You are on page 1of 9

Fatigue Crack Initiation and Propagation of Binder-Treated

Powder Metallurgy Steels


S.J. POLASIK, J.J. WILLIAMS, and N. CHAWLA

Many of the targeted applications for powder-metallurgy materials, particularly in the automotive
industry, undergo cyclic loading. It is, therefore, essential to examine the fatigue mechanisms in these
materials. The mechanisms of fatigue-crack initiation and propagation in ferrous powder-metallurgy
components have been investigated. The fatigue mechanisms are controlled primarily by the inherent
porosity present in these materials. Since most, if not all, fatigue cracks initiate and propagate at the
specimen surface, surface replication was used to determine the role of surface porosity in relation
to fatigue behavior. Surface replication provides detailed information on both initiation sites and on
the propagation path of fatigue cracks. The effect of microstructural features such as pore size and
pore shape, as well as the heterogeneous microstructure on crack deflection, was examined and is
discussed. Fracture surfaces were examined to elucidate a mechanistic understanding of fatigue
processes in these materials.

I. INTRODUCTION from enhanced bonding of particles. Additionally, there is


a reduction of fine-particle dusting, resulting in more effi-
SINTERED powder-metallurgy components are in- cient use of alloy powders.
creasingly replacing wrought alloys in many applications,
Sintered ferrous materials are typically characterized by
due to their high performance, low cost, and ability to be
processed to near-net shapes. a porous and heterogeneous microstructure which develops
The main thrust toward higher performance in powder- from incomplete diffusion of alloying elements during sin-
metallurgy alloys has been achieved by introducing alloying tering. Due to the incomplete diffusion of alloying elements,
additions such as Mo, Mn, Cu, and Ni. While alloying addi- multiple phases are formed. The detrimental effect of poros-
tions increase the strength of the Fe alloy, in elemental ity on the mechanical properties of powder-metallurgy com-
form they may oxidize or diffuse inhomogeneously into ponents is fairly well known.[4–10] Under monotonic tensile
the surrounding Fe particles. Introduction of the alloying loading, porosity reduces the effective load-bearing cross-
additions during melting atomization to form prealloyed Fe sectional area and acts as a stress-concentration site for
particles is effective, but significantly decreases the com- damage, decreasing both strength and ductility.[7] One may
pressibility of the powders. In conventional powder-metal- characterize the porosity in these materials as either intercon-
lurgy processing, “diffusion-alloyed” powders have typically nected or isolated in nature. Isolated porosity results in more
been used. This process involves bonding iron and alloying homogeneous deformation, while interconnected porosity
particles through an intermediate annealing step which causes an increase in the localization of strain at relatively
allows partial diffusion and sintering of the particles to take smaller sintered regions between particles. Thus, for a given
place, prior to pressing and sintering. Binder treatment of amount of porosity, interconnected porosity is more effective
the powder mixture prior to pressing and sintering is a new in reducing macroscopic ductility.[7,9]
and effective technique to increase the compressibility of Porosity also significantly affects fatigue performance,
prealloyed powders, eliminating the diffusion-alloying step, although the role of porosity in fatigue is somewhat different
while still minimizing segregation.[1–5] In the binder-treated than that in tension. In many investigations,[4–6,8–10] crack
process, a polymeric binder mechanically bonds the alloying initiation occurred at pores or pore clusters located at or
additions to the larger iron particles (Figure 1). The enhance- near the surface of the specimen. Holmes and Queeney[8]
ment in compressibility of the powder mixture is accompa- proposed that the relatively high stress concentration at
nied by a smaller path for diffusion during sintering, which pores, particularly surface pores, is responsible for localized
results in enhanced densification of the sintered product. slip leading to crack initiation. In general, an angular pore
Burnout of the binder is accomplished either by a debinding creates a higher stress concentration and stress-intensity fac-
step (at an intermediate temperature below the sintering tor than a round pore.[10] Christian and German[9] showed
temperature) or by heating at a relatively slow rate until the that total porosity, pore size, pore shape, and pore separation
sintering temperature is reached. Other advantages of binder are important factors that control the fatigue behavior of
processing include faster and more consistent flow into the
powder-metallurgy materials. Pores have also been proposed
die cavity and an increased green strength, which results
to act as linkage sites for crack propagation through
interpore ligaments.[4,5,10]
In this study, we have examined the fatigue behavior of
S.J. POLASIK, Undergraduate Research Assistant, J.J. WILLIAMS, a binder-treated Fe-0.85Mo-Cu-Ni alloy. Comparisons are
Postdoctoral Research Fellow, and N. CHAWLA, Assistant Professor, are
with the Department of Chemical and Materials Engineering, Arizona State
made with the fatigue resistance of diffusion-alloyed materi-
University, Tempe, AZ 85287-6006. als of similar composition. Recent limited data indicate that
Manuscript submitted May 10, 2001. binder-treated materials have similar tensile and fatigue

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 33A, JANUARY 2002—73


pore perimeter. For the shape-form factor, a value of “1”
denotes a perfectly round pore, and values that approach
zero denote increasingly irregular pores.
Uniform-diameter cylindrical specimens for tensile and
axial fatigue were machined from the sintered rectangular
blanks. The specimens had a diameter of 5 mm and a gage
length of 15 mm.[12] All testing was conducted on a servohy-
draulic frame. Tensile tests were conducted in strain control
at a constant strain rate of 10⫺3/s, while fatigue testing was
conducted in load control, with an R-ratio (␴min/␴max) of ⫺1
and a frequency of 40 Hz (40 cycles per second). Fatigue
specimens were hand polished, using diamond paste, to a 1
␮m finish. Surface replication was conducted by interrupting
the fatigue test and placing the sample under a relatively
small tensile load (⬃0.25 ␴max) to avoid closure of any
fatigue cracks during the replicating procedure. The entire
Fig. 1—SEM micrograph of alloying additions bonded to iron particles by gage section was then bathed with acetone and subsequently
the binder.
covered with cellulose acetate tape. After the solvent (ace-
tone) evaporated and the replicating tape dried, the replicas
were flattened onto double-stick tape and placed between
properties to those of diffusion-alloyed materials.[4,5] A thor- two microscope slides. Using an optical microscope, digital
ough examination of crack initiation and propagation in micrographs were taken of the entire replicated surface at
these materials, particularly the behavior of short fatigue various fatigue cycles. The largest crack present on the last
cracks, was conducted. Short cracks range in size from a replica prior to failure was identified on prior replicas until
fraction of a millimeter to several millimeters and have been the point of crack initiation. The resolution of the surface
shown to propagate at much faster rates than long fatigue replicas allowed identification of cracks greater than about
cracks under the same driving force.[11] Additionally, short 15 ␮m. Details of the surface-replication technique may be
cracks propagate at stress intensities well below the long- obtained elsewhere.[13,14]
crack stress-intensity factor, (⌬Kth). It will be shown that
short fatigue cracks are important in causing fatigue damage
and failure in these materials. By using surface replication, a III. RESULTS AND DISCUSSION
detailed and in-situ understanding of fatigue-crack initiation
and propagation processes in these materials was obtained. A. Microstructure
Since the preferred initiation site for fatigue cracks is fre- During sintering, multiple phases are formed due to local-
quently the specimen surface, surface replication is an ideal ized variations in composition. The microstructure of the Fe-
means to study crack initiation and growth of short cracks. 0.85Mo-1.5Cu-1.75Ni-0.6 graphite alloy, in the as-sintered
The heterogeneous nature of the microstructure in these condition, is shown in Figure 2. The residual porosity from
materials results in significant crack deflection, which has sintering may be characterized as either primary or second-
also been modeled to adequately characterize the effective ary (Figure 2(a)). Primary porosity consists of larger pores
stress intensity for fatigue-crack growth. which result from geometric packing of the particles or from
binder burnout. Secondary porosity is typically smaller in
nature and may be attributed to residual porosity from liquid-
II. MATERIALS AND EXPERIMENTAL
phase formation and diffusion of alloying additions, such as
PROCEDURE
copper. A heterogeneous microstructure consisting of mar-
Powder mixtures of an Fe-0.85Mo prealloy, 1.5 pct Cu, tensite, nickel-rich ferrite, and divorced pearlite (termed
1.75 pct Ni, and 0.6 pct graphite were binder treated using “divorced” since it is not lamellar, as in conventional pearl-
a proprietary process developed by Hoeganaes Corp.[1,2,3] ite) was observed, as shown in Figure 2(b), and nickel-rich
All powders were pressed into rectangular blanks to a green regions surrounding pores were also observed (Figure 2(c)).
density of 7.0 g/cm3 and sintered at 1120 ⬚C for 30 minutes Results of the pore-size and shape analyses are shown in
in a 90 pct N2-10 pct H2 reducing atmosphere. The as- Figure 3. The pore-size and shape analyses show that the
sintered microstructure of the powder-metallurgy alloys was microstructure consists of pores with sizes below 300 ␮m2
examined by etching with a 2 pct Nital solution. Digital and relatively irregular shapes (shape factors are between
image-analysis techniques were used to determine the pore 0.4 and 0.7). The largest fraction of pores has an area of
morphology (pore size and shape distribution). In order to about 75 ␮m2. The irregularity of the pores suggests a greater
characterize the pore structure, both the pore size and shape amount of local stress concentration and, subsequently, a
distribution were determined from optical micrographs of larger amount of potential fatigue-crack initiation sites than
cross-sections of the samples. The pore size was estimated by in a similar material with perfectly spherical porosity.
measuring the pore area, while pore shape was characterized
using a shape-form factor (F ):
B. Tensile Behavior
4␲A
F⫽ 2 [1] The tensile stress-strain behavior of the Fe-0.85Mo-
P
1.5Cu-1.75Ni-0.6 graphite alloy is shown in Figure 4. Table
where A is the measured pore area and P is the measured I summarizes the tensile properties of this alloy. Porosity

74—VOLUME 33A, JANUARY 2002 METALLURGICAL AND MATERIALS TRANSACTIONS A


(a)
(a)

(b)
(b)
Fig. 3—Distributions in (a) pore size and (b) pore shape in powder-metal-
lurgy alloy.

(c) Fig. 4—Tensile stress-strain behavior of the binder-treated alloy.


Fig. 2—Microstructure of powder-metallurgy material tested in as-sintered
condition: (a) primary and secondary porosity, (b) heterogeneous nature
of the microstructure, and (c) nickel-rich areas from incomplete sintering
of Ni.
clearly decreases the elastic modulus over that of wrought-
alloy steel. Macroscopic ductility due to strain localization
of the sintered neck regions is also significantly lower than

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 33A, JANUARY 2002—75


Table I. Tensile Properties of Fe-0.85Mo Binder-Treated
Alloy
0.2 Pct
Proportional Offset Ultimate
Elastic Limit Yield Tensile Strain-to-
Modulus Stress Stress Strength Failure
(GPa) (MPa) (MPa) (MPa) (Pct)
121 233 526 774 1.83

that of most fully dense wrought alloys. The locally hard


phases shown in Figure 2 may also contribute to a higher
work-hardening rate and lower ductility than that exhibited
in wrought materials. The tensile strength of these materials
has also been shown to increase with Mo content.[5] An
increase in fraction of carbides with increasing Mo content
may account for this behavior. It is important to note that
the onset of microplasticity, given by the proportional-limit
stress (taken here as the point at which the stress deviates (a)
1 pct from linearity) or the stress at which the stress-strain
curve begins to deviate from linearity, takes place at a stress
much lower than the 0.2 pct offset yield stress. The
microplasticity can be attributed to the localized stress con-
centrations and yielding that takes place in sintered necks
between very closely spaced pores.[15]
Fractographic examination of tensile surfaces indicated
localized microvoid growth and coalescence at sintered
necks (Figure 5(a)). Regions where particle contact was not
present, and, thus, sintering did not take place, were also
apparent. Cleavage fracture, in what are thought to be pearl-
itic colonies, was also observed. It should be noted that
damage was more predominant in the form of localized
ductile rupture of sintered necks. Given the lower strength
and localized higher ductility of these necks, it is plausible
that it is the dominant mode of damage during tensile
deformation.

C. Fatigue Behavior
The stress vs cycles behavior of the Fe-Mo steel is shown (b)
in Figure 6 (fatigue run-out was taken as 107 cycles). The Fig. 5—SEM micrographs of tensile fracture: (a) ductile rupture in localized
results of the present investigation are compared to the sinter bonds and (b) cleavage in large pearlitic grains.
results of Chawla et al.,[5] who studied an Fe-Mo alloy with
0.5 pct Mo (using similar specimen dimensions and surface-
preparation techniques to this study). A distinct increase in and Eo is the elastic modulus of the as-sintered material.
fatigue strength with increasing Mo content is observed, Figure 7(a) shows the evolution of the damage parameter
which, as in the case of tensile behavior, may be attributed for specimens in the low-cycle fatigue (LCF) and high-cycle
to a higher fraction of carbides. fatigue (HCF) regime. The number of cycles to the onset of
Quantification of stress-strain behavior can shed some crack initiation of the LCF specimen, from surface-replica-
insight into the fatigue mechanisms in these alloys, particu- tion measurements, is denoted by Ni . The onset of crack
larly relating to cyclic hardening or cyclic softening. The initiation correlates fairly well with the point at which DE
width of the stress-strain hysteresis loop at zero stress was shows a noticeable increase. As expected, the extent of dam-
used to determine the cyclic plastic strain amplitude, and age was much lower in the HCF regime.
the slope of the loop, i.e., the secant modulus, was used to The plastic-strain amplitude seems to be much more sensi-
estimate fatigue damage. In this analysis, a useful parameter tive to cyclic hardening and cyclic softening (Figure 7(b))
to quantify the evolution of damage is the damage parame- than the damage parameter. In the LCF regime, the plastic-
ter[16] (DE), defined as strain amplitude decreases relatively early in the fatigue life,
E indicating cyclic hardening, followed by a gradual increase,
DE ⫽ 1 ⫺ [2] i.e., cyclic softening, until failure takes place. Cyclic soften-
Eo
ing correlates well with the onset of crack initiation and an
where E is the secant modulus at any given number of cycles increase in damage. Lindstead and Karlsson[17] conducted

76—VOLUME 33A, JANUARY 2002 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 8—Ratio of cycles to initiation to cycles to failure. Note that in the
Fig. 6—Comparison of binder-treated alloys with 0.5 pct Mo[5] and 0.85 LCF regime a significant fraction of life is spent in crack propagation,
pct Mo. An increase in Mo results in an increase in fatigue life. while in the HCF regime, most of the fatigue life is spent in crack initiation.

strain-controlled LCF fatigue tests on powder-metallurgy


stainless steels and observed that, for a given strain, the peak
stress in compression was higher than that in tension. This
was attributed to the inability of pores to transfer load in
tension, while in compression, collapse of the pores takes
place, and a higher degree of load transfer to the sintered
regions can take place. The early onset of cyclic hardening
was attributed to the high stress concentrations and, thus,
high localized plasticity. The localized stresses cannot be
relieved easily because of the surrounding interconnected
porosity, which causes the cyclic hardening. At lower strain
amplitudes, the size of the plastic zone is sufficiently small
to cause a gradual increase in the hardening process. In the
HCF regime, both the damage parameter and plastic-strain
amplitude remained relatively unchanged since, in this
regime, a large fraction of the total strain is elastic in nature.
(a) As mentioned previously, one of the most important con-
trolling factors in the fatigue resistance of powder-metal-
lurgy materials is porosity. Fatigue cracks tend to initiate
near pores or pore clusters because of the higher localized
stress intensity associated with these defects. On the other
hand, microstructural heterogeneities surrounding the pores,
such as the locally stronger Ni-rich areas shown in Figure
2(c), may contribute to a decrease in the stress intensity at
the pores and increase the number of cycles to fatigue-crack
initiation. Typically, cracks initiate at pores located at or
near the surface of the specimen, because the stress intensity
is higher there than at a pore in the interior.[6] The number
of cycles required for crack initiation (Ni) was determined
from surface replication, and the fraction of the life of the
specimen spent initiating a crack (Ni/Nf) was plotted as a
function of cycles to failure (Nf) in Figure 8. This approach
provides a quantitative estimate of the fraction of fatigue life
spent in initiation and propagation. From this relationship, it
can be seen that in the LCF regime, cracks initiate very
(b) early (15 pct) and the majority of the fatigue life is spent
Fig. 7—(a) Damage parameter and (b) plastic strain data from stress-strain in crack propagation, while in HCF, about 80 pct of the
hysteresis measurements during fatigue. The specimens tested in the HCF fatigue life is spent initiating the crack. Interestingly, the
regime (210 MPa) exhibit cyclic hardening and gradual cyclic softening,
while specimens tested in the LCF regime (300 MPa) remain relatively
same relationship between Ni/Nf and Nf has been observed
unchanged. Ni, the number of cycles to crack initiation determined from in other systems where fatigue-crack initiation takes place
surface replication, for the LCF specimen is indicated. at exogeneous surface defects.[18–21] One can rationalize this

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 33A, JANUARY 2002—77


Fig. 9—Fatigue crack propagation in Fe-Mo alloy. Note the fatigue crack gradually linking between pores and following a torturous path, presumably d ue
to localized microstructural inhomogeneities (arrows indicate loading axis).

behavior by an incubation concept to describe the fatigue-


crack initiation process in these materials. At intermediate
to low stresses, crack initiation did not occur immediately
upon cycling. Rather, a period of cyclic loading, i.e., an
incubation period, was required to create a localized defor-
mation that resulted in cracking initiation at a pore or cluster
of pores. The number of cycles required to initiate the fatigue
crack appears to be a function of the maximum applied
stress, with lower stresses requiring longer incubation times
for a given pore size and shape. This is to be expected,
because the contribution to the total strain amplitude is pri-
marily elastic in nature. Defect size and shape also play a
role, as larger and more irregular defects are more susceptible
to crack initiation for a given applied stress and will naturally
lead to lower fatigue strength. (a)
After a fatigue crack has initiated at surface or subsurface
pores, it tends to propagate and grow through the interpore
ligaments, using pores as linkage sites. The evolution of
fatigue-crack growth, from surface-replication measure-
ments, is shown in Figure 9. While other investigators have
postulated crack linkage as a possible mechanism,[22,23] sur-
face replication allows in-situ monitoring of the crack initia-
tion and growth process and confirms the linkage of smaller
cracks to form a final crack that causes failure of the material.
Fatigue fractography provided additional insight into
fatigue damage. A scanning electron microscope micrograph
of a fatigue-crack initiation site is shown in Figure 10(a).
The intermittent nature of fatigue-crack growth during each
cycle can be observed through the existence of localized
fatigue striations in various fractured sinter bonds (Figure
10(b)). Striations in conventional steels and alloys are rela-
tively homogeneous and of a single orientation. The forma-
tion of striations has been attributed to plastic strains at the (b)
crack tip, which cause localized slip on planes of maximum Fig. 10—(a) Fatigue crack initiating pore on specimen surface and (b)
shear.[24] The crack front undergoes repetitive blunting and fracture surface after fatigue showing fatigue striations and localized duc-
sharpening during propagation during cyclic loading. Unlike tile rupture.

78—VOLUME 33A, JANUARY 2002 METALLURGICAL AND MATERIALS TRANSACTIONS A


striations in wrought materials, however, the striations in
porous sintered materials seem to be highly localized. This
may be due to the multiple sites of favorable orientation and
size, with respect to the loading axis, that undergo blunting
and sharpening and give the step-like appearance of the
striations. Similar localized fatigue striations were also
observed by Rodzinak and Slesar.[25]
Ductile rupture is also present on the fatigue fracture
surface, as seen in Figure 10(b). The ductile failure regions
are highly localized in sinter bonds and seem to develop
from microvoid nucleation and coalescence. The localized
necking and microvoid growth, while present in the crack-
initiation stage, are predominately found in the later stages
of fatigue. Small spherical inclusions, which likely contrib-
uted to microvoid formation, were found at the bottom of
some microvoids (Figure 10(b)). The composition of the
inclusions was identified as MnS, although Fe oxides were (a)
also found. This is consistent with other authors’ observa-
tions of precipitates and oxides at the periphery of the Fe
particles and at previous particle boundaries.[6] Cleavage
fracture was also observed on the fracture surface. Since
this type of failure was seen in the tensile tests and, therefore,
in fast fracture, cleavage failure may be associated with the
fracture-surface region corresponding to the final, fast crack
propagation leading to failure. Generally, when this cleavage
fracture occurred, it was found to be in large pearlitic grains,
similar to what is shown in Figure 5(b).

D. Fatigue-Crack Growth Rate and Modeling of Crack


Deflection
The fatigue-crack growth rate of short cracks was calcu-
lated from a surface-replication measurement of fatigue
cracks. We use the approach developed by Raju and New-
man[26] for a surface crack in a cylinder (c) to calculate the (b)
stress intensity (K): Fig. 11—da/dN vs stress intensity ranges (⌬K ) as calculated using Raju
and Newman analysis[26] for (a) LCF and (b) HCF.
⌬K ⫽ F⌬␴ 冪␲ Qa [3]

where a is the crack depth, Q is the shape factor for an


ellipse (defined as 1 ⫹ 1.464(a/c)1.65 for a/c ⱕ 1), ⌬␴ is the was taken to be perpendicular to the loading axis). The
tensile component of the fatigue stress, and F is a boundary influence of crack deflection on the stress-intensity factor
correction-size factor which accounts for crack size, crack in the crack was also calculated and is discussed later in
shape, and the ratio of the crack size to specimen diameter. this section. The da/dN data in Figure 10 show that while
Note that only the tensile component of the fatigue-stress several cracks initiate and link together, some fatigue cracks
range is used to compute ⌬K, since the crack growth can show a deceleration in crack growth rate, followed by arrest.
be attributed primarily to the tensile component of the fatigue Conversely, some cracks are arrested, but, after a dormant
stress. Raju and Newman[26] calculated the values for F using period, continue to grow. Crack arrest or deceleration may
the finite-element method. The value for the a/c ratio used be attributed to microstructural barriers such as grain bound-
for the calculations in this study was determined from the aries, locally hard regions (such as Ni-rich regions), interac-
fracture surface and was measured to be about 0.58. This tion with other cracks, or even porosity. Short-crack growth
agrees well with the work of Lindstedt et al.,[10] who meas- also seems to be dependent on fatigue stress, with higher
ured an a/c ratio of about 0.67 in a powder-metallurgy stain- crack growth rates at higher stresses, for nominally similar
less steel. stress-intensity ranges. In addition, many of these cracks
Figure 11 shows the crack growth rate vs stress-intensity measured through surface replication are small cracks, and
factor in the LCF and HCF regimes. The stress intensity for the plastic zone is of comparable size to the crack size,
propagation of the short crack is significantly lower than requiring higher stresses for propagation. Crack deflection
the Kth value obtained from long-crack fatigue growth of may also take place due to microstructural barriers in the
these materials. This reinforces the importance of quantify- material. Nickel-rich regions have been proposed as possible
ing short-fatigue-crack nucleation and growth in these mate- microstructural obstacles for crack growth.[27,28]
rials. The value for the crack length (2c) was taken as the Since crack deflection is significant in these materials, it
horizontal projection of the crack (the length of the crack is important to address the effect of deflection on the overall

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 33A, JANUARY 2002—79


Fig. 12—Crack deflection profile used in the model proposed by
Suresh.[29,30]

stress intensity of the fatigue crack. We use the model devel-


oped by Suresh[29,30] to introduce an approximate measure
of deflection in the crack growth rate. The model assumes
a deflected crack profile, shown in Figure 12.
The crack travels a length of S in mode I, then deflects
at an angle of ␪ over a length of D. The deflection segment
D ⫹ S is repeated, assuming that a given segment is much
smaller than the total length of the crack and that the degree (a)
of deflection in two adjoining segments takes place in oppo-
site directions. The contribution from deflection is estimated
by the average angle and by D/(D ⫹ S), taken over the entire
crack profile. The stress-intensity range of the deflected
portion of the crack is ⌬kdef, while that of the straight segment
is ⌬KI . The effective driving force in the total segment (⌬KI)
is given by the weighted average of the effective stress-
intensity factor across the deflected length D and that across
the straight length S:
D⌬kdef ⫹ S⌬KI
⌬KI ⫽ [4]
D⫹S
where ⌬kdef ⬇ ⌬KI ((cos3 (␪ /2))2 ⫹ (sin (␪ /2) cos2
(␪ /2)2)1/2. Simplifying Eq. [5], we can obtain an expression
for ⌬KI in terms of ⌬KI:


冢冣
(b)
⌬KI ⬇ ⌬KI
冢Dcos2

D⫹S
2
⫹S
冣 [5] Fig. 13—da/dN curves illustrating the higher stress intensity required to
propagate a crack at a given growth rate when deflection occurs for (a)
LCF and (b) HCF.
In a straight crack, the effective stress intensity (⌬keff) is
equal to the applied stress intensity (⌬KI,app). In order to
propagate a deflected crack at the same rate as the straight
crack, ⌬keff ⫽ ⌬KI, so ⌬KI,app may be written as a longer fraction of the total life in LCF (in which cracks
initiate at approximately 10 pct) than in HCF (in which
␪ ⫺1 initiation occurs at around 70 pct of the life span), making
⌬KI,app ⬇ ⌬keff
冢 Dcos2 冢冣
2
⫹S
冣 [6]
the crack-deflection model most useful in crack growth in
the LCF regime.
D⫹S
Measurements of S, D, and ␪ were taken from digital optical IV. CONCLUSIONS
micrographs of the surface replica just prior to failure. Figure
The following conclusions can be made concerning the
13 shows the effective stress-intensity increase required to
fatigue behavior of binder-treated Fe-0.85Mo-1.5Cu-
propagate a deflected crack at the same crack growth rate
1.75Ni-0.6 graphite alloys.
as a straight crack of equal length (measured experimentally
from surface replication). For large interpore spacings, the 1. The microstructure of the alloy consisted of a heteroge-
heterogeneous microstructure in these materials seems to neous microstructure with areas of divorced pearlite, mar-
effectively decrease the crack propagation rate. However, tensite, and nickel-rich ferrite. Two types of porosity
with a higher porosity content and smaller interpore spacing, were characterized: (1) primary porosity, due to particle
the microstructural effects on deflection and subsequent packing and binder burnout, and (2) secondary porosity
crack propagation will most likely be outweighed by the formed due to the copper particles forming a liquid phase
porosity contribution to crack deflection. The stress and and diffusing into Fe particles during sintering.
strain concentration around pores, for instance, will most 2. An increase in Mo content increases the fatigue resistance
likely influence the crack path more than microstructural of the alloy, presumably due to an increase in the fraction
features within an interpore ligament. Cracks propagate for of carbides.

80—VOLUME 33A, JANUARY 2002 METALLURGICAL AND MATERIALS TRANSACTIONS A


3. During fatigue, an initial period of cyclic hardening was 4. N. Chawla, S. Polasik, K.S. Narasimhan, M. Koopman, and K.K.
Chawla: Int. J. Powder Metal., 2001, vol. 37, pp. 49-57.
followed by a period of gradual cyclic softening. This 5. N. Chawla, T.F. Murphy, K.S. Narasimhan, M. Koopman, and K.K.
behavior was most pronounced in the LCF regime. Addi- Chawla: Mater. Sci. Eng. A, 2001, vol. A308, pp. 180-88.
tionally, the incubation period required for crack initiation 6. A. Hadrboletz and B. Weiss: Int. Mater. Rev., 1997, vol. 42, pp.
correlated very well with the onset of cyclic softening. 1-44.
4. The fraction of the life required to initiate a fatigue crack, 7. H. Danninger, D. Spoljaric, and B. Weiss: Int. J. Powder Metall.,
1997, vol. 33, pp. 43-53.
Ni/Nf , increased with a decrease in fatigue-stress ampli- 8. J. Holmes and R.A. Queeney: Powder Metall., 1985, vol. 28, pp.
tude. Thus, crack initiation occurred early in the life of 231-35.
LCF specimens, and a majority of the fatigue life was 9. K.D. Christian and R.M. German: Int. J. Powder Metall., 1995, vol.
spent in the crack-propagation mechanism. In contrast, 31, pp. 51-61.
10. U. Lindstedt, B., Karlsson, and R. Masini: Int. J. Powder Metall.,
since crack initiation occupied a much higher fraction of
1997, vol. 33, pp. 49-61.
fatigue life in the HCF regime, defect size has a higher 11. S. Suresh: Fatigue of Materials, 2nd ed., Cambridge University Press,
impact on the fatigue life, so pore size was the more Cambridge, United Kingdom, 1998, p. 541.
dominant factor in this regime. 12. N. Chawla, C. Andres, J.W. Jones, and J.E. Allison: Metall. Mater.
5. Short fatigue cracks initiated at surface pores or pore Trans. A, 1998, vol. 29A, pp. 2843-54.
13. M.H. Swain: Small Crack Test Methods, ASTM STP 1149, ASTM,
clusters and propagated at faster rates and lower stress Philadelphia, PA, 1992, pp. 34-56.
intensities than long fatigue cracks. Linkage of these 14. M. Caton, J.W. Jones, J.M. Boileau, and J.E. Allison: Metall. Mater.
small fatigue cracks to form the final critical-size crack Trans. A, 1999, vol. 30A, pp. 3055-68.
caused failure, which was characterized by in-situ surface 15. J.A. Lund: Int. J. Powder Metall. Powder Technol., 1984, vol. 20, pp.
replication. Localized microstructural obstacles resulted 141-48.
16. Z.R. Xu, K.K. Chawla, A. Wolfenden, A. Neuman, G.M. Liggett, and
in significantly torturous fatigue growth. N. Chawla: Mater. Sci. Eng., 1995, vol. 203A, pp. 75-80.
6. Application of a crack-deflection model provided a quan- 17. U. Lindstedt and B. Karlsson: Advances in Powder Metallurgy &
titative estimate to the role of crack deflection. It was Particulate Materials, compiled by T.M. Cadle and K.S. Narasimhan,
demonstrated that the effective stress intensity at the crack MPIF, Princeton, NJ, 1996, vol. 5, pp. 17-35.
18. K.V. Sudhakar: Int. J. Fatigue, 2000, vol. 22, pp. 729-34.
tip of a deflected crack must be increased significantly 19. D.A. Lukasak and D.A. Koss: Composites, 1993, vol. 24, p. 262.
to achieve similar crack growth rates to those observed 20. N. Chawla, L.C. Davis, C. Andres, J.E. Allison, and J.W. Jones: Metall.
for a straight crack of equal length. Mater. Trans. A., 2000, vol. 31A, pp. 951-57.
21. S.M. McGuire and M.E. Fine: Metall. Mater. Trans. A, 1996, vol.
27A, pp. 1267-71.
ACKNOWLEDGMENTS 22. H. Drar and A. Bergmark: Fatigue Fract. Eng. Mater. Struct., 1997,
vol. 20, pp. 1319-30.
The authors gratefully acknowledge Dr. K.S. Narasimhan, 23. H. Drar, Mater. Characterization, 2000, vol. 45, pp. 211-20.
24. C. Laird: Fatigue Crack Propagation, Special Technical Publication
Hoeganaes Corp., for supplying the materials and for the 415, ASTM, Philadelphia, PA, pp. 131-68.
financial support for this work. 25. D. Rodzinak and M. Slesar: Powder Metall. Int., 1980, vol. 12, pp.
127-30.
26. I.S. Raju and J.C. Newman: Fracture Mechanics, ASTM STP
REFERENCES 905, J.H. Underwood, R. Chait, C.W. Smith, D.P. Wilhem, W.A.
Andrews, and J.C. Newman, eds. ASTM, Philadelphia, PA, 1986, pp.
1. N. Chawla, G. Fillari, and K.S. Narasimhan: in Powder Materials: 789-805.
Current Research and Industrial Practices, F.D.S. Marquis, ed., TMS, 27. S. Carabajar, C. Verdu, A. Hamel, and R. Fougeres: Mater. Sci. Eng.,
Warrendale, PA, 1999, p. 247. 1998, vol. A257, pp. 225-34.
2. F.J. Semel: Advances in Powder Metallurgy and Particulate Materials, 28. T.M. Cimino, A.H. Graham, and T.F. Murphy: Advances in Powder
Metal Powder Industries Federation, Princeton, NJ, 1989, p. 9. Metallurgy and Particulate Materials, Metal Powder Industries Feder-
3. S.H. Luk and J.A. Hamill, Jr.: Advances in Powder Metallurgy and ation, Princeton, NJ, 1998.
Particulate Materials, Metal Powder Industries Federation, Princeton, 29. S. Suresh: Metall. Trans. A, 1983, vol. 14A, pp. 2375-85.
NJ, 1993, p. 153. 30. S. Suresh: Metall. Trans. A, 1985, vol. 16A, p. 249.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 33A, JANUARY 2002—81

You might also like