by inclusion fracture
due to impinging slip and dislocation pileup is given by[2]
(

2k)N i1/2

4 ( )WI

1/2

[5]

where is the shear modulus of the inclusion, is the


inclusion size, and WI is the specific fracture energy per unit
area of inclusion. Other forms of the model for treating
crack initiation at debonded inclusions and notches are also
available in the literature.[2] Several modifications of the
basic model have also been made by Mura and co-workers,
which are described in a series of recent articles.[8,9,2124]
The crack-initiation model by developed Venkataraman et
al.[8,9] incorporates a surface-energy term (s) and a slipirreversibility factor (e), which has a value between 0 and 1.
The stress-life relation obtained by Venkataraman et al.[9] is
(

2k)N i

d s
0.37
eh d

1/2

[6]

which predicts a grain-size dependence that is contradictory


to Eq. [4] and experimental data.
The scaling law for crack initiation proposed by Chan[25]
is an extension of the CoffinManson LCF equation to
include a microstructure-size parameter. The modified
strain-life equation gives a grain-size dependence of fatiguelife initiation that is similar to that described in Eq. [4]. The
model of Chang et al.[26] is intended to treat fatigue-crack
initiation at inclusions as the result of slip impingement and
dislocation pileups. The form of the model by Chang et
al.[26] is similar to Eq. [5]. The statistical model by Ihara and
METALLURGICAL AND MATERIALS TRANSACTIONS A

Tanaka[27] considers the accumulation of stochastic damage


within the dislocation-cell structure formed during the
fatigue process.
Recently, Harvey et al.[28] proposed a fatigue-crack initiation based on the deepening of slip steps into a fatigue crack
during strain cycling. Slip-height displacements on the order
of nanometers were measured using atomic-force microscopy. Fatigue-crack initiation was assumed to occur when the
cumulative slip-height displacement exceeded a threshold
crack-opening displacement. According to this approach, the
number of cycles to fatigue-crack initiation is determined
by the large-crack fatigue-crack-growth threshold (Kth),
according to[28]
Ni

K2th
4ysEfphs

[7]

where E is the Youngs modulus, ys is the yield stress, p


is the plastic-strain range, f is the fraction of plastic-strain
range that produces the surface displacement, and hs is the
slip spacing.
The review indicated that only a few existing fatiguecrack-initiation models contain microstructural-size parameters. None of the existing fatigue-crack-initiation models
contain crack length in the formulation. As a result, none
of the crack-initiation models can be used to predict the size
of a fatigue crack at initiation. For models based on the
random irreversible slip process,[18,19,28] there is little distinction between the crack-initiation and crack-growth processes, since both involve deepening of the surface-slip steps.
For these models, fatigue cracks initiate early at minute
sizes and propagate to greater lengths at subsequent fatigue
loading. The advance of crack-detection techniques such as
atomic-force microscopy makes it feasible to measure crack
sizes as small as 100 nm or less.[28] However, detection of
cracks in this range is difficult and may not be practical or
necessary. The crack size at initiation is uncertain in most
structural alloys, and the ability to predict its occurrence
has not been established. Thus, there exists a need for a
quantitative definition of crack initiation as well as a methodology that can predict the cycles to initiation of a fatigue
crack of a given size.
III. FORMULATION OF CRACK SIZEBASED
FATIGUE-CRACK-INITIATION MODELS
The formulation of crack sizebased fatigue-crack-initiation models is described in this section. The dislocationdipole model of Tanaka and Mura[1] will be extended to
treat crack initiation along the slipband. Subsequently, the
approach will be utilized to derive the model for treating
crack initiation at inclusions and notches. The essence of
the fatigue-crack-initiation process by the dislocation-dipole
mechanism, operating in a surface grain, is depicted in Figure
1. During fatigue loading, irreversible slip occurs in a favorably oriented surface grain, leading to dislocation motion
on a slip plane (I) and dislocation pileups at grain boundaries.
Upon unloading, dislocations with opposite signs are activated on an adjacent plane (II), producing reverse slip and
the formation of vacancy and interstitial dislocation dipoles
at the ends of the double pileup. Representing the dislocation
pileups in terms of continuum (i.e., continuously distributed)
dislocations, Tanaka and Mura showed that the change in the
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 1A schematic showing the accumulation of dislocation dipoles generated by irreversible slip during fatigue loading. The coalescence of the
dislocation dipoles was postulated by Tanaka and Mura[1] as the process
that a fatigue crack initiates along a planar slipband.

stored strain energy (U ) of dislocations per unit thickness is


given by[1]
U

(1 v)( 2k)2

[8]

during stress cycling between 1 and 2, such that 1


2. The number of fatigue cycles to crack initiation was
then obtained by equating the stored energy in the dislocation
dipoles to the specific fracture energy (Ws) per unit area.
The energy balance leads one to[1]
2NiU 4dWs

[9]

which gives
( 2k)2 Ni

4Ws
(1 v)d

[10]

when Eq. [8] is substituted into Eq. [9]. Equation [10], first
derived by Tanaka and Mura, describes, correctly, the inverse
relationship between fatigue-crack-initiation life and the
grain size (d ). The crack size, however, is arbitrary and not
predicted by this formulation.
The crack size can be incorporated into Eq. [10] by considering the energy of the fatigue-crack formation. According
to Mura and Nakasone,[23] the Gibbs free-energy change
(G) associated with the nucleation of a fatigue crack from
a double pileup of dislocation dipoles is[23]
G We Wm 2cs

[11]

where We is the elastic strain energy stored in dislocations,


Wm is the mechanical energy released due to the opening
up of a crack, c is the half-length (or depth) of the incipient
fatigue crack, and s is the surface energy of the crack.
A detailed description of the energetics of fatigue-crack
initiation, presented in the Appendix, indicates that the Gibbs
free-energy change is a function of the fatigue cycle, as
given by[23]
G z1N 2 z2N 2 2z3Ns

[12]

where z1, z2, and z3 are given by Eqs. [A7], [A8], and [A9],
respectively. The first term (z1N 2) is the stored energy of
VOLUME 34A, JANUARY 200345

N 0, crack initiation can also be defined on the basis of


G 0, and the resulting fatigue cycle at crack initiation
would be twice as large as that predicted by the instability
criterion (G/N 0). The instability criterion is preferred
here because it yields a more conservative result. It should
also be noted that most of the work done by the external
load is expected to spent by irreversible slip and is dissipated
as heat. Only a small portion of the external work is stored
in the dislocation substructure.
The approach of Mura and Nakasone[23] is based on the
assumption that all the dislocation dipoles in the slipbands
contribute to the formation of the fatigue crack. The number
of dislocation dipoles (neq) in a single slipband is given by[8]

(a)

bh

(b)
Fig. 2Energetics of fatigue crack formation: (a) the formation of a fatigue
crack of length (2c) or crack depth (c) at the tip of a planar slipband, and
(b) the free-energy change associated with the crack initiation process.

G
0
N
and the number of cycles to reach the maximum free-energy
change is taken to be the number of cycles to fatigue-crack
initiation. At G/N 0, the free-energy change is still
positive (G 0) and crack initiation is not spontaneous,
but must overcome the surface-energy barrier. The energy
barrier can easily be provided by the stored strain energy
of the dislocation dipoles, since the dipole structure is unstable with respect to the crack configuration. Instead of G/
46VOLUME 34A, JANUARY 2003

[13]

where h is the slipband width and Weq is the strain energy


stored in the dislocation dipoles of a single or an equivalent
slipband. In general, not all dislocation dipoles contribute
to the formation of the fatigue crack, at least not immediately
at crack initiation.[8] Under this circumstance, it is more
appropriate to consider fatigue-crack initiation as occurring
only at a portion of the slipband, as shown in Figure 2. The
surface energy required to create this crack would be 2cs ,
and the strain energy within the incipient crack that would
be released to form the crack is cWeq/d. The maximum in
the Gibbs free energy (Eq. [12]) can be postulated to occur
when the condition given by[9]
cWeq
d

[14]

Weq 2ds

[15]

2cs

the dislocation dipoles, the second term (z2N ) is the elasticstrain-energy release rate (Eq. [A3]) due to the opening of
the incipient crack during the initiation process, and the third
term (2z3Ns) is the surface energy associated with the crackformation process. Equation [12] reduces to the theory of
brittle fracture[69] or to the theory of elastic/plastic fracture[70,71] subject to small-scale yielding under monotonic
loading when the stored energy of dislocation (first term) is
ignored and N is taken to be unity for monotonic loading.
The free-energy change described by Eq. [12] is shown
schematically as a function of fatigue cycle in Figure 2.
Initially, the free-energy change is positive (G 0), indicating that there is an energy barrier for crack initiation. The
free-energy change reaches a maximum where G/N
0, which indicates that the dislocation-dipole structure and
the incipient crack are energetically in quasi equilibrium.
Beyond the maximum, G/N 0 and the dislocationdipole structure is unstable with respect to crack formation.
Mura and Nakasone[23] postulated that the onset of crack
initiation occurs when the Gibbs free-energy change reaches
a maximum, as described by[23]

Weq

neq 0.05

is satisfied, leading to

at incipient crack initiation. The crack length (c) at crack


initiation is given by the number of dislocations (nc) that
contribute to crack formation, according to[9]
c ncb

[16]

where b is the magnitude of the Burgers vector. The value


of nc is obtained from Eq. [13] by replacing Weq by cWeq/d,
since only that amount of strain-energy density is required to
overcome the surface energy. This substitution leads one to[9]

bh d

nc 0.05

cWeq

[17]

which can be combined with Eq. [16] to give


2

d
c 0.005
h

[18]

at fatigue-crack initiation. Imposing the condition that s


Ws leads one to
( 2k)N i1/2

8 2
(1 v)

1/2

1/2


h c
d d

[19]

as the criterion for the initiation of a fatigue crack of length


2c or depth c. For fatigue-crack initiation in polycrystalline
materials, the applied-stress range ( ) can be related to
the shear-stress range ( ) on the slip plane through the
Taylor factor (M ), leading to
METALLURGICAL AND MATERIALS TRANSACTIONS A

( 2Mk)N i

8M2 2
(1 v)

1/2

1/2


h c
d d

[20]

when the exponent of Ni has been generalized to , where


0 1. An exponent not limited to 1/2 is necessary,
because it provides better agreement with experimental data.
Furthermore, experimental evidence and statistical simulation of fatigue-crack initiation by irreversible slip indicated
that the value of is not a constant, but depends on the
stacking-fault energy[19] and on the degree of slip irreversibility.[9,18,19]
Using a similar procedure, the governing equation for
fatigue-crack initiation by slip impingement at an inclusion
can be derived. In this case, the surface energy for slipband
cracking in Eq. [18] is replaced by the corresponding surface
energy for creating two fracture surfaces in an inclusion (i).
The resulting equation is then combined with Eq. [5] to give

( 2Mk)N i

8( )

1/2

Mh2
d(h d )

(a)

[21]
1/2

for fatigue-crack initiation at inclusions.


IV. STRESSINITIATION LIFE
CALCULATIONS
The applications of the proposed model (Eq. [20]) for
computing the stressinitiation life curves of a number of
structural alloys are demonstrated in this section. Material
constants in the model include the shear modulus ( ), Poissons ratio ( ), and grain size (d ). The term 2Mk represents
the fatigue limit below which fatigue-crack initiation does
not occur. These parameters and the fatigue limit were evaluated from available experimental data in the literature. The
Taylor factor for the optimally oriented grain is taken to be
2, and the parameter has a universal value of 0.005 based
on the investigation by Venkataraman et al.[9] The half-crack
length or crack depth (ci) at final failure is taken to be about
1000 m, unless an experimental value is available. The
slipband width (h) is generally not available, and it is used
as a fitting parameter in the stress-life calculations.
To determine the material constants, the fatigue limit was
first evaluated from the experimental data. A log-log plot
of - e vs Nf was then constructed to obtain the fatiguestrength exponent from the slope of the linear plot. This is
illustrated in Figure 3(a) for the PM Astroloy alloy.[29] After
specifying a final crack depth, the slipband width was evaluated by fitting the crack-initiation model to the experimental
data. Once h was determined, S-Ni curves were calculated
for various crack depths. The calculated curves of PM
Astroloy for four different ci values are shown in Figure 3(b).
A summary of the material constants for PM Astroloy[29] is
presented in Table I, together with those for other
alloys.[3039]
Figures 4(a) and (b) compare the calculated S-Ni curves
against experimental data for PM Astroloy[29] and wrought
IN100[30] at 600 C to 650 C. The experimental data corresponded to specimen fracture, which was simulated by crack
lengths of about 800 to 1000 m. For both alloys, the cycles
to crack initiation at a given stress range are reduced with
decreasing crack lengths. At high stresses, the cycles to
METALLURGICAL AND MATERIALS TRANSACTIONS A

(b)
Fig. 3Fatigue-life curves for PM Astroloy: (a) plot of -e used to
determine the material constants in the fatigue crack initiation model, and (b)
computed stress (S)initiation life (Ni) compared with experimental data.[29]

initiate a crack length of one-half of the grain size are comparatively short. Near the fatigue limit, most of the fatigue
life is spent in the initiation of a grain-sized crack.
The number of cycles to crack initiation is available for
a number of steels. Applications of the proposed fatiguecrack-initiation model to these steels are presented in Figure
5 for 4340[31] and three other low-carbon (LC) steels.[32] The
finite grain size for the 4340 steel is about 1 m,[4] while
it is 7.8, 20.5, and 55 m for LC steels A through C,
respectively. In all cases, the experimental data at failure
were used to determine the model constants. Using this set
of material constants, the fatigue-crack-initiation the model
was used to compute the S-Ni response for other crack sizes.
Figure 5(a) shows that the agreement between the calculated
curve and initiation data for 38 m is good at high stresses.
It is not as good, but reasonable, near the fatigue limit. There
are no experimental data available for crack sizes less than
38 m. Good agreement is also obtained for LC steels A
through C, as shown in Figures 5(b) through (d), respectively.
In all three cases, the crack lengths at crack initiation were
about 20 to 90 m.
The effect of grain size on fatigue-crack initiation is best
illustrated by the results shown in Figure 6(a) for Ti-6Al4V[33] and in Figure 6(b) for lamellar TiAl alloys.[34,35,36]
The Ti-6Al-4V alloy was heat treated to become a bimodal
VOLUME 34A, JANUARY 200347

Table I. Summary of Material Parameters for Various Structural Alloys

Materials
PM Astroloy[29]
IN100[30]
4340 steels[31]
LC steels A[32]
LC steels B[32]
LC steels C[32]
Ti-6Al-4V[33]
Lamellar TiAl[34,35,36]
X7075[37]
2024 Al[37]
X2024 Al[37]
ULTIMET[38]
Cu[39]

T, C

Grain Size
d, m

25
650
600
25
25
25
25
25
25, 650
25

20
20
20
1
7.8
20.5
55
6
150
30
220
50
50
80
25

25
25
25
25

R
0
0
*
1
1
1
1
0.1
0.1
1
1
1
1
0.05
1

Fatigue Limit
c 2 Mk,
MPa

Poissons
Ratio

500
580
600
600
476
340
300
335
310
100
100
110
70
437
75

0.3
0.3
0.3
0.3
0.3
0.3
0.3
0.3
0.3
0.3
0.3
0.3
0.3
0.3
0.3

Shear Modulus
, MPa

Fatigue Strength
Exponent,

Slipband Width
h, m

0.54
0.44
0.50
0.50
0.783
0.616
0.664
0.5
0.2
0.225
0.225
0.308
0.197
0.351
0.63

0.15
0.05
0.03
1.5 103
0.3
0.4
3
0.03
8 103
1.2 102
1.2 101
9 102
1.9 102
7.5 102
5

8.01
6.72
6.72
7.76
7.76
7.76
7.76
4.4
4.4
2.58
2.58
2.58
2.58
8.9
4.84

104
104
104
104
104
104
104
104
104
104
104
104
104
104
104

*Strain-controlled tests at zero minimum strain.

(a)

(b)
Fig. 4Comparisons of computed and measured S-Ni curves for Ni alloys
at 600 C to 650 C: (a) PM Astroloy[29] and (b) IN100.[30]

microstructure of primary grains and lamellar


colonies with an average grain size of 6 m. The S-Ni curve
is typical, exhibiting an increase in fatigue strength with a
48VOLUME 34A, JANUARY 2003

decrease in fatigue life, even though there is considerable


scatter near the fatigue limit. In contrast, the S-Ni curve for
lamellar TiAl alloys is relatively flat. The fatigue strength
is not very sensitive to the number of cycles to crack initiation or the fatigue life. The calculated S-Ni curve for the
grain-sized crack (ci 75 m) is similar to that for 500
m. Thus, fatigue failure in the lamellar alloys is imminent
once a grain-sized crack has been initiated.
Grain size affects the S-Ni curves of X-7075 Al[37] in a
different way. As illustrated in Figure 7(a), an increase in the
grain size lowers the fatigue strength of X7075-Al relatively
uniformly over the 104 to 107 cycles regime, by shifting the
entire S-Ni curve to a lower stress. This behavior is faithfully
reproduced by the crack-initiation model (Figure 7(a)) by
allowing the fatigue limit to vary with the grain size. Nonuniform slip has been shown to lower the fatigue strength in
2024-Al[37] (Figure 7(b)). The crack-initiation model is capable of simulating this aspect of the fatigue process through
the slipband-width parameter. Localized slip results in a
smaller slipband-width, while uniform slip would lead to a
large slipband width.
Application of the crack-initiation model to ULTIMET,[38]
a Co-based alloy, is presented in Figure 8(a). The grain size
of the alloy was 80 m, and the size of just-initiated slipband
cracks was on the order of 100 to 200 m. The calculated
S-Ni curves are in agreement with the experimental data
(Figure 8(a)). Comparison of the model calculation and
experimental data of Cu[39] is presented in Figure 8(b). The
model simulates the initiation of grain-sized microcracks
accurately, but the computed curve for specimen failure is
not as good. In particular, the model underpredicted the
fatigue limit.

V. CRACK-SIZE PREDICTIONS
One of the important features of the proposed model is
the ability to predict the crack size at fatigue-crack initiation. The crack length at initiation can be computed based
on two methods, utilizing either Eq. [18] or [20]. The use
METALLURGICAL AND MATERIALS TRANSACTIONS A

(a)

(b)

(c)

(d )

Fig. 5Measured S-Ni data compared against model calculations for steels at ambient temperature: (a) 4340 steel,[31] (b) LC steel A,[32] (c) LC steel B,[32]
and (d ) LC steel C.[32]

of Eq. [20] for computing the crack size is demonstrated


in Figures 3 through 8 for selected alloys. The validity of
Eq. [18] for computing the fatigue-crack size at initiation
is evaluated in this section. The evaluation involved a
review of the experimental data for fatigue-crack size at
initiation. A summary of the experimental data[31,32,38,4053]
for the crack size at initiation reported in the literature is
presented in Table II.
Figure 9 shows the measured values of 4340 steels[31]
and LC steels[32] as a function of grain size. Despite the
large scatter, the crack size is seen to increase with the
grain size, according to a power law with an exponent of
about 2. The theoretical crack size was computed based on
Eq. [18] using a slipband with of 0.4 m and a surface
energy of 2 J/m2. This particular h value was obtained from
the S-Ni calculations described earlier, while the s value
of 2 J/m2 is typical for cleavage fracture in metals. At a
given grain size, the computed crack sizes are five orders
of magnitude lower than those observed experimentally,
even though the correct grain-size dependence (slope of
the linear plot) was obtained. This discrepancy is attributed
to the use of a surface energy that does not reflect the
plastic energy dissipated during irreversible slip along the
slipband. To obtain good agreement between model calculations and measured crack sizes, the surface energy
METALLURGICAL AND MATERIALS TRANSACTIONS A

required is 1 105 J/m2, as shown in Figure 9. Thus, it is


imperative that the proper surface energy, one that includes
the plastic dissipation in the slipband, be used in the cracksize calculation when Eq. [18] is used. The surface energy
for cleavage fracture in metals would underpredict the crack
size at initiation considerably.
The fatigue-crack size detected at initiation for other
alloys is compared against the model calculations. From
Figure 10, it appears that experimental data form roughly
two scatter bands, corresponding to two sets of h and s
values. These correlations might be fortuitous, since it is
obvious from Eq. [18] that the crack size depends on h,
s , and the shear modulus. The results underscore the need
for more accurate values for h and s if Eq. [18] is used
to predict the fatigue-crack size at initiation.
The values of h are expected to range from the slipplane spacing to the grain size. The former corresponds
to localized slip on a single slip plane, while the latter
corresponds to uniform slip distributed evenly within the
entire grain. Computed h values for various alloys are correlated against the grain size in Figure 11, which shows that
the slipband widths lie between the two bounds. For a
given alloy, the h value appears to increase slightly with
increasing grain size.
VOLUME 34A, JANUARY 200349

(a)

(a)

(b)
Fig. 6Comparisons of computed and measured S-Ni curves for Ti-based
alloys: (a) Ti-6Al-4V[33] and (b) lamellar TiAl alloys.[34,35,36]

VI. SMALL-CRACK BEHAVIOR


The crack-size-based fatigue-crack-initiation model
allows one to investigate the growth behavior of just-initiated
small cracks. This can be accomplished by differentiating
Eq. [20] with respect to the fatigue cycle, and the resulting
fatigue-crack-growth rate (dc/dN ) expression is given by
dc
2
dN

(1 v)
8M 2

2Mk

d 3 21
c
h2

1
2

[22]
which is reduced to

(1 v)
dc

dN
8M 2

2Mk

d3
h2

[23]

and the crack-growth rate is independent of crack length


when 1/2. In contrast, dc/dN increases with crack length
when 1/2, but dc/dN decreases with crack length when
1/2.
The fatigue-crack-growth response of small cracks in Ti6Al-4V[54] was computed using Eq. [23] using the materials
constants obtained from S-Ni calculations and shown in
Table I. Equation [23] was used because 1/2 for this
50VOLUME 34A, JANUARY 2003

(b)
Fig. 7Calculated and observed S-Ni responses in Al alloys:[37] (a) grain
size effect in X-7075 Al and (b) slip morphology effect in 2024 Al.

alloy. The computed crack-growth rates as a function of


the applied stress for small cracks are presented in Figure
12, which shows that dc/dN is zero at a stress below the
fatigue limit, but increases rapidly with increasing stresses
above the fatigue limit. The crack-growth rate at a constant
stress and R ratio is, however, a constant, independent
of the crack length or the stress-intensity-factor range. A
comparison of the computed da/dN based on the crackinitiation model against experimental data of small
cracks[54] and large cracks[55] is shown in Figure 13 for
Ti-6Al-4V. Small cracks in Ti-6Al-4V were observed to
propagate at stress-intensity ranges below those of the
large-crack threshold, but they exhibited the same growth
rates as those of the large crack at K Kth when the
length of the small cracks became longer. In contrast, the
computed crack-growth rate based on the crack-initiation
model is a constant, independent of K. Furthermore, the
growth rate is in agreement with experimental data when
the crack size is greater than the grain size, which is about
6 m.
VI. CRACK INITIATION AT A STRESS
GRADIENT
The utilization of the crack-size-based initiation model
for treating fatigue-crack initiation at a stress gradient is
METALLURGICAL AND MATERIALS TRANSACTIONS A

as the maximum length of a fatigue crack that can be initiated


in the notch field. Figure 15 shows a comparison of Eq. [26a]
against experimental data of several structural alloys.[5867] A
summary of the notch data is presented in Table III. There
is reasonable agreement between the experimental data and
model calculations, which is consistent with the previous
finding by Lukas and Klesnil.[68] It is also important to note
that the stresses acting on a small crack emanating from a
notch change from the notch-stress field to the bulk-stress
field as the length of the small crack increases. The transition
point has been calculated as
l*
(k2t 1)1

[26b]

which is based on an analysis by Dowling[31] and is shown


as the dotted line in Figure 15. Fatigue is limited by crack
initiation at stresses below the boundary, but is also limited
by crack growth at stresses above the boundary. From Figure
15, it is apparent that the proposed fatigue-crack initiation
is most pertinent for fatigue of notches with a kt value of
less than 3 or 4, depending on the nominal stress.
For a crack of length c, where 0 c l*, initiated within
the notch-stress field, the average stress acting on the crack
is given by

(a)

(c)

1
c

(x)dx

1
c

ktS

4x

which can be integrated to yield


(c)

ktS
2c

and further simplified to


(b)

4c

1/2

dx

[27]

[28]

(c) ktS 1

Fig. 8Measured and calculated S-Ni curves for a Co alloy (ULTIMET)


and copper: (a) Co alloy, ULTIMET;[38] and (b) Cu.[39]

1/2

[29]

when the approximation


illustrated in this section for notch fatigue. Figure 14 shows
a notch of a root radius ( ) subjected to a remotely applied
stress range (S). The distance directly ahead of the tip of
the notch root is designated as x, while the distance normal to
the notch is designated as y. The elastic-stress-concentration
factor of the notch is kt . According to Neuber,[56] the distribution of the local stress (yy) of an element (x) located directly
ahead of the notch tip ( y 0) is given by[56,57]
yy ktS

4x

1/2

[24]

whose distribution is depicted by the solid line in Figure 14.


According to the proposed fatigue-crack-initiation model,
crack initiation is feasible only when e . There, the
maximum length of the fatigue crack that can be initiated
in a notch field is obtained from the equality given by

ktS
l*

1/2

[25]

which can be rearranged to yield

l* 1

ktS
e

METALLURGICAL AND MATERIALS TRANSACTIONS A

[26a]

4c

1/2

1 4c
1 4c

2
8

[30]

is invoked on the basis of the Binomial theorem. Substitution


of Eq. [29] into [20] leads one to

ktS 1

c
e N i

8M
(1 v)
2

1/2

[31]
1/2

dd
h

for crack initiation in a notch field, where e 2Mk is


the fatigue limit.
Figure 16(a) shows the calculated S-Ni curves for initiating
a crack depth of 200, 500, and 1000 m from a notch with
a kt value of of 4. The calculations are based on Eq. [31]
using material parameters for Ti-6Al-4V. Figure 16(a) indicates that at a given stress, the cycles to initiate a crack
increase with increasing crack length. At a large crack length
(i.e., 1000 m), the calculated S-Ni curve approaches that
of the smooth specimen. It occurs when the condition

kt 1

c
1

[32]

VOLUME 34A, JANUARY 200351

52VOLUME 34A, JANUARY 2003

METALLURGICAL AND MATERIALS TRANSACTIONS A

ferrite fine Nb
precipitates
bi-modal
bi-modal
globular (equiaxed
alpha)
globular (elongated
alpha)
2
equiaxed grains

Carbon steels

4320 steels
HSLA steels

single phase

single phase

Co superalloy

Mo/Mo alloys

IN718
IN100
PM LC
Astroloy

Astroloy

Lamellar TiAl
Ti-8.6Al

40

ferrite mastensite

LC steels (B)
LC steels (C)

Ti-6Al-4V
Ti-6Al-4V

20.5
55

ferrite pearlite
ferrite pearlite

LC steels (A)

60
100
30
80 (50 to 250)

5.0 to 50
60
30

75 to 150

Inclusions
(75 to 150 m
Al2O1)
inclusions
(75 to 150 m
SiO2)

slipbands

11
29
20
100 to 200

75 to 150

14 to 230
5 to 45
10 to 80
240
50 to 70
18
10 to 20
25

40 to 70

15
10 to 45
18 to 62

10.0 to 50.0
0.2

150
12.7
20 to 70
100 to 140
90
3.1
118
377
104
12 to 250

2
50 to 70

Initial Crack
Length,
2ci , m

slipbands
slipbands
slipbands
pores
slipbands
carbides
carbides
slipbands

elongated alpha

38
150
20
100
30

primary alpha
alpha
alpha

slipbands in
ferrite
gb
slipbands

slipbands
slipbands

slipbands
grain boundaries
(gb)
slipband
gb
gb/slipbands
gb/slipbands
slipbands

Initiation Site

6
10
5

5.0 to 23.0
7

100
130
600
7.8

single phase
single phase
single phase
ferrite pearlite

Pure Al
Fe-3 pct Si

10
350

single phase
single phase

Microstructure

Grain Size, d
(m)

SEM

SEM

SEM
SEM
SEM

SEM
OM
OM
SEM

SEM

SEM
SEM
SEM

SEM

OM

OM
OM
OM
Ni Nf Np

OM

Measurement
Method

320
96

340
650
650
860

750

720
750
750

450
385
375
586

331
240
265
245
245
245
184
279

Stress
Amplitude,
MPa

310

680

680
680
680

420
360
350
450

270
225
232
232
187
158
158
279

Fatigue
Limit,
MPa
5

Reference

[45]

Kunio and Yamada[44]

Taira et al.[32]

Smith[42]
Tanaka et al.[43]

McGreevy and Socie


Bhat et al.[53]

3 105
6 105
2 106
1.4 106
5 106
6.6 105
3.9 105
3.5 106

2 10 Thompson et al.[13,42]
3 104 Usami[40]

Nf ,
Cycles

1 107
4.8 107
2 107

Weiss et al.[51]
Wei et al.[30]
Jablonski[52]

Davidson and Chan[48,49,50]

Chan and Shih[34]


Wagner et al.[41]

Weiss et al.[51]

1 106 Jiang et al.[38]

6.9 104
3.5 103

1.7 103

8.4 104
8 104

1 103 5.1 103

Peters et al.[46]
1 103 1.3 104 Demulsant and Mendez[47]
1 103 9.1 103

9 103

10 104
1 105
6 105
3 105
1.1 106
2.2 105
1.3 105
9.0 105
1 107

1 10
5 103

Ni ,
Cycles

Summary of Microstructure, Grain Size, Fatigue Crack Initiation Site, Initial Crack Length, Measurement Method, and Fatigue Conditions for Various Alloys

Pure Cu

Material

Table II.

Fig. 11Correlation of deduced values of the slipband width with grain


size for various structural alloys.

Fig. 9A comparison of measured and computed crack sizes at crack


initiation for steels. The crack lengths computed based on the surface energy
(s 2 J/m2) is several orders of magnitude smaller than the experimental
values. A large surface energy on the order 1 105 J/m2 is required to
give the experimentally observed crack lengths. The finding suggests that
the plastic energy dissipated by irreversible slip during the crack initiation
process must be taken into account in the surface energy term.

Fig. 12Fatigue crack growth rates of just initiated small cracks as a


function of the applied stress ranges for Ti-6Al-4V.

Fig. 10Experimental data of crack lengths at initiation for various structural alloys compared against model calculations.

is satisfied. Thus, it is important to apply Eq. [31] judiciously


to the length scale where fatigue-crack initiation dominates.
For kt 4, fatigue-crack initiation dominates at c/ 0.067
(Figure 15), corresponding to c 100 to 200 m. Thus,
the calculated S-Ni curve shown in Figure 16 is valid for c
200 m. Calculated S-Ni curves for large crack depths
(200 m) might be inaccurate, since crack growth has not
been considered in these calculations. A comparison of the
calculated and measured S-Ni curves for smooth[33] and
notched[59,60] specimens of Ti-6Al-4V is presented in Figure 16(b).
VII. DISCUSSION
A unique contribution of this article is the development
of a microstructure-based fatigue-crack-initiation model that
METALLURGICAL AND MATERIALS TRANSACTIONS A

relates explicitly the stress range and fatigue cycle to the


crack-size and microstructural-size parameters. The energetics of fatigue-crack formation have been clarified and firmly
established on the basis of the Orowan modification of the
Griffith theory of brittle fracture. Such a relation is important
for providing a scientific basis for predicting the crack size
at fatigue-crack initiation within a surface grain as well as
at a notch root. Furthermore, the proposed model allows one
to predict the crack-extension rate for an incipient small
crack at initiation. The crack-size-based model, therefore,
provides a link in the gap between the crack-initiation and
crack-propagation approaches that has, heretofore, been
missing.
One of the important findings of the survey performed as
part of this investigation is that the crack size at initiation,
reported in the literature, is usually on the order of the grain
size or greater. Cracks of smaller size have been reported.
Even in these cases, the observed crack sizes are considerably larger than those predicted on the basis of Eq. [18] and
a surface energy of 2 J/m2, which is a reasonable value for
the energy to create two new surfaces in metals. The large
discrepancy between the computed and observed crack sizes
shown in Figures 9 and 10 necessitates a broader view of
what contributes to the surface-energy term in Eq. [18].
VOLUME 34A, JANUARY 200353

Fig. 15Calculated crack depth at initiation normalized by the notch radius


compared against experimental data for several structural alloys.[5869]

Table III. Summary of Notch Stress Concentration Factor


(kt), Notch Radius ( ), and Crack Depth (c) for
Selected Alloys
Material
Ti-6Al-4V
(STOA)[58]
Ti-6Al-4V (MA)[59,60]
Fig. 13A comparison of the calculated fatigue crack growth response of
just initiated small cracks against experimental data of small and large
cracks in Ti-6Al-4V.[54,55]

Ti-6Al-4V (BA)[59,60]
Ti-10V-2Fe-3Al[61]
2024-T4[62]
2124-T4[62]
7075-T6[63,64]
Mild steels[65,66]
LC steels[67]

Fig. 14A schematic depicts the formation of a fatigue crack within the
stress field of a notch subjected to a remotely applied stress range of S.

The concept of a surface energy for fatigue-crack formation is a direct extension of Orowans modification of the
Griffith energy for brittle fracture.[69] As in quasi-static fracture,[70,71] the surface energy dissipated by fatigue is considerably larger than that required to create two fresh surfaces.
As a result, it is not surprising that the apparent surface
energy required to nucleate a fatigue crack of a size commensurate to the experimental data is much larger than the typical
value of 2 J/m2 for the surface energy. It is also encouraging
that the apparent surface energy inferred from Eq. [18] and
54VOLUME 34A, JANUARY 2003

kt

Notch Radius,
mm

2.25
4.1
12.6
4.1
12.6
2
2
2
8
12
8.48

1.27
1.59
0.25
1.59
0.25
2
2
2
0.1
0.254
0.16

Crack Depth,
m
16
150
950
265
635
10
64

to 370
to 2540
to 5080
to 5080
to 2800
to 20
to 500
76
550
800
167 to 854

the experimentally observed fatigue-crack sizes at initiation


are comparable to the experimental values of the plastic
work dissipated during fatigue-crack propagation. Previous
studies by Fine, Davidson, and co-workers[7276] indicated
that the plastic work dissipated during fatigue-crack growth
was in the range of 104 to 106 J/m2 for various alloys including steels, Al alloys, and Ti alloys. During fatigue-crack
initiation, most of the energy provided by the applied load
is expended in irreversible slip and dissipated as heat, while
that stored in the dislocation dipoles is comparably small.
Thus, the surface-energy term in Eq. [18] should be identified
with the plastic work dissipated by irreversible slip during
the crack-initiation process and not with the energy to create
free surfaces.[77] In the event that fatigue-crack initiation
indeed occurs at the true surface energy, the crack size at
initiation would be so small (0.1 m) that it is certainly
below the detection limit of most conventional crack-length
measurement methods. Subsequent growth of these fatigues
to a detectable size would invariably involve a surface energy
on the order of the 104 to 106 J/m2 reported for the growth
of fatigue cracks in a number of structural alloys.[7276] Thus,
there is both theoretical and experimental evidence to support the notion that the s term in Eq. [18] is dominated
by the plastic work spent in irreversible slip during crack
formation. Most, if not all, of this plastic work is likely to
be dissipated as heat.
METALLURGICAL AND MATERIALS TRANSACTIONS A

Three length-scale parameters are included in the crackinitiation model, which are the slipband width, grain or
microstructural size, and the crack size. A simple and yet
rigorous way for extending the microstructure-based crackinitiation model to a probabilistic design and life-prediction
framework is to describe the distributions of the three lengthscale parameters (h, d, and c) in terms of the means (h, d,
and c) and randomness (Xh , Xd , and Xc).[5] The latter are
dimensionless parameters that describe the randomness or
statistical distribution of the parameter of interest. Upon
proper substitution, Eq. [20] can be expressed as
( 2Mk)N i
(a)

(b)
Fig. 16S-Ni curves for notch fatigue: (a) computed S-Ni curves for initiation to several crack sizes at a notch tip compared to that for smooth bar
fatigue, and (b) computed S-Ni curves compared against experimental data
for Ti-6Al-4V.[33,59]

The difficulty with identifying the proper value for the


surface-energy term can be avoided if Eq. [20] is used for
relating the cycles to initiation at a given crack length or
depth. This approach also offers the advantage that it can
be readily adopted into a life-prediction scheme for treating
both fatigue-crack initiation and growth, as demonstrated
earlier by Dowling[31] and Socie et al.[78] Instead of experimental estimates of crack length, as used by Dowling,[31]
the crack-size-based crack-initiation model, if verified by
further experimental means, allows one to compute the number of cycles to initiation for the desired crack length, as
well as the uncertainty in fatigue-crack-initiation life due to
variations in microstructural features or defect sizes.
The formulation of the crack-initiation model is based on
the accumulation of dislocation dipoles on planar slip bands.
This type of crack-initiation process is most prevalent in
planar-slip materials, where the formation of dislocation
cells and patterning is difficult. For cell-forming materials,
the proposed crack-initiation model might be applicable once
PSBs are formed. Under this circumstance, the ladder-like
dislocation structure within the PSBs is similar to the dislocation pileup structure in planar-slip materials. Coalescence
of edge-dislocation dipoles in the PSBs has also been proposed as the mechanism of void formation in cell-forming
materials.[79,80]
METALLURGICAL AND MATERIALS TRANSACTIONS A

8M 2 2
(1 v)
1/2

1/2

1/2


h c
d d

[33]


Xh Xc
Xd Xd

to provide a probabilistic relation between fatigue-crackinitiation life and variations in microstructural-length scales.
If desired, the fatigue limit (2Mk) and the Taylor factor can
be described via the probabilistic treatment. Furthermore,
the same approach can be applied to Eqs. [21] and [29] to
obtain the pertinent equations for crack initiation at inclusions and notches, respectively.
The proposed crack-initiation model has been evaluated
against experimental data in the literature. Since not all
the microstructural-length parameters in the crack-initiation
model were measured, the values of the parameters such as
slipband width and the crack size at initiation or fracture
could only be estimated or, in some cases, fitted to the
experimental data of crack-initiation life. Even though it is
not a rigorous proof of the crack-initiation model, the
deduced slipband width and crack size at initiation or fracture
are reasonable and consistent with experimental data when
they are available in the literature.
According to the crack-initiation models as given in Eqs.
[20], [21], and [29], the number of cycles to crack initiation
increases with increasing crack length and decreasing grain
size. Both findings are supported by extensive experimental
observations reported in the literature.[31,32,37] In addition,
the number of cycles to initiation increases with increasing
slipband width. The slipband width parameter can be interpreted as a measure of the inhomogeneity of slip within the
grain. A small slipband width accompanies inhomogeneous,
localized slip, and it leads to easy crack initiation and a
lower crack-initiation life. In contrast, homogeneous slip
results in a large slipband width and a higher crack-initiation
life. The detrimental effect of localized slip and the beneficial
effect of uniform slip on fatigue-crack initiation are well
known and have been demonstrated by Starke and Lutjering.[37] From Figure 11, it is evident that the slipband width
is related to the grain size. The origin and nature of the
relationship between the slipband width and grain size is
not known. Undoubtedly, the widening of the slipband must
be related to the work-hardening behavior of the material
within the slipband. Work hardening of the slipband is not
addressed in the current model, but it can be incorporated
using crystal-plasticity theories. In the current investigation,
the slipband width is used, essentially, as a fitting parameter.
Experimental determination of this material parameter is
also desirable. It is also of interest to note that the slipband
VOLUME 34A, JANUARY 200355

Fig. 17Crack length vs fatigue cycle curves for three grain sizes computed
using the proposed crack initiation model compared with the current
approach for assigning a specified crack length after the low-cycle fatigue
life has been exhausted. The calculations were performed for crack sizes
below the crack detection limit of current nondestructive evaluation
methods.

values obtained in this study are consistent with the widths


of PSBs or extrusions reported in the literature.[6,1517,81]
For lifing of components, it is a common practice to
compute a crack-initiation life based on low-cycle-fatigue
data generated by testing smooth specimens and a crackgrowth life based on fatigue-crack-growth data obtained
from large-crack fracture-mechanics specimens.[3] The
crack-initiation life is defined in terms of a specified crack
size that corresponds to the resolution of the crack-detection
method. For a complex geometry, the crack-initiation life
corresponds to the cycles to initiate a surface crack of 750
m in length.[3,82] The current approach is illustrated as the
dashed line in Figure 17. The proposed crack-initiation
model provides a framework for providing several pieces of
information that cannot be obtained by the current approach,
which include: (1) computing the initiation portion of the
crack length vs fatigue-cycle curve, (2) delineating the crackinitiation and the crack-growth regimes, as well as their
transition point, and (3) calculating the variation of the crackinitiation response due to microstructure variability. This is
demonstrated in Figure 17, which shows the calculated crack
length as a function of fatigue cycles in the crack-initiation
regime for Ti-6Al-4V with three grain sizes. In Figure 17,
an existing crack of 750 m in length is shown as the dotted
line. The current approach of assigning a crack length of
750 m after the low-cycle-fatigue life has been exhausted
is represented by a step function and is shown by the dashed
line in Figure 17. In comparison, the computed crack length
cycles curves are continuous, sensitive to the grain size, and
different from that depicted by the step function at small
crack sizes that are below the current crack-detection limit.
The enhanced ability of being able to compute the crack
size at initiation for a crack size below the crack-detection
limit is a unique attribute of the proposed crack-initiation
model.

1. The crack size at initiation can be incorporated into a


fatigue-crack-initiation model by considering the fracture
energy dissipated by the irreversible slip process during
fatigue. A set of crack-size-based fatigue-crack-initiation
models has been developed for treating crack initiation
at slipbands, inclusions, and notches.
2. The number of cycles to initiation increases with increasing crack sizes and slipband widths, but decreases with
increasing grain sizes, inclusion sizes, and notch-stressconcentration factors (decreasing notch radius).
3. The crack sizes at initiation, calculated based on the
surface energy, are several orders of magnitude smaller
than the experimental values. For realistic crack sizes,
the surface-energy term must be interpreted as the plastic
energy dissipated by irreversible slip during the fatiguecrack-initiation process.
4. The crack-initiation model leads to a crack-growth-rate
equation for stage I fatigue that shows different characteristics than those of stage II large cracks.
5. Crack initiation at blunt notches (kt 4) is treated by the
proposed model. The applicability of the crack-initiation
model to sharp notches (kt 4) might be limited, because
fatigue-crack growth is the controlling failure process in
sharp notches.
6. Capabilities of the crack-size-based initiation model
include computations of the stressinitiation life curves,
fatigue-crack growth curve for just-initiated cracks,
crack-length vs fatigue-cycle curve for crack sizes below
the current crack-detection limit, and the variability of
these properties due to microstructure variations.
APPENDIX
The energetics of fatigue fracture
According to Rice,[83,84] irreversible slip in metals can be
related to internal dislocation arrangements and described
in terms of changes in the internal variables. These internal
variables could conceptually be considered as state variables
that provide a full thermodynamic characterization of constrained equilibrium states corresponding to a given internal
dislocation arrangement, applied stress, and temperature. For
crack initiation from a pileup of dislocation dipoles, the
internal variable would be the elastic-strain energy stored
in the dislocations. The free-energy change associated with
the coalescence of dislocation dipoles into a fatigue crack
is the difference in the energy (2cs) expended in forming
a crack and the elastic-strain energy released by the dislocations and crack opening, resulting in[22,23]
G We Wm 2cs

[A1]

which is also shown as Eq. [11] in the text. The We term


corresponds to the stored-energy, which is comprised of the
self-energy and the interaction energy of the dislocation
dipoles. The value of We is given by
We

1bd 2 ( 2k)2 N 2
4 2A

[A2]

with
A b/2 (1 v)

VIII. CONCLUSIONS
The conclusions reached in this investigation are as
follows.
56VOLUME 34A, JANUARY 2003

where 1 is a numerical constant. The Wm term is Irwins


strain-energy release rate due to the crack opening[71] and
is given by[22,23]
METALLURGICAL AND MATERIALS TRANSACTIONS A

Wm

1v 2
(KI K2II)dc

[A3]

Ni

where KI and KII are the Mode I and II stress-intensity


factors, respectively. Integrating [A3] lead one to
Wm

(1 v)( 2k) d b N
8 A2
2

2 2 2

[A4]

bd( 2k)N
A

[A5]

which is combined with Eqs. [A1], [A2], and [A4] to give


G z1N 2 z2N 2 2z3Ns

[A6]

where

1bd 2 ( 2k)2
4 2 A

[A7]

(1 v)( 2k)2 2d 2b2


8A2

[A8]

bd ( 2k)
A

[A9]

z1
z2
and

z3

The partial derivation of G with respect to N gives


G
2z1N 2z2N 2z3 s
N
and setting G/N 0 leads one to
Ni

z3 s
z 1 z2

[A10]

when G/N 0 is taken as the condition for crack initiation. The crack size at initiation is then obtained by substituting Eq. [A10] into [A5], leading to
ci

z3
bd( 2k)
s
A
z 1 z2

[A11]

for initiation in which all the dislocation dipoles in the pileup


contribute to the formation of the crack. A simpler cracksize expression has been obtained and is shown in Eq. [18]
when only a portion of the dislocation pileup contributes to
crack initiation. For both cases, ci is proportional to the
surface energy, s . The surface energy is considered to consist of an elastic component and a plastic component. The
latter reflects the plastic work spent by the irreversible slip
that accompanies the crack formation and the opening up
of the crack surfaces.
Instead of using G/N 0 as the crack-initiation criterion, the onset of crack initiation can be considered to occur
at G 0. Using Eq. [A6], this crack-initiation criterion
leads one to
G z1N z2N 2z3 s 0
METALLURGICAL AND MATERIALS TRANSACTIONS A

2z3 s
z 1 z2

[A13]

which is twice as large as that shown in Eq. [A10].

as the elastic-strain-energy release rate associated with the


opening up of the incipient fatigue crack. Furthermore, the
incipient crack length can be related to the dislocation pileup,
as given by[23]
c

and

[A12]

ACKNOWLEDGMENTS
This research was performed while the author was an
ERLE Visiting Scientist at Air Force Research Laboratory
(AFRL). The work was supported by AFRL through Contract No. F33615-99-C-5803, Dr. James M. Larsen, Technical Manager. Technical assistance by Mr. J. Lawson, M.
Dent, and J. Jira, AFRL, Dr. V. Nagarajan of Systran Federal
Corporation, and clerical assistance by Ms. L. Salas, Southwest Research Institute, is acknowledged.
REFERENCES
1. K. Tanaka and T. Mura: ASME J. Appl. Mech., 1981, vol. 48, pp.
97-103.
2. K. Tanaka and T. Mura: Metall. Trans. A, 1982, vol. 13A, pp. 117-23.
3. B.A. Cowles: Mater. Sci. Eng., 1988, vol. A103, pp. 63-69.
4. J.-P. Bailon and S.D. Antolovich: in Fatigue Mechanisms: Advances
in Quantitative Measurements of Physical Damage, ASTM STP 811,
J. Lankford, D.L. Davidson, W.L. Morris, and R.P. Wei, eds., ASTM,
Philadelphia, PA, 1983, pp. 313-49.
5. K.S. Chan: Metall. Trans. A, 1993, vol. 24A, pp. 2473-86.
6. K.S. Chan and T.-Y. Torng: ASME Trans., J. Eng. Mater. Technol.,
1996, vol. 118, pp. 379-86.
7. R. Tyron and T.A. Cruse: ASME J. Eng. Mater. Technol., 1997, vol.
119, pp. 65-70.
8. G. Venkataraman, Y.W. Chung, and T. Mura: Acta Metall. Mater.,
1991, vol. 39, pp. 2621-29.
9. G. Venkataraman, Y.W. Chung, and T. Mura: Acta Metall. Mater.,
1991, vol. 39, pp. 2631-38.
10. M.R. Mitchell: in Fatigue and Microstructure, M. Meshii, ed., ASM,
Metals Park, OH, 1978, pp. 385-437.
11. L.F. Coffin, Jr.: Trans. ASME, 1954, vol. 76, pp. 931-50.
12. S.S. Manson and M.H. Hirschberg: in Fatigue: An Inter-Disciplinary
Approach, Syracuse University, Syracuse, NY, 1964, pp. 133-78.
13. N. Thompson, N.J. Wadsworth, and N. Louat: Phil. Mag., 1956, vol.
1, pp. 113-26.
14. J.C. Grosskreutz: in Metal Fatigue DamageMechanism, Detection,
Avoidance, and Repair, ASTM STP 495, S.S. Manson, ed., ASTM,
Philadelphia, PA, 1971, pp. 5-60.
15. C. Laird: in Fatigue and Microstructure, M. Meshii, ed., ASM, Metals
Park, OH, 1978, pp. 149-203.
16. M.E. Fine and R.O. Ritchie: in Fatigue and Microstructure, M. Meshii,
ed., ASM, Metals Park, OH, 1978, pp. 245-78.
17. L. Remy: in Fatigue 84, C.J. Beevers, ed., EMAS, Warley, United
Kingdom, 1984, vol. I, pp. 15-30.
18. A.S. Cheng and C. Laird: Fat. Fract. Eng. Mater. Struct., 1981, vol.
4, pp. 343-53.
19. A. Saxena and S.D. Antolovich: Metall. Trans. A, 1975, vol. 6A, pp.
1809-28.
20. K. Tanaka and T. Mura: Mech. Mater., 1981, vol. 1, pp. 63-73.
21. M.R. Lin, M.E. Fine, and T. Mura: Acta Metall., 1986, vol. 34, pp.
619-28.
22. G. Venkataraman, Y.-W. Chung, Y. Nakasone, and T. Mura: Acta
Metall. Mater., 1990, vol. 38, pp. 31-40.
23. T. Mura and Y. Nakasone: J. Appl. Mech., 1990, vol. 57, pp. 1-6.
24. T. Mura: Mater. Sci. Eng., 1994, vol. A176, pp. 61-70.
25. K.S. Chan: Scripta Metall. Mater., 1995, vol. 32 (2), pp. 235-40.
26. R. Chang, W.L. Morris, and O. Buck: Scripta Metall., 1979, vol. 13,
pp. 191-94.
27. C. Ihara and T. Tanaka: Fat. Fract. Eng. Mater. Struct., 2000, vol.
23, pp. 375-80.
28. S.E. Harvey, P.G. Marsh, and W.W. Gerberich: Acta Metall. Mater.,
1994, vol. 42, pp. 3493-3502.
29. Isomoto and N. Stoloff : Mater. Sci., 1990, vol. A124, pp. 171-81.
30. W. Wei, H. Floge, and E.E. Affeldt: Scripta Metall. Mater., 1991, vol.
VOLUME 34A, JANUARY 200357

25, pp. 1757-61.


31. N. Dowling: Fat. Eng. Mater. Struct., 1979, vol. 2, pp. 129-38.
32. S. Taira, K. Tanaka, and M. Hoshina: Fatigue Mechanisms, ASTM
STP 675, ASTM, Philadelphia, PA, 1979, pp. 135-73.
33. R.S. Bellows: Improved High Cycle Fatigue Life Prediction, Final
Report No. F33615-96-C-5269, University of Dayton Research Institute, Dayton, OH, 2001.
34. K.S. Chan and D.S. Shih: Metall. Mater. Trans. A, 1997, vol. 28A,
pp. 79-90.
35. K.S. Chan and D.S. Shih: Metall. Mater. Trans. A, 1998, vol. 29A,
pp. 73-87.
36. J.M. Larsen: Air Force Research Laboratory, Wright-Patterson Air
Force Base, Dayton, OH, unpublished research, 2001.
37. E.A. Starke, Jr. and G. Lutjering: in Fatigue and Microstructure, M.
Meshii, ed., ASM, Metals Park, OH, 1978, pp. 205-43.
38. L. Jiang, C.R. Brooks, P.K. Liaw, and D.L. Klarstrom: in Superalloys
2000, T.M. Pollock, R.D. Kissinger, R.R. Bowan, K.A. Green, M.
McLean, S. Olson, and J.J. Schirra, eds., TMS, Warrendale, PA, 2000,
pp. 583-91.
39. W. Hessler, H. Mullner, B. Weiss, and R. Sticker: Met. Sci., 1981,
vol. 15, pp. 235-40.
40. S. Usami: in Small Fatigue Crack, R.O. Ritchie and J. Lankford, eds.,
TMS, Warrendale, PA, 1986, pp. 559-83.
41. L. Wagner, J.K. Gregory, A. Gysler, and G. Lutjering: in Small Fatigue
Cracks, R.O. Ritchie and J. Lankford, eds., TMS, Warrendale, PA,
1986, pp. 117-28.
42. G.C. Smith: Proc. R. Soc., 1957, vol. A242, pp. 189-97.
43. K. Tanaka, M. Hojo, and Y. Nakai: in Fatigue Mechanisms: Advances
in Quantitative Measurement of Physical Damage, ASTM STP811,
J. Lankford, D.L. Davidson, W.L. Morris, and R.P. Wei, eds., ASTM,
Philadelphia, PA, 1983, pp. 207-32.
44. T. Kunio and K. Yamada: in Fatigue Mechanisms, ASTM STP675,
J.T. Fong, ed., ASTM, Philadelphia, PA, 1979, pp. 342-70.
45. T.E. McGreevy and D.F. Socie: Fat. Fract. Eng. Mater. Struct., 1999,
vol. 22.
46. M. Peters, A. Gysler, and G. Luetjering: in Titanium 80 Science and
Technology, H. Kimura and O. Izumi, eds., AIME, Warrendale, PA,
1980, vol. 3, pp. 1777-85.
47. X. Demulsant and J. Mendez: Fat. Fract. Eng. Mater. Struct., 1995,
vol. 18, pp. 1483-97.
48. D.L. Davidson and K.S. Chan: Acta Metall., 1989, vol. 37 (4), pp.
1089-97.
49. D.L. Davidson and S.J. Hudak, Jr.: Metall. Mater. Trans. A, 1995, vol.
26A, pp. 2247-57.
50. D.L. Davidson: Fat. Fract. Eng. Mater. Struct., 2000, vol. 23, pp.
445-52.
51. B. Weiss, R. Stickler, and A. Fathulla: in Short Fatigue Cracks,
R.O. Ritchie and J. Lankford, eds., TMS, Warrendale, PA, 1986, pp.
471-97.
52. D.A. Jablonski: Mater. Sci. Eng., 1981, vol. 48, pp. 189-98.
53. S.P. Bhat, R.S. Cline, and Y.-W. Chung: M.E. Fine Symp., P.K. Liaw,
J.R. Weertman, H.L. Marcus, and J.S. Santner, eds., TMS, Warrendale,
PA, 1991, pp. 49-59.
54. Y.N. Lenets: Improved High Cycle Fatigue Life Prediction, Final
Report No. F33615-96-C-5269, University of Dayton Research Institute, Dayton, OH, 2001, Appendix 3C.
55. R.E. deLaneuville and J.W. Sheldon: Improved High Cycle Fatigue
Life Prediction, Final Report No. F33615-96-C-5269, University of
Dayton Research Institute, Dayton, OH, 2001, Appendix 3B.
56. H. Neuber: Theory of Notch Stress, translated by F.A. Raven, Edwards,
Ann Arbor, MI, 1946.

58VOLUME 34A, JANUARY 2003

57. M.M. Hammouda, R.A. Smith, and K.J. Miller: Fat. Eng. Mater.
Struct., 1979, vol. 2, pp. 139-54.
58. M.A. Moshier, T. Nicholas, and B.M. Hillberry: Fatigue and Fracture
Mechanics, ASTM STP1417, W.G. Reuter and R.S. Piascik, eds.,
ASTM, West Conshohocken, PA, 2002, vol. 33, in press.
59. G.R. Yoder, L.A. Cooley, and T.W. Crooker: Fracture Mechanics:
16th Symposium, ASTM STP868, M.F. Kanninen and A.T. Hopper,
eds., ASTM, Philadelphia, PA, 1985, pp. 392-405.
60. G.R. Yoder, L.A. Cooley, and T.W. Crooker: Proc. 23rd Structures,
Structural Dynamics and Materials Conf., CP823, Part 1, American
Institute of Aeronautics and Astronautics, New York, NY, 1982, pp.
132-36.
61. G.R. Yoder, L.A. Cooley, and R.R. Boyer: in Microstructure, Fracture
Toughness and Fatigue Crack Growth Rate in Titanium Alloys, A.K.
Chakrabarti and J.C. Chesnutt, eds., TMS, Warrendale, PA, 1987, pp.
209-29.
62. S.S. Manson: Exper. Mech., 1965, vol. 5, pp. 193-226.
63. J.C. Grosskreutz and G.G. Shaw: Technical Report No. 66-96, Air
Force Materials Laboratory, Wright-Patterson Air Force Base, Dayton,
OH, May 1966.
64. S.S. Manson and M.H. Hirschberg: Technical Note D-3146, NASA,
Cleveland, OH, June 1967.
65. N.E. Frost and D.S. Dugdale: J. Mech. Phys. Solids, 1957, vol. 5, pp.
182-92.
66. M.H. El Haddad, T.H. Hopper, and K.N. Smith: Eng. Fract. Mech.,
1979, vol. 11, pp. 573-84.
67. K. Tanaka and Y. Nakai: Fat. Eng. Mater. Struct., 1983, vol. 6, pp.
315-27.
68. P. Lukas and M. Klesnil: Mater. Sci. Eng., 1978, vol. 34, pp.
61-68.
69. A.A. Griffith: Phil. Trans. R. Soc. A, 1921, vol. 22A, pp. 163-98.
70. E. Orowan: Rep. Progr. Phys., 1948, vol. XII, p. 185.
71. G.R. Irwin: Fracturing of Metals, ASM, Cleveland, OH, 1948, pp.
147-66.
72. P.K. Liaw, M.E. Fine, and D.L. Davidson: Fat. Eng. Mater. Struct.,
1980, vol. 3, pp. 59-74.
73. P.K. Liaw, S.I. Kwun, and M.E. Fine: Metall. Trans. A, 1981, vol.
12A, pp. 49-55.
74. M.E. Fine and D.L. Davidson: in Fatigue Mechanisms, J. Lankford
et al., eds., ASTM STP811, Philadelphia, PA, 1983, pp. 350-70.
75. D.L. Davidson and J. Lankford: in Environment-Sensitive Fracture of
Engineering Materials, Z.A. Foroulis, ed., TMS-AIME, Warrendale,
PA, 1977, pp. 581-94.
76. D.L. Davidson: Morris E. Fine Symp., P.K. Liaw, J.R. Weertman,
H.L. Marcus, and J.S. Santner, eds., TMS, Warrendale, PA, 1991, pp.
355-62.
77. S.R. Bodner, D.L. Davidson, and R.J. Lanford: Eng. Fract. Mech.,
1983, vol. 17, pp. 189-91.
78. D.F. Socie, N.E. Dowling, and P. Kurath: Fracture Mechanics: 15th
Symp., ASTM STP833, R.I. Sanford, ed., ASTM, Philadelphia, PA,
1984, pp. 284-99.
79. J.G. Antonopoulus, L.M. Brown, and A.T. Winter: Phil. Mag., 1976,
vol. 34, pp. 549-63.
80. U. Essman and H. Mughrabi: Phil. Mag., 1979, vol. 40, pp. 731-56.
81. P.J.E. Forsyth: Proj. R. Soc., 1957, vol. A242, pp. 198-202.
82. Y. Dai, N. Marchand, and M. Hongoh: Can. Aero. Space J., 1993,
vol. 39, pp. 35-44.
83. J. Kestin and J.R. Rice: Proc. Critical Review of Thermodynamics
Symposium, E.B. Stuart, G.-O. Benjamin, and A.J. Brainard, eds.,
University of Pittsburgh School of Engineering Publication, Pittsburgh,
PA, 1970, pp. 275-98.
84. J.R. Rice: J. Appl. Mech., 1970, vol. 37, pp. 728-37.

METALLURGICAL AND MATERIALS TRANSACTIONS A

Você também pode gostar